Sunteți pe pagina 1din 1514

PREFACE

These two parts bring together a number of authoritative, state-of-the-art reviews and
contributions, written by well recognized experts in the field of"flow and rheology of
non-Newtonian fluids." Knowledge of non-Newtonian behavior is of vital importance to
a variety of manufacturing processes including, for example, mixing, shear-thickening,
fibre spinning, coating, and molding.

This work covers areas such as bio- and food- rheology, electro-rheological fluids,
polymers, flow in porous media, and suspensions. Complex and industrial flow situations
are dealt with via analytical, as well as, numerical methods. In Chapter 1, a critical
account of advances made in the area of flow-induced interactions in circulation is
presented. Chapters 2 & 3 deal with shear-thickening in biopolymeric systems, and with
the rheology of food emulsions, respectively. The next six chapters are on complex flows,
in particular, Chapter 4 discusses worm-like micellar surfactant solutions. Chapter 5
covers time periodic flows. Chapter 6 communicates on secondary flows in tubes of
arbitrary shape. Chapter 7 relates effects of non-Newtonian fluids on cavitation. Chapter
8 discusses viscoelastic Taylor-Vortex flow. Chapter 9 deals with non-Newtonian mixing.
This is followed by two chapters on computational methods relevant to homogeneous
viscoelastic fluids at the macro-level. The next major section is on constitutive equations
and viscoelastic fluids. Chapter 12 discusses recent advances in transient network theory.
Chapter 13 deals with theories based on fractional derivatives and Chapter 14 involves
kinetic theory. Chapters 15 and 16 put forward new concepts approaching the constitutive
structure of polymeric melts. The next chapter communicates the theory of flow of
smectic liquid crystals. Part A ends with an overview of extensional flows.

Volume B starts with a section on electro-rheological fluids. The first two chapters in the
section summarize the constitutive theories for electro-rheological fluids from the
continuum and molecular points of view. Chapter 21 relates a comprehensive approach
to the constitutive structure of electromagnetic fluids, and the following two chapters deal
with the properties of electro-rheological fluids. The next section covers some industrial
flows related to drag reduction, and paper coating. Polymer processing and the related
rheology are discussed in Chapters 26-29. In particular, the rheology of long
discontinuous iber thermoplastic composites, thermo-mechanical modelling of polymer
processing, injection molding and flow of melts in channels with moving boundaries are
covered in Chapters 26-29 respectively. Free surface viscoelastic and liquid crystalline
polymer fibers and jets, and numerical sinmlation of melt spinning of polyethylene fibers
are the subject of Chapters )0 and 31, respectively.
vi

Section 9 contains two chapters dealing with foam flow and non-Newtonian flow in
porous media. Section 10 discusses four chapters on various aspects of
suspension.Chapter 34 reviews and puts forth new ideas on the fluid dynamics of fine
suspensions. Chapter 35 deals with concentrated suspensions. This section ends with a
discussion on fiber suspensions and fluidized beds.

The last section of Part B contains a discussion on transport-phenomena involving heat


and mass transfer in rheologically complex systems and an account of a new one-
dimensional model for viscoelastic diffusion in polymers.

With the publication of this work we hope to update and complement earlier work in a
diverse range of topics. These two volumes should be of interest to all those engaged in
basic, as well as applied research.

The information presented herein is equally valuable for practising engineers who are
constantly dealing with complex situations involving non-Newtonian materials. The
contents of the two volumes are accessible to those with a background in engineering
and/or pure sciences. We would like to take this opportunity to thank the contributors
who, despite their busy schedules, kindly agreed to participate.

Dennis A. Siginer
New Jersey Institute of Technology
Newark, NJ, USA

Daniel DeKee
Tulane University
New Orleans, LA, USA

Raj P. Chhabra
Indian Institute of Technology
Kanpur, UP, India
vii

LIST OF CONTRIBUTORS

Advani, Suresh G. Cairncross, R. A.


University of Delaware Department of Chemical Engineering
Department of Mechanical Engineering Drexel University
Spencer Laboratory Philadelphia, Pennsylvania 19104, USA
Newark, Delaware 19716, USA

Agassant, J. F. Carreau, Pierre J.


Centre de Mise en Forme des Mat6riaux Center for Applied Research on Polymers
Ecole des Mines de Paris CRASP
URA CNRS 1374 Department of Chemcical Engineering
BP 207 Ecole Polytechnique, Montreal
06904, Sophia Antipolis, FRANCE QC H3C 3A7, CANADA

Bagley, Edward B. Chhabra, R. P.


756 S. Columbus Department of Chemical Engineering
Morton, Illinois 61550-2428, USA Indian Institute of Technology
Kanpur, INDIA 208016
Bakhtiyarov, Sayavur I.
Space Power Institute Co, Albert
231 Leach Center Department of Chemical Engineering
Auburn University University of Maine
Aubum, Alabama 36849-5320, USA Orono, Maine 04469-5737, USA

Bechtel, Stephen E. Conrad, Hans


Department of Aerospace Engineering Department of Materials Science and Engineering
Applied Mechanics and Aviation North Carolina State University
The Ohio State University Raleigh, North Carolina 27695, USA
Columbus, Ohio 432 I0, USA

Blumen, A. Couniot, A.
Theoretical Polymer Physics Siemens-Nixdorf Information Systems S.A.
Freiburg University LoB "Major Projects"
Rheinstr. 12, 79104 Freiburg, GERMANY Chaussee de Charleroi 116_
B- 1060 Brussels, BELGIUM

Bousfield, Douglas W. Coupez, T.


Department of Chemical Engineering Centre de Mise en Forme des Materiaux
University of Maine Ecole des Mines de Paris
5737 Jennes Hall URA CNRS 1374
Orono, ME 04469-5737 USA BP 207, 06904, Sophia Antipolis, FRANCE

Brito-De La Fuente, Edmundo Creasy, Terry


Food Science and Biotechnology Department University of Southern California
Chemistry Faculy "E" Center for Composite Materials
National Autonomous University of Mexico VHE 602 MC0241
UNAM, 04510 Mexico, D.F., MEXICO Los Angeles, California 90089-0241, USA

Brunn, Peter O. De Kee, D.


Universitat Erlangen-Nurnberg Department of Chemical Engineering
Lehrstuhl fur Stromungsmechanik Tulane University
Caueerstr. 4 New Orleans, Louisiana 70118, USA
D-91058 Erlangen, GERMANY
Demay, Y.
Buyevich, Yuri A. Centre de Mise en Forme des Materiaux
Center for Risk Studies and Safety Ecole des Mines de Paris
University of California Santa Barbara URA CNRS 1374
6740 Cortona Dr. BP 207, 06904, Sophia Antipolis, FRANCE
Santa Barbara, Califomia 93117, USA
viii

Dintzis, Frederick R. Gallegos, C.


USDA Departamento de Ingenieria Quimica
ARS Universidad de Huelva
National Center for Agricultural Utilization Resarch Escuela Politecnica Superior
Peoria, Illinois 61604, USA La Rabida, 21819
Palos de la Ftra (Huelva), SPAIN
Dulikravich, George S.
Aerospace Engineering Department Goldsmith, Harry L.
233 Hammond Building Department of Medicine
The Pennsylvania State University The Montreal General Hospital
University Park, Pennsylvania 16802, USA 1650 Ave Cedar
Montreal, Quebec H3G 1A4, CANADA
Dunwoody, James
Department of Applied Mathematics & Theoretical Physics Hoyt, Jack W.
The Queen's University 4694 Lisann Street
Belfast BT7 INN NORTHERN IRELAND San Diego, Califomia 92117, USA

Dupret, F. lsayev, A. I.
CESAME Institute of Polymer Engineering
Unite de Mecanique Appliquee The University of Akron
Universite catholique de Louvain Akron, Ohio 44325-0301, USA
Avenue G. Lemaitre 4-6
B- 1348 Louvain-la-Neuve, BELGIUM Kanu, Rex C.
Department of Industry and Technology
Duming, Christopher J. Ball State University
Department of Chemical Engineering and Applied Chemistry Muncie, Indiana 47306, USA
Columbia University
New York, New York 10027, USA Khayat, Roger E.
Department of Mechanical & Materials Engineering
Fong, C. F. Chan Man The University of Western Ontario
Department of Chemical Engineering London, Ontario, CANADA N6A 5B9
Tulane University
New Orleans, Louisiana 70118, USA Kim, Kyoung Woo
Fiber Research Center
Forest, M. Gregory Sunkyong Industries
Department of Mathematics Su Won, 440-745, KOREA
University of North Carolina
Chapel Hill, North Carolina 27599-3250, USA Kim, Sang Yong
Department of Fiber and Polymer Science
Franco, J.M. College of Engineering
Departamento de Ingenieria Quimica Seoul National University
Universidad de Huelva San 56-1, Shinlim-Dong, Kwanak-Ku
Escuela Politecnica Superior Seoul 151-742, KOREA
La Rabida, 21819
Palos de la Ftra (Huelva), SPAIN Kornev, Konstantin G.
Institute for Problems in Mechanics
Friedrich, Chr. Russian Academy of Sciences
Freiburg Materials' Research Center 101 (1) Prospect Vemadskogo
Freiburg University Moscow 117526, RUSSIA
Stefan-Meier-Str.21
79104 Freiburg, GERMANY Kwon, Youngdon
Department of Textile Engineering
Fruman, Daniel H. Sung Kyun Kwan University
Groupe Phenomenes d'Interface Su Won, 440-746, KOREA
Ecole Nationale Superieure de Techniques Avancees
91761 Palaiseau Cedex - FRANCE
Lavoie, Paul Andre Prakash, J. Ravi
Center for Applied Research on Polymers Department of Chemical Engineering
CRASP Indian Institute of Technology
Department of Chemcical Engineering Madras, INDIA, 600 036
Ecole Polytechnique, Montreal
QC H3C 3A7, CANADA Rajagopal, K. R.
Texas A&M University
Leonov, Arkadii I. College Station, Texas, 77842-3014, USA
Department of Polymer Engineering
The University of Akron Rozhkov, Aleksey N.
Akron, OH 44325-0301, USA Institute for Problems in Mechanics
Russian Academy of Sciences
Leslie, Frank M. 101 (1) Prospect Vernadskogo
Mathematics Department Moscow 117526, RUSSIA
University of Strathclyde
Livingstone Tower Schiessel, H.
Richmond Street Theoretical Polymer Physics
Glasgow G 1 1XH, SCOTLAND Freiburg University
Rheinstr. 12, 79104 Freiburg, GERMANY
Letelier, Mario
Universidad de Santiago de Chile Shaw, Montgomery T.
Santiago Department of Chemical Engineering and Polymer Program
CHILE University of Connecticut
97 North Eagleville Road
Mal, O. U-136
CESAME Storrs, Connecticut 06269-3136, USA
Unite de Mecanique Appliquee
Universite catholique de Louvain Siginer, Dennis A.
Avenue G. Lemaitre 4-6 Department of Mechanical Engineering
B- 1348 Louvain-la-Neuve, BELGIUM New Jersey Institute of Technology
Newark, New Jersey 07102, USA
Neimark, Alexander V.
TRI/Princeton Steger, R.
601 Prospect Ave. Rheotest Medingen GmbH
P.O. Box 625 RodertalstraBe 1,
Princeton, New Jersey 08542-0625, USA D-01458 Medingen b. Dresden, GERMANY

Overfelt, R. A. Tang, P. H.
Space Power Institute Department of Chemical Engineering and Applied Chemistry
231 Leach Center Columbia University
Auburn University New York, New York 10027, USA
Auburn, Alabama 36849-5320, USA
Tanguy, Philippe A.
Padovan, J. Department of Chemical Engineering
Department of Mechanical Engineering Ecole Polytechnique Montreal
The University of Akron P.O. Box 6079
Akron, Ohio 44325-0301, USA Station Centre-ville
Montreal, H3C 3A7 CANADA
Petrie, Christopher J. S.
Department of Engineering Mathematics Tanner, R. I.
University of Newcastle upon Tyne Department of Mechanical and Mechatronic Engineering
Newcastle upon Tyne NE1 7RU, UNITED KINGDOM The University of Sydney
NSW 2006, AUSTRALIA
Phan-Thien, Nhan
Department of Mechanical and Mechatronic Engineering Tao, Rongjia
The University of Sydney Department of Physics
NSW 2006, AUSTRALIA Southern Illinois University at Carbondale
Carbondale, Illinois 62901, USA
Vanderschuren, L. Zhang, Y.
Shell Research S.A. Institute of Polymer Engineering
Avenue Jean Monnet 1, The University of Akron
B-1348 Louvain-la Neuve, BELGIUM Akron, Ohio 44325-0301, USA

Vergnes, B. Zhou, Hong


Centre de Mise en Forme des Materiaux Department of Mathematics
Ecole des Mines de Paris University of North Carolina
URA CNRS 1374 Chapel Hill, North Carolina 27599-3250, USA
BP 207, 06904, Sophia Antipolis, FRANCE
Zook, C.
Verhoyen, O. Institute of Polymer Engineering
CESAME The University of Akron
Unite de Mecanique Appliquee Akron, Ohio 44325-030 l, USA
Universite catholique de Louvain
Avenue G. Lemaitre 4-6
B- 1348 Louvain-la-Neuve, BELGIUM

Verleye, V.
TECHSPACE AERO
Route de Liers 121
B-4041 Milmort, BELGIUM

Vincent, M.
Centre de Mise en Forme des Materiaux
Ecole des Mines de Paris
URA CNRS 1374
BP 207, 06904, Sophia Antipolis, FRANCE

Vossoughi, Shapour
Department of Chemical and Petroleum Engineering
University of Kansas
4006 Learned Hall
Lawrence, Kansas 66045-2223, USA

Wang, Qi
Department of Mathematical Sciences
Indiana University-Purdue University at Indianapolis
Indianapolis, Indiana 46202, USA

Wu, C. W.
Research Institute of Engineering Mechanics
Dalian University of Technology
Dalian 116024
People's Republic of China

Yoo, Jung Yul


Department of Mechanical Engineering
College of Engineering
Seoul National University
Seoul 151-742, KOREA

Yziquel, F.
Center for Applied Research on Polymers
CRASP
Department of Chemcical Engineering
Ecole Polytechnique, Montreal QC H3C 3A7, CANADA
FLOW-INDUCED INTERACTIONS IN THE CIRCULATION

Harry L. Goldsmith

McGill University Medical Clinic, Montreal General Hospital,


Montreal, Quebec H3G 1A4, Canada

1. I N T R O D U C T I O N

Many years ago, it was suggested to the English physiologist L.E. Bayliss by
the eminent Cambridge authority on fluid mechanics, G.I. Taylor, that rather
than work on the highly complex rheology of blood, he would be better
spending his time studying the flow of marmalade through sewage pipes. That
advice did not deter Dr. Bayliss [1,2] and a host of physiologists, rheologists,
chemical and mechanical engineers from engaging in the problem, which had,
after all, an honourable parentage in the person of Jean Leonard Marie
Poiseuille, who published as early as 1839 [3,4], and in this century, the
Swedish pathologist, Robin F~thraeus [5,6]. Quite apart from the clinical value
of discovering how blood cells travel in the vessels of the circulation, there is,
for the rheologist, a fascination about this fluid, since it is able to flow, even at
hematocrits (packed red blood cell (RBC) volume concentration) well above
40%, with such ease compared to other suspensions and emulsions [7]. Thus, it
is possible to pack a column of mammalian red cells containing less than 2%
trapped plasma into a 2 mm diameter glass tube, and then suck it out quite
easily into a micropipette whose tip has a diameter of only 0.5 mm. By
comparison, oil-in-water emulsions reach the quasi-solid consistency of
margarine at much lower concentrations.
Clearly, in such a concentrated suspension, interactions between the blood
corpuscles and between the corpuscles and the vessel walls play an important
role in determining the mechanics of the motion. This chapter is concerned
with a description of these interactions. We begin by taking the reader through
a brief description of blood, seen as a dispersion of charged colloidal-size
particles in a Newtonian suspending medium, and the main flow regimes to
which it is subjected. To understand the cell and cell-wall interactions, one
must begin with the macroscopic, overall non-Newtonian rheology of blood,
and then to correlate these rheological properties with the flow properties of
the individual cells, studied first in isolation, then at increasing concentrations
as interactions become increasingly important, i.e. the microrheology of the
blood. This part of the chapter will mainly involve the erythrocytes or red
cells, since they occupy >99% by volume and -96% by number of the
particulate phase of blood, and thus effectively determine both the macro- and
microrheology of the suspension. However, the most recent advances in the
field have been made at the submicroscopic level, where the molecules and the
forces involved in cell-cell and cell-wall adhesion have been investigated, thus
entering the realm of what has been called molecular rheology. At this level,
the chapter focusses not on the red cells but on the platelets and leukocytes,
since it is they that are involved in key physiological and pathophysiological
events through their role in blood coagulation and a g g r e g a t i o n - the platelet,
and in i n f l a m m a t i o n - adhesion of leukocytes to the vessel wall and their
passage into the extravascular space. In fact, adhesion processes between cells
and between cells and the vessel wall are particularly critical in the circulation
for all three classes of cells. The biophysics and rheology of certain of these
adhesive reactions and the means of measuring forces of adhesion are dealt
with in Sections 4 and 5. Where possible, the phenomena described are
discussed in terms of fluid mechanical and colloid chemical theory.

1.1 B l o o d as a colloidal d i s p e r s i o n
Mammalian blood is a dispersion of three classes of negatively charged
particles (the cells or corpuscles; Table 1) in an aqueous solution (the plasma)
containing dissolved salts and proteins (Table 2). As pointed out above, from
the point of view of its rheological properties, blood is essentially a
concentrated suspension of red blood cells

1.1.1 Cell membrane


The biconcave red cell has a very thin lipid bilayer membrane (-10 nm)
containing proteins linked to a submembraneous skeleton of actin/spectrin.
The membrane encapsulates a 33% Newtonian salt solution of hemoglobin
having a viscosity ,--6x that of the plasma at 37~ The ease of deformation of
the red cell is mainly due to the properties of the membrane acting as a two-
dimensional incompressible material, which deforms at constant surface area
and constant volume [8], and when osmotically swollen does not increase its
surface area. The deformed cell remembers its undeformed shape and
recovers it within a fraction of a second upon removal of the forces of
deformation. By means of micropipette aspiration techniques, the viscoelastic
properties of the membrane have been extensively characterized. The elastic
deformation is described by three fundamental, independent deformations [9]3
elongation or shear of the membrane having a shear modulus of 6-9 0-
mN m -~, an isotropic expansion of the membrane surface without shear or
Table 1
Mean values of physical parameters of human blood cells
Cell Shape Major Axis Volume Number Volume
3 g1-1 Fraction
gm gm
Erythrocyte Biconcave 8.3 88 5.4x106 0.46 (m)
(Red Cell) disc 4.8xl 0 6 0.42 (f)
Thrombocyte Oblate 3.1 7.5 2.5x105 1.9x10 -3
(Platelet) spheroid
Leukocytes Ruffled 7.0x103 1.2x10 -3
(White Cells) spheres (total) (total)
Neutrophil 7.8 240 3.7x103 8.8x10 -4
Lymphocyte 6.0 115 2.5x103 2.9x10 -4

Table 2
Mean values of concentrations of ions and proteins in plasma
SALTS PROTEINS
Ion Concentration Name Molecular Concentration
meq./litre Weight gm/litre

Na + 150 Albumin 69,000 44.1

K+ 4 Fibrinogen 340,000 3.0


Ca++ 5
Mg++ 2

C1- 111 t~l-globulin 200,000 2.7


HCO3- 28 ~2-globulin 5.9
PO 4- 3
13-globulin 90,000 8.8
SO4-- 1 ),-globulin 160,000 14.2

bending, having an area expansion modulus of 450 mN m-', and a bending of


the membrane with a very low bending modulus of 1.8 + 0.2x10 -19 N m. The
resistance of the membrane to the rate of deformation is characterized by its
viscosity; the area viscosity has been found to have a value of 0.6-1.2x10 -3
mN s m -1 [9].

1.1.2 Charge and stability.


The negative charge on the red cell membrane is due to N-acetyl and N-
glycolneuraminic acid (sialic acids) residues attached via a gycosidic linkage to
other molecules. Measurements of the electrophoretic mobility of red cells in
buffers of physiological ionic strength = 0.145 yield values of the ~-potential
-15 mV [10]. When treated with the enzyme neuraminidase to remove the
sialic acid residues, the amount of charge lost, as calculated from the change in
electrophoretic mobility, is underestimated 2-3x, when compared with the
amount of sialic acid released. The reason for the discrepancy is thought to be
due to the presence of a layer of polyelectrolytes on the surface of the lipid-
protein bilayer through which the solvent can flow [11]. The sialic acid
charges are distributed within this layer, and not anchored to a rigid surface as
in solid colloidal particles such as charged latex spheres. The layer of
polyelectrolyte, called the 'glycoprotein calyx' is believed to extend out to -- 7-
10 nm beyond the bilayer, and its thickness is a function of the ionic strength.
The presence of such a layer would in part account for the fact that red cells
suspended in physiological saline (e.g. Ringers, a solution of ions mimicking
those in plasma, to which albumin may be added) do not aggregate. If treated
as a rigid colloidal charged particle, calculations of the double layer thickness
at physiological ionic strength yield a value o f - - 8 nm, i.e., within the
glycocalyx boundary and so close to the membrane that one would have
expected the cells to be subject to van der Waals attractive forces. In fact,
application of the DLVO theory of colloidal stability [12,13] to red cells
clearly indicates that the cells should aggregate in physiological saline [11,14].
However, red cells do aggregate in plasma in the presence of fibrinogen, as
well as in buffers containing plasma expanders such as dextran [14,15] and
polyvinylpyrrolidone. Within the primary aggregate, or rouleau, the cells lie
in a regular array as shown in Figure 1, discoid face to discoid face, and
deform so that the adjoining faces are locally parallel and have maximum
contact. Studies of aggregation in dextran solutions have been interpreted in
terms of cross-bridging by polymer molecules long enough to simultaneously
adsorb on the surfaces of two neighbouring cells [16]. In the absence of
reliable evidence that dextrans are adsorbed on red cells, this interpretation has
been questioned and an alternate mechanism, based on depletion flocculation,
proposed [17]. In this scenario, surfaces are initially repulsive and stabilized
by an electrostatic stress IIo~exp(-z/2~), where z is distance and ~ is the
electrostatic decay constant. With addition of non-adsorbing polymer, an
osmotic pressure difference builds up between the mid point of the
intermembrane gap and the external bulk solution. The osmotic pressure
induces an attractive depletion stress in which the decay constant is now the
Figure 1: A network of rouleaux of human red blood
cells photographed at rest on a microscope slide.

correlation length of the polymer ~, i.e., IIpeXp(-z/~), where ~, characterizes


the range of the depletion zone. At low [polymer] there is no adhesion since
the electrostatic stress is less than the depletion stress. With increasing
[polymer], the osmotic pressure difference increases, and so does the attractive
depletion stress, becoming larger than the electrostatic stress, and aggregation
occurs. This continues with increasing polymer concentration. However, at
the same time, ~ decreases thereby reducing the range and magnitude of the
depletion term. Ultimately, the repulsive electrostatic stress once again wins
out and the tendency to aggregation is reduced resulting in electrostatic
stabilization. These two competing stresses account well for the observed
aggregation vs [dextran] curves [15].

1.2 Blood vessels and flow regimes


Table 3 gives values of the vessel and flow parameters in the human
circulation. On leaving the left ventricle of the heart, blood enters the
ascending aorta having a resting diameter of-.25 mm, and is accelerated to
mean linear velocities as high as 0.9 to 1.4 m s-~. Almost immediately, the
blood is subjected to branching flows, the vessel diameters decrease and their
numbers increase rapidly, and hence the cross-sectional area of the vessel bed.
Values given for the dog show that the area increases from 5 c m 2 in the large
arteries, to 20 cm 2 in the small arteries to 1,360 c m 2 in the capillaries [18].
Broadly speaking, one can distinguish three flow regimes:
(i) A rapid, pulsatile flow in the aorta and larger arteries where inertial
effects due to acceleration and deceleration of the fluid predominate.
The mechanics of the flow have been modelled regarding blood as a
homogeneous fluid, neglecting its particulate character [19].
(ii) A much weaker pulsatile or quasi-steady viscous flow in the smallest
arteries and arterioles, in which interactions between the blood cells
largely determine the mechanics of the flow.
(iii) A viscous flow in the capillaries whose diameter is smaller than that of
the red cell, and which can be described as a bolus flow of a train of
cells surrounded by a thin lubricating film of plasma at the vessel wall,
with pockets of plasma (boluses) between cells [21]. The red ceils
readily deform into complex shapes [22,23], and their motion has been
modeled [24]. The poorly deformable white cells travel more slowly,
causing red cells to accumulate behind them [25,26].
Due to the multiplicity of branching, the non-uniformity of vessel diameters
and the existence of curved vessel segments, sudden changes in velocity and
direction of the bloodstream result in secondary flows having radial
components, and not infrequently in flow separation and the formation of
recirculation zones. Such flows have been designated by the term "disturbed"
[27], to distinguish them from laminar and turbulent flows, the latter observed
near normal aortic and pulmonary valves, increasing markedly in disease when
the valves become partially stenosed [28].
Branching occurs with such frequency that the disturbance created by the
flow at one bifurcation has not had time to dissipate before that due to another
branch comes into play. It is well known that these effects are particularly
important with regard to the genesis of thrombosis and atherosclerosis, and
clearly that they involve interactions between the corpuscles and fluid with the
vessel wall. Such interactions, which will be dealt with in Section 6 below, can
result in injury to the vessel wall and to the blood cells themselves.

2. M A C R O R H E O L O G Y OF BLOOD

The rheological properties of blood resemble that of a pseudoplastic fluid,


being a strong function of the rate of deformation. At low shear rates the
apparent viscosity decreases with increasing shear rate as the network of
rouleaux (Figure 1) breaks up. At moderate and high shear rate, the apparent
viscosity decreases as shear-induced deformation of the red cells increases.

2.1 Viscometry in Couette Flow


2.1.1 Effect of shear rate on apparent viscosity
The role of red cell deformation and aggregation is shown in Figure 2 in a
plot comparing the apparent relative viscosity, T/r at 37~ of 45% red cell
suspensions in plasma and in albumin-saline with that of aldehyde-hardened
(still biconcave) red cells in albumin-Ringer solution [29]. All suspending
media had the same viscosity, 1.2 mPa s, and measurements were made in a
concentric cylinder Couette device. Since there was no fibrinogen or other
aggregating protein in the albumin-Ringer suspension, no rouleaux were
present and the difference between rh in plasma and albumin-Ringer
suspensions in the range 10-z < G < 5 s-~ was due to the progressive break-up of
aggregates. The further decrease in rh is then ascribed to cell deformation, the
curve eventually flattening out as the blood assumes a quasi Newtonian
character at G > 10 3 s -1. By contrast, r/, for the rigidified cells is almost
constant with increasing G, there being no aggregation or deformation.

10 3

>.,

r,o
o
o 10 2
..=., ~ " ' ~ ' ~ R B C in plasma
> Aggregation ~ Hardened cells in alb.-Ringer
(D
>
...=.
o......-o.........~ ......~ ...... ~. _ ~'~... Deformation ~ C>
........ v ......,~...... o......-':~.. / '----',.--~ " - ~
R B C in alb Ringer "'" 0 "--" (~)
rr

.L,,I ........ I ........ I ........ I ...... A,I , , 9 , .... I

10-2 1 0 "1 1 10 10 2 10 3

Shear Rate, s ~
Figure 2. Relative apparent viscosity vs shear rate in suspensions
of normal RBC in plasma (aggregation and deformation), and in
Ringers-albumin (no aggregation) and hardened RBC in Ringers-
albumin (no aggregation or deformation). From Chien [29] with
permission.

2.1.2 Constituitive equations


A comparison of 11 equations which have been used to describe the
rheological behaviour of human blood was made by Easthope and Brooks [30].
The equation which best represented their data obtained in a concentric cylinder
measuring system was that of Walbum and Schneck [31]:

~'=a 1 exp a2H + G -1-a4H (1)

where H = hematocrit, and a~-a4 are adjustable constants, functions of


temperature, macromolecular composition of the suspending phase etc. The
equation was developed from a power law relationship between shear stress and
shear rate, using regression curve fitting of trial equations built by successive
inclusion of variables of decreasing influence. The model does not allow for a
yield stress.
Because of the complexity and high concentration of blood, it has not been
possible to develop constituitive equations based on a mechanistic model of
blood flow, without the use of empiricism or approximations. Such an equation
is that due to Casson [32] first used by Scott Blair [33]"
,~.1/2._ aol/2 + a l G a ~ 2 (2)

Here, the suspension is modelled as containing particles which can aggregate at


low shear rates to form rod-like particles, whose length increases with
decreasing shear rate. The similarity between these aggregates and rouleaux
(Figure 1) makes the model attractive. However, the yield stress, a0 does not
become zero except when H = 0. Nevertheless, Equation (2) represents the
rheological data on blood over a limited shear rate range with one set of
constants. Also, the experimental determination of yield stress poses real
problems, since aggregation at very low shear rates leads to migration of
rouleaux away from the walls of the Couette cylinders (syneresis), and
sedimentation is then also a problem [34].
Another equation based on mechanistic modelling is that due to Quemada
[35,36]"

~r _ (1 _ 0.5kH) -2, where k __ k0 + k o o t-~rl/2


--r and Gr- G (3)
1 + G" " F 1/2 - Gr
Here, the suspension properties are represented by the coefficients k0 and k..,
the intrinsic viscosities at zero and infinite shear stress of the particles which
predominate at those shear rates, and G~, the critical shear rate, which can be
considered to be the inverse of the relaxation time for the dominant structural
unit causing the suspension to be non-Newtonian. The coefficients are functions
of hematocrit, suspending phase composition, etc. The model has been found to
be almost as good as the Walbum-Schneck equation [31] with the advantage that
the coefficients have a physical meaning. Further discussion of this model and
its application may be found in reference [37], and a comprehensive discussion
of blood rheology in reference [38].

2.2 Blood flow in cylindrical tubes


The use of macroscopic rheological data obtained in Couette instruments and
continuum models to predict flow behavior in l a r g e vessels appears satisfactory
as long as there is no appreciable cell aggregation resulting in syneresis and
sedimentation. Both effects will occur in horizontally positioned tubes at mean
linear velocities < 1 tube diameter/second [39]. In vertically positioned tubes,
10

aggregation leads to syneresis and the two-phase flow of a core of aggregates


surrounded by a cell-depleted peripheral layer, with a reduced pressure
gradient at a given volume flow rate [40]. With these exceptions, continuum
mechanics appears to be satisfactory for vessels having diameters > 500 ktm.
As the tube diameter decreases, however, and approaches the dimensions of the
red cells, blood no longer acts as a continuum and the following effects arise.

2.2.1 The Fdhraeus effect


As the tube diameter decreases below 500 gm, the measured instantaneous
hematocrit in the tube, Hr, is found to be smaller than the hematocrit, HR, in the
stirred inflow reservoir, or even in the discharge, Ho (providing no screening
effects occur at the tube entrance). The ratio Hr/Ho decreases with decreasing
diameter, as shown in Figure 3, until the tube diameter falls below - 15 gm,
when it increases again. The effect is due to the existence of a slower moving
cell-depleted peripheral layer of low hematocrit, surrounding a faster moving
central core of higher hematocrit. When mixed, the tube hematocrit is smaller
than the reservoir or discharge hematocrit. It can be shown that:
Hr <u> <Up>
HD = <Uc~, = < f i c > ( 1 - H r ) +H r (4)

where <u> is the average tube blood velocity, and <Up> and <Uc> are the
average plasma and blood cell velocities [6].

! v ,

o, i
HD
0.6

04 i ~ Critica,,lDiameter
92 I~)0 I000
Diameter, l~m

Figure 3. F~thraeus effect for human red cells, HR = 40-45%, for


all literature data for suspensions in tubes, at flow rates ensuring
no RBC aggregation (cross-hatched region; also 9 F~hraeus data
[6]). Critical diameter (-2.7 g m ) ~ the smallest tube through
which a shuman RBC can flow. From [6], with permission.
11

2.2.2 The Fdhraeus-Lindqvist effect


The effect refers to the original observation [5] that the hydrodynamic
resistance of blood and other red cell suspensions decreases as vessel diameter
decreases below 300 gm diameter. The most obvious explanation for a lower
resistance or a lower effective viscosity (computed using the Poiseuille-Hagen
equation to distinguish it from the apparent viscosity, determined from the
measured shear rate and shear stress, as in Couette instruments [37]), is that it is
a consequence of the F~ihraeus effect, i.e. a lower tube viscosity due to a lower
hematocrit. It may also be due to the rheological effect of a non-uniform
distribution of cells across the vessel lumen or to the failure of the continuum
model of the suspension, as will be discussed in the next section.

3. M I C R O S C O P I C C O R R E L A T E S OF M A C R O S C O P I C F L O W
BEHAVIOUR

The general problem of microrheology is the prediction of the macroscopic


rheological properties of a material from a detailed description of the elements
of which it is composed. In the case of blood, the elements are the individual
corpuscles (effectively the red blood cells) each surrounded by the suspending
fluid, the plasma. What is most striking about the mechanics of the motion is
the fact that, under physiological conditions, the cell finds itself subjected to
shear stresses and considerable particle crowding, such that it is continually
distorted from the biconcave resting shape. The flow behaviour and
interactions of individual red cells has been studied using microrheological
techniques, in particular the travelling microtube, a device for tracking the
motions of cells and colloidal particles through vertically mounted precision
bore glass tubes of 50-200 gm diameter, while photographing or videotaping
through a high resolution microscope [41-43]. The microscope axis is fixed,
and the tube is mounted within a chamber attached to a vertically mounted
sliding platform supporting a syringe infusion-withdrawal pump. Both the
platform and the syringe pumps are driven hydraulically by continuously
variable speed electronically-controlled DC motor drives.

3.1 Rotation and deformation of red cells


At shear stresses z < 0.03 Nm -2, isolated human RBC rotate with periodically
varying angular velocity, but for small perturbations due to Brownian
diffusion, maintaining their biconcave shape [43]. The rotational orbits are
similar to those previously found for rigid discs [44,45], in accord with theory
[46] applied to rigid spheroids, as illustrated in Figure 5 for the angular motion
of the axis of revolution of an RBC and a 4-cell rouleau:

= 1G[ 1 + B(re)]COS 2q~ (5)


dt
12

where ~ is the azimuthal angle of the axis of revolution with the diametrical,
X2-axis (Figure 4; = + 90 ~ in positions 1 and 5, and 0 ~ in position 3 in Figure
5), B(re)= (r 2 - 1)/(r 2 + 1), r~ is the equivalent ellipsoidal axis ratio (axis of
revolution/diametrical axis) and T the period of rotation through 2n, given by
[47]:

T - 27r (re + re 1)
- G(r----~ (6)
Integration of Equation (5) yields:

tan~=re( Gt )
re + l / r e (7)

As predicted for an oblate spheroid having <re> = 0 . 3 8 , at any given instant,


the largest fraction of red cells in the tube (49%) are found with their major
axes within + 20 ~ of the direction of flow (r = 90 + 20 ~ [43]). Similar results
have been obtained with isolated platelets, except that here, the effects of
Brownian rotary motion are appreciable [48], as has been found with other
colloidal-size particles, such as doublets of latex spheres [49].
With increasing shear stress, the rotational motion progressively deviates
from that predicted by theory, as cells spend more time aligned with the flow.
At "t"> 0.1 N m -2, a large fraction are seen lying in the median plane of the tube
without apparently rotating, instead aligning themselves at a constant angle to
the flow; moreover, they are deformed with an increase in the major,
diametrical axis [43]. Such behaviour, resembling that of the deformation of
liquid drops in immiscible viscous fluids [7,50], has been studied with red cells
suspended in higher viscosity media such as buffered low molecular weight
dextran [43,51]. It has been shown that there exists a critical shear stress above
which the membrane begins to rotate about the interior of the cell, in what has
been called a tank-treading motion, an unfortunate term, since the membrane
motion is likely transmitted into the interior of the cell resulting in circulation
patterns within the hemoglobin solution. As predicted by a two-dimensional
theory applicable to Couette flow [52], an increase in the ratio of external to
internal viscosity promotes the stationary orientation of the particle.

3.2 Lateral migration of cells


The redistribution of blood cells in narrow tubes, which is at the root of the
Fhhraeus and Fhhraeus-Lindqvist effects, is due to a net lateral migration of red
cells away from the tube wall. There exists a substantial body of theoretical
and experimental knowledge on the effect of the vessel wall on the motions of
suspended rigid or deformable model particles (the reader is referred to
reviews by Brenner [53] and Leal [54]) and of blood cells, including platelets
and leukocytes [55,56].
13

X
3

r
'
"-1' U ~ GX
2
X
2
Figure 4. Rotation of the axis of revolution of a spheroid (heavy line)
defined by the Cartesian (Xi) and polar (0, ~0) coordinates constructed
at the particle centre of rotation and origin of a Couette shear field.

360
A i _
B,| I ......

270 _ i
! 4~~ ~
i/1
i1) i Ot~" r red cell I
i

180 ~.'~_._. ~r:;0;35 .....


-o
Zol i
"- i

1~ 4-cell rouleau
90 - ,~ r'i 1"1 -
i
I

0.25 0.50 0.75 1.00


t FLOW
T

Figure 5. Rotation of a single human RBC and 4-cell rouleau in Poiseuille


flow at G < 20 s-~. A: Variation of the angle ~ with time t during an orbit
having the period T. The line drawn through the solid circles was
computed from Equation (7); the dashed line corresponds to uniform
angular velocity. B" The same particles, drawn from cinemicrographs at
orientations corresponding to positions 1 to 5. From [43], with permission.
14

The vessel boundary exerts its influence on the flowing blood cells not only
by retarding their translational and rotational velocities, an effect appreciable
within one or two cell diameters from the wall, but also by generating radial
components in the cell velocity. There are two established mechanisms by
which migration across the streamlines can occur [45,47,53-55,57]: migration
due to particle deformation at low Reynolds number, and migration due to
inertia of the fluid at moderate and high Reynolds number. Here, we concern
ourselves mostly with the former mechanism.
Deformable particles (liquid drops, flexible fibers), suspended in
Newtonian media undergoing Poiseuille flow in the creeping flow regime,
migrate away from the wall towards the axis [45]. In the case of a fluid drop, it
has been shown to be an effect of the particle disturbance flow which generates
a flow with a radial component in the neighbourhood of the drop. In the
absence of fluid inertia, rigid spheres and spheroids do not migrate laterally
across the streamlines [45]. Under conditions of negligible fluid inertia,
particle Reynolds number < 10 -6, the latter defined in terms of the particle
translational slip velocity (particle [u] - fluid [U]) of a sphere, radius b at the
axis of a tube, radius R, in Poiseuille flow [58]:

Single human red cells and rouleaux also migrate radially inward, as illustrated
in Figure 6, for red cells in Ringers solution and in buffered dextran solutions
[55]. As previously found with fluid drops [45], the rate of migration increases
with particle deformation (it is much greater in dextran than in Ringer solution;
Figure 6) and increases rapidly with increasing ratio of particle to tube
diameter. Such migration, although severely inhibited by the particle-crowded
conditions in normal blood, nevertheless appears to result in a thin (-- 4 ktm
wide) cell-depleted layer at the wall and a significant lowering of the
hydrodynamic resistance in many blood vessels.

3.3 Cell interactions at normal hematocrits


Long before hematocrits approach those normally present in the circulating
blood (N 40-45%), even at 0.1% there are frequent two- three- and some multi-
body collisions, and these, as will be seen in Section 4, can be studied to obtain
information on the forces at play during such shear-induced interactions. As the
hematocrit exceeds 10%, there are constant collisions between all cells, and
above 30%, particle crowding becomes a factor and actually contributes to the
deformation of the red cells from their resting biconcave shape. Such
deformation can be observed even at shear stresses < 0.03 N m -2 at which the
isolated single cell in plasma would have rotated as a rigid disc (as in Figure 5).
Since one cannot observe the motions of individual blood cells in the interior
of whole blood flowing through narrow tubes, due to continual reflection and
15

~ -; ~, ~ ~
!
r
RBC in
E 40~ Dextran Ceins
O 0.2
or)

I,,,,-
X
X
A 0.4
(.9
V

0.6 , 1 , I ....

0.4 0.6 0.8 1.0


r/R
Figure 6. Measured inward migration of RBC in plasma Ringers
solution compared to that in 25% buffered dextran having 35x
the plasma viscosity. <G> is the mean tube shear rate, x is axial
distance from the tube entrance. From [55], with permission.

refraction of the transmitted light, it was necessary to render the blood


transparent. This was achieved by preparing "ghost red cells" by osmotic
hemolysis of normal red cells in buffer of 1/10 the ionic strength of plasma.
The cells were washed to remove the hemoglobin solution, and then slowly
returned to their normal salt and protein environment, whereupon they assume
their former biconcave shape [59,60]. Over a range of volume concentrations
from 10-80%, the measured apparent relative viscosities of red cell and derived
ghost cell suspensions did not differ significantly. Tracer normal blood cells
were added to the ghost cell suspensions and the motions of these, clearly
visible in the interior of the tube, photographed and analyzed.

3.3.1 Deformation due to shear and particle crowding


Figure 7 illustrates the continually changing deformation of a red cell in a
55% ghost cell suspension. The cells no longer undergo angular rotation but
spend much of their time aligned and deformed in the direction of flow. The
membrane likely rotates about the interior in an irregular fashion.
Deformation of rouleaux in ghost cell suspensions is also observed [60]. The
ability of red cells to deform and squeeze past each other in flow is the
microscopic correlate of the macroscopic flow behaviour of blood in Couette
viscometers which exhibits such remarkably low viscosity at moderate and high
shear rate, relative to that of concentrated suspensions of rigid particles and
even that of concentrated emulsions.
16

0 sec. 0.36sec. 0.72 sec. 1.08 s e c .

O.072sec.

' - - - - 10 ~ - - - - ~
I II - 5 0 p m
v" ~--

I
I
I
E

Figure 7. Tracings from photomicrographs of the deformation of a


tracer RBC at 7.2 ms intervals in a 55% ghost cell suspension being
tracked in robe flow. From [61], with permission.

3.3.2 Velocity distributions at high concentrations


The effects of particle crowding on the velocity distribution had previously
been explored in the tube flow of concentrated suspensions of macroscopic size
rigid spheres and discs. Here, the suspensions were made transparent to
transmitted light by matching refractive indices of particle and suspending
phase, and then adding visible tracer quantities of particles of the same size and
shape but having a different refractive index [62]. It was shown that for 0.05 <
b/R < 0.15, when particle volume concentrations exceeded 20%, the velocity
distributions became blunted in the centre of the tube with a core of radius r~ in
which particle velocities, u(r) were maximum and constant, = UM, and < U(O),
the centreline velocity in Poiseuille flow (as illustrated in the upper panel of
Figure 8). This region was designated as partial plug flow although this did not
imply that the profile was mathematically flat, only that there was no
17

1.0 ~ . l l l ' ]
Rigid ~ I
......~ Spheres ~ I

0.5
,~ " ~ ........... -
-I
r Complete Par
o
R Plug Flow , PlugFlow____~____;t~ ;
0 "- o 00000fO~176176

- : ~ e u , , , e -
0.5
9I .o~" Flow
......~ F'o'~
I .....~ i - Drops
1.0 L.... ................i
9 ~" ,

1.0 ' - ~ s I
t Cells~

0.5- 9 c = 32% ....~'0~

-~ 0 Plug Flow
- .__L_
I
~]__t_- ~~ /

.(7""
0.5 ~ D -
iscs Q

1.0 ........ ........ I I


O- 0.25 0.50 0.75 1.0
u(r)

U(O)

Figure 8. Dimensionless velocity profiles of visible tracer particles


in 32% transparent suspensions of model rigid particles and ghost
red cells: plot of relative radial distance vs. particle translational
velocity + centreline velocity in Poiseuille flow at the same volume
flow rate. Upper panel: Rigid spheres (b/R = 0.11) exhibiting
complete plug flow compared to emulsion droplets (b/R = 0.09)
with only partial plug flow. Lower panel: Rigid discs (b/R = 0.08)
compared to ghost cells (b/R = 0.11). From [62], with permission.

measurable velocity gradient. Nevertheless, these suspensions were still quasi-


Newtonian in that the velocity profile was independent of flow rate at low R%,
and the pressure drop per unit length, AP, was directly proportional to volume
18

flow rate, Q [62]. In fact, it was shown that, despite the continual erratic
(Brownian diffusion-like) radial displacements of the spheres in those regions
of the tube in which shear-induced interactions occurred, over relatively short
time periods, the displacements were reversible in time and space [62,63].
At a given b / R , the degree of blunting increased with increasing volume
fraction, c, of the disperse phase, and at a given c, it increased with increasing
b/R. Blunting of the velocity distribution was not due to aggregation of the
spheres or discs whose interactions are due solely to hydrodynamic forces, nor
due to a redistribution of particles in the tube which could have resulted in the
two-phase flow of a more viscous and concentrated central core of suspension
surrounded by a less concentrated and less viscous peripheral layer. With the
exception of the exclusion layer closer than one particle radius from the wall,
sphere centres were shown to be uniformly distributed in the tube [62]. Rather,
the effect is due to particle interactions in the crowded suspension, and has been
treated by Skalak [64] who has interpreted the interactions of the particles in
terms of passing versus non-passing motions of adjacent particles. It should be
noted, however, that more recent work has shown that redistribution of rigid
particles in variable shear fields can and does occur in the creeping flow regime
providing particle interactions proceed for long enough times [65-69].
As shown in the lower panel of Figure 8, at comparable particle volume
concentrations, the ghost cell suspensions exhibited blunting of the velocity
profile, similar to that Of the discs, but with significantly lower r c [59,70]. As
with suspensions of rigid particles, r c increased with increasing b / R and
concentration. However, as expected, given the shear and concentration-
dependent deformation of the ghost and red cells, there is no quasi Newtonian
behaviour: the degree of blunting decreased with increasing Q as the parabolic
Poiseuille velocity profile was approached, accompanied by a decrease in
effective viscosity [71 ], and the motions were not reversible in time and space.
In this respect, the non-Newtonian behaviour of red cells and ghost cell
suspensions resembled that of concentrated emulsions of deformable liquid
droplets in which the droplets are also deformed not only by the shear stress,
but also through particle crowding. As shown by work in transparent oil-in-oil
emulsions, the droplets were distorted from prolate ellipsoids into irregular
shapes [72]. Moreover, the shear-induced deformation led to some migration
of the droplets away from the wall (opposed by an outward dispersive force
due to the continual interactions of droplets in the core of the emulsion), and as
with the red blood cells, such migration was not reversible in time and space.
In addition (see upper panel of Figure 8) the velocity distribution was less
blunted than that of a suspension of rigid spheres at the same particle volume
fraction and b / R , and was flow rate dependent: as the flow rate increased and
the droplets became increasingly deformed, the velocity distribution gradually
approached the parabolic Poiseuille velocity profile [72,73].
19

3.3.3 Convective dispersion of cells


As pointed out above, the resistance to crowding of the blood corpuscles at
normal hematocrits opposes any substantial inward radial migration of red
cells, such as observed at low hematocrits. At the microscopic level, this
resistance is seen as a series of continuous collisions between red cells which
results in a marked dispersion of all blood cells and the surrounding plasma, as
illustrated in Figure 9 for a tracer red cell and a leukocyte in a 40% ghost cell
suspension. Measurements of the mean square radial fluctuations <Ar2> over
small time intervals At of the paths of tracer red cells and 2 ktm diameter latex
microspheres were used to compute dispersion coefficients D~ defined as:
Dr= <AP>/2At (9a)
At shear rates from 2-20 s -~, Dr for RBC and latex spheres ranged from
- 1 0 -12 to >10 -11m 2 s -1 [55,59], values 2 to 3 orders of magnitude greater than the

0.9

0.8
/
- 40

"l ~
"~
o.7

2
~PMN

5O 7O
Q
"
Q
5

~ 0. I .-i.
~ Q
"0 1,0~ u u , r I n

"- 35
"~ 0.9 "1:
tr 3

0.8 30

20

,
0 5 lO 15 20

Time,s
Figure 9. Radial dispersion of the paths of a polymorphonuclear
leukocyte (PMN) and an RBC in 40% ghost cell suspensions in tube
flow. Asterisks denote times when particles collided with the wall.
The slightly deformed PMN and markedly deformed RBC are
shapes seen while tracking the cells. From [74], with permission.
20

Brownian translational diffusion coefficient, Dr, for the isolated red cell
suspended in plasma, viscosity 7/- 1.8 mPa s at 23~
Dr = kBTr/K, ri = 4.4 x 10 -14 m E S -1 (9b)
where kB is the Boltzmann constant, TK the absolute temperature and
K, = 5.23x10 -5 m, the translational resistance coefficient for motion along or
transverse to the axis of revolution (2.4 l-tm) of a rigid spheroid, re = 0.38 [75].
The above values of radial diffusion coefficients in ghost cell suspensions are
in fair agreement with blood platelet translational diffusion coefficients
obtained from measurements of their dispersion in flowing blood, which
increased from < 1 0 -12 m 2 s -1 in plasma to 2.5 x 10 -~ m E s 1 at 50% hematocrit in
flowing blood at a mean tube shear rate <G> - 100 s -~ [76].
As a consequence of the enhanced radial dispersion of the other blood cells
by the red cell motions, is an increase in the two-body collision rate between
platelets in shear flow, which is discussed in Section 4. More important is the
effect of the red cells in increasing the frequency of platelet-wall collisions with
increasing hematocrit, an effect of considerable significance in hemostasis (the
normal plugging of cut or damaged vessel wall, a process initiated by platelets)
as well as in thrombosis (abnormal, pathological adhesion and growth of
platelet thrombi on the walls of arteries), particularly in regions of disturbed
flow, as described in Section 6. It has been shown that adhesion of platelets to
various artificial surfaces [77,78] as well as artery subendothelium [79,80]
(artery stripped of the endothelial layer thereby exposing collagen fibres to
which platelets strongly adhere) in flowing blood is markedly enhanced by the
introduction of red cell or ghosts cells into the suspending medium.
The extent to which red cell deformation or size may affect the increase in
the effective diffusion constant of cells and solute has also been studied. The
transport of molecular solute was found to increase slightly with cell size and
rigidity [81], and the wall adhesion of platelets observed to increase
substantially with red cell size [81]. Augmented transport in the shear flow of
concentrated suspensions of model particles and red blood cells has been
thoroughly reviewed by Zydney and Colton [82]. They proposed a model of
augmented solute transport based on shear-induced particle migrations and the
concomitant dispersive fluid motion induced by these particle migrations.
Augmented solute transport was defined as (D~e/DsF ) -- 1, where Dee is the
effective solute diffusion coefficient measured in the sheared suspension, and
D SF i s the solute diffusion coefficient in the absence of flow. If particle
rotations are assumed to be unimportant, D ~ - DSF + Dp, Op being the particle
diffusion constant. Augmented diffusion is predicted to vary as the Peclet
number, Pe - b2G/OsF.

3.3.4 Redistribution of blood cells in tube flow


At normal hematocrits and in vessels of diameters > 1 mm, the thin
21

peripheral layer of--4 gm thickness will likely have a negligible effect on the
distribution of red cells across the lumen. As mentioned in Section 2.2,
however, at low flow rates (corresponding to mean velocities <U> < 1 tube
diameter s-~) aggregation of red cells becomes an important factor.
Paradoxically, it is at low, and not high flow rates, that inward migration of
red cells in plasma or high molecular weight dextran buffer is most pronounced
in tube flow. Here, the formation and rapid inward drift of rouleaux of red
cells results in the two phase flow of an inner core of a network of rouleaux
surrounded by a peripheral cell-depleted layer in which can be seen single red
cells, small rouleaux and an apparently large number of platelets and leukocytes
[40,56,60,83-85]. Such two-phase flow has also been induced in small vessels in
the microcirculation of animals by intravenous injection of gelatin or
fibrinogen and of dextran. Both in vivo and in vitro it has been shown that the
formation of the red cell core is associated with the 'margination' (displacement
to the periphery) of white blood cells [56,84,85]. Cine films of blood flow
taken in tubes of 100-340 gm diameter clearly show that the effect is due to the
outward displacement of white cells by the inwardly migrating network of
packed red cell rouleaux [56,85]. Here, an "inverse" F~hraeus effect occurs:
since <u~> is less than <up> [Equation (4)], the leukocyte concentration in the
tube is greater than that in the infusing reservoir [56].
That such two-phase flow results in a decrease in hydrodynamic resistance,
R - AP/Q, had earlier been shown in studies of the oscillatory flow of
concentrated suspensions of rigid neutrally buoyant spheres where inertial
effects led to a small but significant inward migration of the particles from the
tube wall [86]. The effect was later demonstrated in the flow of mammalian
blood, where it was also shown that in vertically positioned tubes, enhanced red
cell aggregation in the presence of 250 kDa dextran resulted in a lower
effective viscosity at <U> < 2 s-~ [39]. By contrast, in horizontally positioned
tubes, where there is an asymmetric distribution of red cells and aggregates due
to sedimentation of the core, the effective viscosity continues to increase with
decreasing <U>. The relation between the radius of the core, r~, and
hydrodynamic resistance has been studied in buffered 110 kDa dextran
suspensions [40], and as shown in Figure 10 in a tube, R = 172 gm,
hydrodynamic resistance at first increased with decreasing <U> until
aggregation brought about syneresis and a shrinking diameter red cell core.
That the effect is due to aggregation was confirmed by experiments in 10%
buffered albumin in which rouleau formation was totally absent, and R
continued to increase with decreasing <U> down to < 0.2 s-~ [40]. As expected,
in buffered albumin suspensions there was no margination of leukocytes [56].
Redistribution of platelets in flowing blood has also been observed. In the
case of platelets in arterioles [87,88], and 1-2.5 gm diameter latex microspheres
in robes of-- 200 gm diameter [89,90], particle number concentrations near the
vessel wall have been shown to be higher than in the core of the flowing
22

1.0 ~ , ,

0.9

re

I-I 9 Citrated Blood


0.8 O Heparinized Blood
9 RBC in 1.5% Dextran

0.7
i i i i i i i i

3.0

co 0
,- 2.5
E
t~
Q- 2.0

.4 1.5
----it 9 1.5% Dextran ~

1.0 i i i i 1 i i i

0.05 0.1 0.2 0.5 1.0 2.0 5.0 10 20 50


<U>, Tube Diameters s 1

Figure 10. Development of a peripheral cell depleted layer in a 34%


suspension of RBC in 1.5% buffered dextran 110 in a 172 gm
diameter vertically mounted tube. The decrease in the relative
width, r JR, of the red cell core of aggregates (upper panel) occurs
in parallel with a decrease in hydrodynamic resistance (lower panel)
as mean tube velocity decreases below 1 s -~. For RBC in albumin
buffer (no aggregation) R continues to increase with decreasing
<U>. From [40], with permission.

suspension ("near wall excess"). The effect, which requires the presence of red
cells, has been modelled by adding a lateral drift term to the convective
diffusion equation for platelet transport in flowing blood [91]. The reason for
the net outward drift is believed to arise from the inward migration of the
more rapidly migrating red cells resulting in a marginal layer of lower red cell
concentration, whose width continuously fluctuates and should be viewed in a
statistical sense. The outward lateral drift of platelets or microspheres occurs
23

because of a net flux of particles from a region of higher red cell concentration
and hence higher red cell collision rate to a region of lower red cell
concentration and lower collision rate. Such a drift has been shown to occur in
alloys [92]. The fact that the location of the near wall excess of microspheres
occurs a few microns away from the wall is due to the fact that the particles are
physically repelled, and at sufficiently high Rep, fluid dynamically repelled
from the wall through inertial effects. It should be noted that the above
hypothesis does not attribute the motion of platelets toward the wall to their
exclusion by red cells in the interior of the tube, as is the case for leukocytes.

4. SHEAR-INDUCED TWO-BODY I N T E R A C T I O N S : F R O M
CHARGED C O L L O I D A L P A R T I C L E S TO BLOOD CELLS

Microrheological techniques, both in tube (travelling microtube [41-43]) and


in Couette flow (Rheoscope [93]) have been used to study two-body interactions
between charged colloidal particles and between blood cells. The left panel of
Figure 11 illustrates a two-body collision between equal-sized rigid spheres as it
appears when the particles are tracked by moving the tube upward with a
velocity equal to that of the downward flowing fluid at the mid-point of the axis
between the two spheres. The collision shown is one in which the trajectories
of the spheres are symmetrical, as previously observed in the case of neutral
macroscopic particles in viscous media [47,94]. The particles separate along
paths having the same radial coordinate, r, as those of the paths of approach. In
the case of colloidal-size charged latex spheres, however, interaction forces due
to double layer repulsion and attraction due to van der Waals forces come into
play when sphere surfaces approach to within a distance h - 100 nm [95]. This
results in asymmetric collision trajectories [96], as illustrated in the right hand
panel of Figure 11. These can be analyzed and hydrodynamic theory used
[95,97] to show that net interaction forces as small as 10-~3 N are detectable. In
the case of latex spheres in aqueous solutions of simple electrolytes, the
interaction forces have been interpreted by applying the DLVO theory of
colloid stability [12,13,95] and thereby obtain values of the Hamaker constant
and the retardation parameter of the van der Waals force [96].

4.1 Theoretical Considerations


The fluid mechanical problem of predicting the trajectories of two
interacting, neutrally buoyant rigid spherical particles of equal size, b,
suspended in a Newtonian fluid of Viscosity 77, undergoing simple shear flow
has been solved [97-99], and extended to the case of unequal-sized spheres
[100]. More importantly, it has been extended to the case when interaction
forces, Fi, t(h), other than hydrodynamic operate at h < 100 nm [95,101]. The
relative velocity of the sphere centres a distance s apart, is then given by [95]:
24

C(s*)Fint(h) (10)
-~ = A ( s * ) G b sin20 sin2q~ +
dt 3 rcb rl

and the angular motion of the doublet axis by:


dO 1
dt = -4 B(s*)G sin 20 sin 2q~ (11)

d~ 1
dt = -2 G[ 1 + B(s*) cos 2q~] (12)

-u(r) -u(r)
t t
i X3
i
i
i Repulsion
i
i

,, ,,
i Xm
8

X~

X3

Figure 11. Left: Two-body symmetrical collision in Poiseuille flow between


rigid spheres forming a transient doublet. The collision is tracked by moving
the tube with a velocity-u(r), equal but opposite to that of the fluid at the mid-
point of the doublet where Cartesian and polar coordinates have been
constructed. Right: Projection in the X2X3-plane of the asymmetric paths, due
to double layer repulsion, of the centres of two colliding 4 ktm polystyrene
latex spheres in 50% aqueous glycerol with l mM KC1. At the centre is the
collision sphere which cannot not be penetrated. From [96], with permission.
25

Here, 0 and r are the respective polar and azimuthal angles relative to X1 as the
polar axis, as shown in Figure 11; A(s*) and C(s*) are known dimensionless
functions of s* = s/b, which have been documented [95,97,99]. Equation (12 is
identical with the Equation (5) describing the angular velocity of a rigid prolate
ellipsoid [46] with B(s*)= (r~(s*)- 1)/(r~(s*)+ 1). Asymptotic expressions for
r~(s*) have been given for large and small s* [97].
Rearrangement of Equation (10) yields the force equation:

3rtbrl ds = A(s*) 3JrbqGb 2 sin20 sin2q~ + Fint(h) (13)


C(s*) dt C(s*)
in which the term on the left represents the hydrodynamic drag force resisting
approach of the particles, and the first term on the right the normal
hydrodynamic force, Fn, between spheres acting along the line of their centres,
being maximal for 0 = Jr/2, i.e. rotation within the X2X3-plane (Figure 11, left),
at r = - ~ / 4 when the force is compressive, and at r = rt/4 where it is tensile.
Equations (10)-(12) have been shown to apply in Poiseuille flow providing b/R
<<1 [96].
When F~n,(h) = 0 the trajectories of approach and recession of the colliding
spheres are symmetrical about the orientation q~- 0 of the doublet axis (Figure
11, left), and are defined by two constants, D and E given by the integrated
form of Equation (10) [96,97]"

x l =+2s*Dfls*) and x z =+2s*fls*)[E+ g(s*)] (14)

where g(s*) and f(s*) are integral functions of s* for which numerical values
have been given [96,97]. When Fi,.(h) :r O, the paths of approach and recession
are no longer symmetrical, and D and E become functions of time, increasing
when F..t(h) > 0 (repulsion), and decreasing when Fi,.t(h) < 0 (attraction).

4.2 Determination of the Interaction Force


In principle, Fi,~(h) can be computed from Equation (10) if one can measure
the observed translational velocity, ds/dt. This turns out to be difficult since the
trajectory may lie out of the median, XzX3-plane of the tube, and even for
collisions in the median plane it is impossible to measure ds/dt when h < 100
nm, just when particle interaction forces begin to be important. Instead, the
experimentally observed trajectories are fitted by numerically integrating
Equation (10) and solving for D and E [96]. In addition, in the case of the
charged latex spheres, by using expressions from the DLVO theory, it is
possible to compute the Hamaker constant, A, for the latex:
F~nt(h) -- Fattr(h) + Frep(h) ,
26

Fattr(h ) =
12h2
I(1+
where the van der Waals attractive force, Fattr(h ), is given by:

1 + 1.77p 2
p<l (15a)

Ab 10.98 0.434 0.0674 p > 1 (15b)


Fattr(h) - 12h2 p p2 I- P3

and the double layer electrostatic repulsion force, Frep(h), is given by:

Frep(h) = 2~Kbe~ / /
1 + e -rh (16)

Here, p = 27rh/A, is the retardation parameter, ~, the London wavelength, K the


dielectric constant of the suspending medium, e0 the permittivity of free space,
gt0 the surface potential (usually taken to be the measured (-potential), and tr the
reciprocal Debye double layer thickness. Analysis of trajectories such as those
shown in Figure 11 for polystyrene latex spheres in 50% aqueous glycerol at 1
mM KC1. led to a value of A = 3x10 -21 Joule in good agreement with a value
(2.7x10 -21 Joule) extrapolated from that determined in water [96].
It should be noted that Equation (10) neglects the effect of the double layer
on the velocity field around the spheres. The layer introduces coupling forces
and torques which should be included in the functions A(s*), B(s*) and C(s*)
[101]. Fortunately, under the conditions of the experiments shown in Figure
11, fight, tcb >> 1, and electroviscous forces and torques were negligible.

4.3 Two-Body Collision and Collision Capture Frequencies


4.3.1 Collisions in the absence of interaction forces
Traditionally, the frequency with which two spheres collide to form doublets
in a sheared suspension containing No singlets was computed assuming, as
originally done by Smoluchowski [102], that all particles move along linear
trajectories until they collide with a sphere. A reference sphere of radius b is
located at the origin of a Cartesian coordinate system and a simple shear field,
and all sphere ccntres which pass through a collision sphere of radius 2b and
cross-sectional area 4~b 2 will undergo collisions. The general case of the two-
body collision frequency, J~, between a reference sphere of radius b~ and
spheres of radius b z , concentration N2 per unit volume, is then given by:
4 3
Js - -~ (1 + b 1/b 2) 3N2Gb 1 (17a)
27

32 3
= --~ NoGb (17b)
for equal-sized spheres of radius b.
The classical Smoluchowski theory has been corrected using the rigorous
hydrodynamic theory of the curvilinear particle collision trajectories described
by Equation (10) with F~,~(h) = 0 [103]. The difficulties in defining a collision
were eliminated by defining a "curvilinear collision sphere" of radius p,,
around the reference sphere. Correction factors to be applied to the results of
the rectilinear theory (collision sphere radius = 2b) were then calculated for the
cross-sectional area of the collision sphere, the collision frequency and the
average doublet lifetime as a function of the ratio p,,,/b. Curvilinear and
rectilinear theories were shown to be compatible at p,,,/b = 2.38, which appears
to correspond to the minimum dimensionless distance of approach of the
spheres recognized experimentally by different observers as a collision [47].

4.3.2 Collisions in the presence of interaction forces


When interparticle forces act between the spheres, the collision capture
frequency per particle, J,, giving the number of collisions that result in
permanent doublet formation, is then defined as"
32
Js = - ~ aoNo Gb3 (18a)

=
8aocG
zc (18b)

where ao = JJJ is the capture efficiency and c is the volume concentration..


Capture efficiencies have also been computed using rigorous hydrodynamic
theory. In this case the collision capture cross-section was found by integrating
the trajectory Equations (10)-(12) forward in time by trial and error [104].
Capture efficiencies were calculated as a function of the ratios of the interaction
forces Fr~p and Fa,,, to the hydrodynamic force for equal sized [104] and
unequal-sized spheres [105]. In the absence of repulsive forces, for equal sized
spheres, ao decreases with increasing G, as would be expected from Equation
(13), since with increasing G both the hydrodynamic drag and the
hydrodynamic normal force between spheres increases while the interaction
force remains constant. When both repulsive and van der Waals forces act, ao
first increases at low G, then decreases at moderate and high G. The values of
ao depends strongly on the ratio of the sphere radii, being appreciably greater
for equal sized particles, since the minimum distance of approach between
spheres actually increases with decreasing radius ratio.
28

Measurements of the number of doublets formed in the Poiseuille flow of 1


~m diameter polystyrene latex spheres at 1 mM KC1 demonstrated that ao
decreased with increasing G in going from tube centre to the periphery [104].
Experiments with aqueous suspensions of 2.6 ~tm diameter polystyrene latex
spheres in Poiseuille flow showed that at 10 mM KC1, many of the two-body
collisions resulted in the formation of permanent doublets [106]. Some of the
spheres of these doublets were in a secondary potential energy minimum at
distances h between 15 and 17 nm, as shown by measurements of the period of
rotation [Equation (6)]. Theory shows that the dimensionless period of
rotation, TG, depends not only on h but also on whether the spheres are capable
of independent rotation [97]. There is a different solution for each of these
cases, and TG at a given s* is markedly greater for independently rotating
spheres than for rigidly coupled spheres.
By contrast, studies carried out in the presence of a cationic polyelectrolyte
revealed the formation of permanent doublets many of which rotated as
dumbbells of rigidly-linked spheres, being in a primary potential energy
minimum, and exhibiting values of TG close to the predicted value of 15.62. At
sufficiently high polymer concentration, polymer bridge formation appeared to
provide the most reasonable explanation of their flow behaviour [107].

4.4 Two Body Collisions Between Blood Cells


In the circulation, aggregation of blood cells and their adhesion to the vessel
wall is most often mediated by cross-bridging of biological polymer adhesion
molecules (carbohydrates and proteins) through non-covalent receptor-ligand
linkages. These interactions occur under flow-induced normal and shear
stresses, and therefore, characterization of the physical strengths of such bonds
is an important prerequisite for understanding adhesion processes such as occur
in cancer metastasis, platelet thrombosis and leukocyte margination and
extravasation. Modem techniques in physical chemistry, cell and molecular
biology have made it possible to relate the microrheology of cell-cell and cell-
wall adhesion to the actual biological polymer molecules (receptor and ligand)
and bonds involved in these processes. One has therefore been able to study the
effect of fluid mechanical forces on intermolecular reactions and deal with the
transduction of fluid mechanical forces via the cell membrane into the cell
interior through the activation of second messenger systems [108-110] and the
relation between the chemistry of cell motility and the forces generated by the
cell [111]. Some studies, concerned with rheology at the molecular level, are
now described in this and the subsequent Section.

4.4.1 Application to shear-induced aggregation of platelets


The kinetics of the adenosine diphosphate (ADP)-induced aggregation of
native human platelets in plasma and of washed cells in physiological salt
solutions subjected to shear in small tubes have been investigated. The ADP
29

concentrations chosen were those just able to induce aggregation, so that the
effects of shear and time of exposure to ADP would be at their most sensitive.
The magnitude and time of exposure to shear were varied independently by
using a double infusion system [112] in which the cells and agonist
simultaneously entered a tiny mixing chamber and flowed through preset
lengths of polyethylene tubing at mean transit times from 0.2 to 86 s and mean
tube shear rates, <G>, from 40 to 1920 s -~ (representative of those in the
circulation; Table 3). The suspensions were collected into glutaraldehyde to
arrest the reaction, and all particles in the volume range 1 to 105 ~tm 3 counted
and sized using an aperture impedance counter. Aggregation was measured by
the decrease in the number of singlets, and the number and size of aggregates.
Adenosine diphosphate-induced aggregation is known to be mediated by the
plasma protein fibrinogen which specifically binds to the activated form of the
platelet membrane integrin glycoprotein complex, GPIIb-IIIa [113]. Most
likely, aggregation occurs via cross-bridging of fibrinogen molecules between
receptors on adjacent cells, (Figure 12) [114]. Successful cross-linking requires
simultaneous binding of opposite ends of the fibrinogen molecule to two cells
immediately after activation of the GPIIb-IIIa complexes, and before their
saturation with fibrinogen, an unlikely scenario given the high plasma
fibrinogen concentration (2x107 molecules available per platelet) and receptor
surface density (5X104 sites per cell). A model of aggregation therefore
requires either a low affinity for fibrinogen binding with the continuous
breaking and making of bonds, or the time-dependent exposure of new

i
t'~, C
r
/ r
Figure 12. Schematic diagram of platelet activation by ADP causing shape
change from smooth discocyte (A) to spiny spheroid (B) with pseudopods,
and transformation of the membrane glycoprotein IIb-IIIa receptors
(inset) into active forms (o) capable of binding fibrinogen (Fb; C) which
can cross-link cells producing doublets. From [115], with permission.
30

receptors that permit cross-linking to occur during the time of collision.


In platelet-rich plasma (no red cells present) prepared from blood
anticoagulated with sodium citrate ([Ca ++] -- 45 ~tM), at 0.2 ~tM ADP, the
presence of a shear-dependent initial lag time indicated that both a weak and a
strong platelet-platelet bond is involved in aggregation at physiological shear
rates [116]. As illustrated in Figure 13 the initial rate of disappearance of
singlets into microaggregates was the highest at the lowest shear rates, but later
in time decreased markedly. As the shear rate increased, the initial rate of
aggregation decreased; nevertheless a relatively high rate of aggregation was
sustained at later times. These results were interpreted using two-body
collision theory even though activated platelets are not rigid spheres, but
spheroechinocytes (spiny spheroids) with numerous pseudopods of appreciable
length capable of significantly increasing the effective collision cross-section of
the cell [117]. With platelets, the interaction force is in part repulsive, due to
like surface charges (negative charges on the platelet surface glycoproteins rich
in sialic acid), and attractive, due in part to van der Waals forces, but mainly
due to polymer bridging by fibrinogen. Since net attractive forces appear to
operate, a 0 was expected to decrease with increasing shear rate.

[ ' I ' I ' ' ' ' '

100 <G> - ~ 41.9 S-1 -[

335 S -1
~ 80
--~ 1335 S-1
-~ 60

t~
~ 40

r9~ 2o

0 20 40 60 80 100
M e a n Transit Time, s

Figure 13. Effect of shear rate on the ADP-induced


aggregation of platelets in plasma, expressed as % singlets
remaining vs mean transit time in the tube (R = 0.6 mm);
[ADP] = 0.2 gM; bars are + 1 S.E.
31

Indeed, ao, computed from Equation (18b) over the first 4 seconds, was
greatest at the lowest shear rate (= 0.26), but decreased markedly to 0.017 and
<10 .3 at the intermediate and highest shear rates, respectively. Thus, it appears
that a weak bond between cells of the microaggregates formed within the first
few seconds, diminished in strength with time, and the low collision frequency
at low shear rate was unable to sustain a high rate of aggregation. As shear rate
increased, the initial decrease in ao led to a decrease in rate of aggregation, yet
at later times the high collision frequency, aided by the arrival of a more shear
stress-resistant bond sustained a high rate of aggregation.
In the presence of red cells at 40% hematocrit, the initial rate of single
platelet aggregation in whole blood is an order of magnitude greater than in
platelet-rich plasma [118,119]. There has been a long-standing discussion on
whether this is due to a chemical (release of ADP from the red cells) or a
physical effect of the red cells. As shown in Figure 14, some aggregation of
platelets in blood at low shear rate was observed in the absence of any added
ADP, and this could be abolished by the addition of an ADP scavenger enzyme
[119]. However, there remained a large effect believed to be a physical one.
An obvious candidate for the increased rate of aggregation is the increased
platelet diffusivity in blood, described in Section 3, which might lead to an

100 ._ - - - - _ - - = - _ - - -

9~ ~~ -- "" -- ~BloodL,H = 40%, no ADP -

8o

60

40

20

0
0 10 20 30 40 5O
M e a n Transit Time, s

Figure 14. Plot as in Figure 13 showing marked effect of RBC


on the ADP-induced aggregation of platelets in blood in tube
flow at <G> = 42 s -a. From [118], with permission.
32

increased two-body collision rate. However, arguments based on calculations


of a Brownian translational motion two-body collision frequency [102]:
JD = 16~bNoDr (19)
show this is unlikely. Here, D r is now the effective diffusion coefficient,
obtained from measurements of the radial dispersion of platelets in blood
undergoing tube flow (of the order of 5 -~ m 2 s-~). Computations indicate
that the two-body collision frequency would increase by 156% at <G> = 42 s -~,
but only by 19% at <G >= 335 s ~, over that in platelet-rich plasma, too small to
explain the large increase in the observed initial aggregation rates [118].
Instead, the hypothesis that an increased a0 results from an increased collision
doublet lifetime, To, was tested. To this end 2.5 ~tm diameter latex spheres
were suspended 40% red cell ghost suspensions in plasma as a model of platelets
in blood and the measured distribution, '~'Dmeas, compared with that in plasma in
the absence of multibody interactions, and with that, TDtheor , predicted from
Jeffery's theory. From Equation (7) with r~ = 1.98 for a doublet of rigidly-
linked, touching spheres [F~,,(h) = 0],, the doublet lifetime is:

- ~
TDtheor- 4"97 tan-l(1 tan q~0) (20)

where ~0 =-~00 is the collision angle when the spheres are first seen to make
apparent contact. In the absence of other interactions, the doublet is predicted
to rotate until the orientation q~ = +r the reflection of the collision angle, is
reached when spheres separate. Figure 15 illustrates the erratic time course of
the ~0-orientation of the axis of a collision doublet, compared to that computed
from Equation (20). For 320 doublets, the m e a n TDmea/'~Dtheor : 1.61 + 1.80
(S.D.) compared to 1.00 + 0.31 for 90 doublets in plasma. Analysis showed
that impedance of the rotation is the likely mechanism by which "~Dmeas > "~Dtheor"
A more sophisticated analysis of shear-induced platelet aggregation uses
population balance mathematics and a laminar shear coalescence equation
[120,121] in which a particle density function expresses the makeup of a
collection of singlets and aggregates of various sizes. The collision efficiency
and an aggregate void fraction (to avoid the problem that platelets and are not
spheres) are introduced and experimental data matched by computer analysis.
A population balance analysis has also been used to model the shear-induced
aggregation and disaggregation of activated neutrophils [122,123].

0 E F F E C T OF FORCE ON THE K I N E T I C S OF C E L L - C E L L A N D
CELL-WALL ADHESION

Adhesion between cells is generally mediated by weak non-covalent


molecular linkages which, in the system at rest, can nevertheless persist for a
33

90

60 ] 9 _ .

Measured

30 I A Theoretical
I I
I I
-30 I 1
-60

-90 , ~ , " I , I i I
0 0.05 0.10 0.15 0.20 0.25
Time, s
Figure 15. Time course of the ~0-orientation of a collision doublet of
2.5 gm latex spheres in a 40% ghost cell suspension (O) compared to
that for an undisturbed doublet (A) formed at the same initial ~ = -~o
and G, predicted to separate at the mirror image q~ = +~Oo. Dashed
lines give time interval = "~Dmeas(> "~Dtheor) o f apparent contact between
the spheres. From [119], with permission.

long time. In contrast, the linkages quickly rupture when mechanical stress is
applied to the cells. Thus, molecular attachments appear to be maintained by
kinetic traps and bond rupture should be regarded as a kinetic process far from
equilibrium, with apparent bond strengths expected to depend on the rate of
force application. With the advent of ultrasensitive mechanical techniques such
as the atomic force microscope [124,125], the biomembrane force probe [126;
Section 5.3] and optical tweezers [127], the properties of single adhesive bonds
at cell interfaces can be tested with nanoscale resolution. In this Section, the
focus is on bonds made and broken in the flowing blood, and discussion is
limited to the kinetics of bond rupture between freely suspended cells in shear
flow, and to the kinetics of leukocyte rolling along, and adhesion to the vessel
wall. In both cases, the experimental results have been modelled using relations
for the effect of an external force on the off rate constant of a single bond.

5.1 Models of Bond Rupture under Force


5.1.1 Bell model
Bell [128] was the first to treat the increased rate of bond dissociation under
force in biological systems, in particular with respect to receptor-ligand bonds,
34

applying to such bonds a model of Zhurkov [129] for the force dependence of
bond rupture based on empirical studies of the fracture of macroscopic solids.
This model is actually an extension of transition state theory in gases [130] in
which it was proposed that intermolecular forces in a gas could be treated as a
one dimensional random walk in a potential energy well. The probability of
escape depends on the depth, E 0, of the well and the natural frequency of the
bond in vacuum, "to (-- 10 -~3 s for atoms in a solid). For a parabolic energy
barrier it was shown that the escape or break-up time in the absence of an
extemal force is given by:
t o = "to exp (E o/kBTK) (24)
where the exponential represents the probability that thermal fluctuations
provide enough energy for the barrier to be surmounted (for the transition
state to be reached). Zhurkov introduced the force dependence by postulating
that the extemal force, f per bond, acts directly along the reaction coordinate x
to reach the value xa at the transition state, and that the force reduces the energy
barrier in a linear fashion; thus
tb= To exp (E 0 - f x ~ )/kBT K (25a)
= t o exp (-z'f) (25b)

where Z = xa/kBTK. Therefore, when f reaches the value f/3 - Eo/xa, t~ = "t'0, and
bond failure is almost instantaneous. By identifying tb with 1/~r the reverse,
or off-rate constant, the connection to bond reaction rates is made:
tr = tCog~ exp (z'f) (26)
where tCog~ = 1~to. The probability of bond breakage Pb in a short time interval,
At, is then given by [ 131 ]"

Pb = 1 -- exp (-tCoffAt) = 1 - exp [ 0te x p(x'f) At] (27)

5.1.2 Evans model


In the Bell model, all the force-driven features are lumped into one
p a r a m e t e r - a length x~, often referred to as the 'range' or 'width' of the bond,
that gives little insight into molecular events. An alternative form was
therefore proposed [132] for forces sustained at a level nearf~:

tb= "gOexp -f- (28)

where ~:0 is again the time for rapid bond rupture and m was used to capture
variations in rupture behaviour from ductile (m < 1) to brittle failure (m > 1).
Evans et al. showed that any realistic technique for putting stress on bonds will
35

have a rise time or rate of loading and this, coupled with the inherent stochastic
nature of receptor-ligand bonds, will lead to the peaks observed in the
histograms of bond strength from mechanical probe experiments [132,133].
Hence, the strength of a single bond is governed by the rate of loading, and they
have used the Brownian dynamics theory of Kramers [130] to show that the rate
follows a general form given by:
to# = to#~ gff) exp [AEo(f)/kJK] (29)
which follows the required phenomenological descriptions for the force-driven
kinetics [134].
These models of bond rupture under force have been used to analyze
experimental data obtained on the break-up doublets of red cells, and on the
roiling of leukocytes on endothelium, as described below.

5.2 Cell-cell adhesion- break-up of doublets of red cells


Studies of two-body interactions between charged latex spheres described
above led to similar experiments on the interactions between sphered red blood
cells. Rigid sphered swollen human red blood cells (SSRC) of known antigenic
type were prepared, and permanent doublets formed by cross-linking the
spheres with the corresponding antibody [42]. Normal red cells were sphered
and swollen in 0.2M buffered glycerol containing sodium dodecyl sulfate, fixed
in 0.08% glutaraldehyde, washed and suspended in buffered media containing
viscous co-solvents such as glycerol, sucrose or 40 kDa dextran. In the absence
of antibody, two-body collisions between the 6.3 l.tm diameter SSRC involving
transient doublet formation were found to be markedly asymmetric with an
apparent repulsion, the trajectory constant E increasing and the minimum h >
50 nm [135], i.e., a distance large compared to that at which double layer
interactions become important. The result was likely due to the interaction,
principally steric, between the glycocalyx of the colliding cells.

5.2.1 Hydrodynamic force of break-up


Instead of studying 2-body collisions, therefore, the model of doublets of red
cell spheres, cross-linked by antigen-antibody linkages was used in order to
measure the hydrodynamic forces required for bond rupture. The SSRC were
suspended in buffered aqueous 48-56% sucrose of known viscosity containing
0.075-0.6 nM of monoclonal anti-blood group antibody. Permanent doublets of
SSRC formed through two-body collisions when the suspensions were sheared
at low G (< 10 s-~). Measurements of TG showed these doublets to be rigidly
coupled [42]. For such a doublet in shear flow, ds/dt = 0, and if it can be
induced to break up, then, from Equation (13), F~,,(h) can be computed:
A(s*)
Fint(h) = - Fn = - C(s*-"'~37cbr/Gb2 sin20 sin2q~ (30)
36

However, the bonds linking the red cell spheres are also subject to
hydrodynamic forces acting normal to the doublet axis, i.e. shear forces. The
general method of calculating forces, torques and translational velocities of
neutrally buoyant rigid spheres in a simple shear field [97], based on the matrix
formulation of hydrodynamic resistances in creeping flow [136], was used to
compute the normal force acting along, and the shear force, F~, acting
perpendicular to the axis of the doublet [137]:
F, = a,(h)rlGb 2 sinZO sin2dp [31]

F s = O~s(h)rlGb2 sin 0
1 - sin 20 COS2~)
/ (2sin20 cos2~- 1)2sin2r /-
+ cos20 C0S2~
1

[32]

where a,(h) and a,(h) are force coefficients weakly dependent on the minimum
distance of approach, h.

5.2.2 Measurement of the hydrodynamic force at break-up


To measure the hydrodynamic force required to break up permanent
doublets of SSRC, the suspensions were subjected to a slowly accelerating
Poiseuille flow in the travelling microtube apparatus, and the particles tracked
and videotaped until break-up occurred. From the videotape replay, the
doublet radial position and velocity yielded the local shear stress as well as the
particle orientation at break-up, enabling F, and F~ to be computed. It was
estimated that, at an antibody concentration of 0.3 nM --- 15 IgM molecules were
available over the estimated 2000 antigenic sites in the contact area of the SSRC
(2.5 x 104/lal) available for cross-linking (between h = 20 and 45 nm [42]).
Although this implies a maximum possible 30 cross-bridges, the actual number
of bonds was most probably much smaller, as was confirmed by the computer
simulation of a stochastic theory of bond rupture (see below). Thus, the
fraction of two-body collisions resulting in permanent doublets at shear rates <
10 s-1, was estimated to be only ---1%; moreover it was unlikely that all available
IgM molecules would be bound to antigen, and many of the bonded antibody
molecules would not participate in cross-linking.
In calculating the forces at break-up, it was first assumed that the cells
separated the instant a critical force to break n bonds of strength f, F = nf, was
reached. Figure 16 shows scattergrams of the normal force at break-up, F,,, as
a function of antibody concentration and suspending phase viscosity. It is
evident that there is considerable overlap in the range of F,, without any
evidence of clustering around a set of discrete force levels corresponding to
different numbers of antigen-antibody crossbridges, although the mean
measured force increased significantly with antibody concentration (Figure 16).
37

250 <>
58%

200 56?
Z
56%
150 50~ m

8 ~
49 o -,~ozx o
6 O
100 46% A ~ 0

5O

0
0.15 0.30 0.60 0.15
[IgM], nM [IgA], nM
Figure 16. Scattergram of the distribution of the normal hydrodynamic
force to separate doublets of SSRC cross-linked by monoclonal IgM and IgA
antibody, suspended in buffered 46-58% sucrose in an accelerating Poiseuille
flow. Also shown are mean values + 1 S.D. From [93], with permission.

The absence of clustering of F, values suggested that bond rupture leading to


doublet break-up was a stochastic process, both time and force dependent, as
shown by Evans et al. in studies using the micropipette aspiration technique
[132]. Thus, in the accelerating Poiseuille flow experiments a doublet might
well have broken up at a separation force lower than that measured at break-up,
had the particle been given enough time under a lower force.

5.2.3 Time and force dependence of break-up


Accordingly, experiments were undertaken in Couette flow using a
transparent counter-rotating cone and plate d e v i c e - the Rheoscope [51,93] - in
which a constant shear stress could be applied virtually instantaneously, and the
time dependence of break-up studied as a function of the applied force.
As predicted, at a given applied force, there was a pronounced distribution in
times to break up, while the average time to break up decreased with increasing
applied force. Figure 17 (left panel) shows the observed temporal distribution
of break-ups plotted as a function of the dimensionless number of rotations
(t/T) from the onset of shear. It can be seen that, in the intermediate and high
force ranges, most of the break-ups occurred within the first 2 to 3 rotations.
In the lowest force range, the fraction of break-ups fluctuated between 1.5 and
4% in the first 7 rotations, after which no more doublets broke up.
38

0.25 0.25
F n <70pN F n = 55 p N
o

o
0.20
I
9.-o.. 70<F n<140pN , 0.20 I
I
9-.o..- Fn=lllpN

I ---i- Fn> 140pN


l ----A- F n = 1 6 7 p N

0.15 - I 0.15 1
O
..I l
9 l
o
- 0.10 0.10 -" I
0
0
A
-9 0.05 "~ 0.05 - i "A..-~ x ~' -~.~-]
o
/A \ "..,"1 9 *,,,4

ra~

0 0 | I I ,1 I I |
0 ] 2 3 4 5 67 8 910 0 1 2 3 4 5 6 7 8 9 10
Number of Rotations Number of Rotations

Figure 17. Left: Fraction of doublets of SSRC, cross-linked by 75 and 150


pM IgM antibody, breaking up per rotation in Couette flow after onset of
shear. Right: Computer simulation of break-up for the best fit parameters
xa = 0.40 nm, to = 100 s, <no> = 4.0. From [93], with permission.

5.2.4 Locus of bond rupture: use of antigen-linked latex spheres


There remained the question of the locus of bond rupture in the SSRC
studies. Micropipette aspiration experiments have shown that it is possible to
extract antigen under force from a normal red cell [132]. Subsequently,
experiments in shear flow showed that blood group antigen extraction can
occur even with fixed cells [ 138] which cast doubt on the assumption that break-
up of the SSRC crossbridge is due to failure of antigen-antibody bonds.
Therefore, studies in which a synthetic blood group B antigen was covalently
linked to carboxyl modified polystyrene latex spheres and the spheres cross-
linked by monoclonal IgM antibody were undertaken [139]. Here, break-up of
doublets of microspheres should occur through the rupture of antigen-antibody
bonds, and would be expected to lead to a significantly different force
dependence on doublet lifetime. As with the SSRC over a comparable range of
normal force, there was a distribution in times to break up. However,
significantly higher forces were required to achieve the same degree of break-
up for doublets of antigen-linked spheres than for SSRC.
39

5.2.5 Monte Carlo simulation of doublet break-up


Using the Bell theory [128], a computer simulation of doublet break-up
under shear was developed in which to and x, [Equation (25)] were varied to fit
the data collected in the Poiseuille and Couette flow experiments. Since bond
formation is thought to be a Poisson process, specification of the average
number of bonds, <nb>, is all that is required to characterize the distribution of
bond numbers used in the computation of the force per bond when fitting the
experimental data. Each rotation was divided into 1000 equal time steps, and
for each step Pb was computed from Equation (27) and the force per bond
calculated from Equation (31) for the current instantaneous values of 0, ~0 and
nb. A random number between 0 and 1 was chosen from a uniform distribution
for each bond remaining. If the number drawn was less than Pb, the number of
bonds was reduced by one, and the force per bond on the remaining bonds
recalculated. Thus, bonds are postulated to break sequentially. The results
obtained for the best fit of the Couette flow data with 75 pM IgM are shown in
the fight hand panel of Figure 17, with x,= 0.40 nm, to = 100 s and <nb> - 2.5
bonds. It is evident that the general features of the real experiments are quite
well reproduced.
Computer simulation of the break-up of doublets of antigen-linked latex
spheres indicated that differences between these and populations of SSRC were
due to a change in bond character. The antigen sphere-IgM bond was less
sensitive to the applied force, had a lower value of x, (0.12 nm) and required a
higher value o f f , (1000 pN) for instantaneous rupture compared to that (360
pN) for the SSRC [139]. This supported the notion that antigen-antibody bonds
were ruptured in the case of the antigen-linked spheres whereas antigen was
able to be extracted from the membrane of the SSRC.

5.3 A Biosurface Force Probe


In Section 5.1 it was pointed out that the strength of a single bond is
governed by the rate of loading, a factor which could not be controlled in the
break-up studies described above. Evans et al. [126], describe an ultrasensitive-
tunable force transducer that can measure forces over an incredible range from
0.01 pN to > 1000 pN - the strength of a covalent b o n d - and in which the rate
of loading can be controlled. As illustrated in Figure 18, the transducer is a
microbead probe attached to a pressurized membrane capsule (e.g. RBC or
lipid vesicle). The pressure, P, is controlled by micropipette suction and sets
the membrane tension Zm:
Rp
Zm = P2( 1 _ Rp/Ro ) (33)

where R 0 and Rp are the radii of the membrane capsule and suction pipette,
respectively. When a small force, f, is applied to the probe, the capsule is
40

-p

Tm

Azt

Y
Figure 18. Schematic diagram of an ultrasensitive-
tunable force transducer formed by a pressurized
membrane capsule. From [140], with permission.

elongated by a displacement, Az,, proportional to the force. The stiffness


constant, kf for the transducer ( f - kyAz,) is given by:
27t-z"m
kf = ln(2Rp/R o) + ln(2R 0/rb) (34)

r b being the radius of circular contact between capsule and microbead. Since
stiffness is proportional to tension, the force sensitivity can be tuned in
operation between 1 l.tN m -~ and 10 mN m -~ simply by changing P, and Azt is
measurable down to 0.01 ~tm using optical techniques. The microbead is
chemically conjugated to separate ligands, one for macroscopic attachment to
the capsule surface and the other for focal bonding to receptors on a biological
surface. The microbead is brought towards the surface beating the receptor,
bond formation is signalled when fluctuations in height (which depend on the
rigidity of the receptor interface) diminish markedly; similarly bond release
shows up when the fluctuations return to the original level. The relative
frequency of formation and release then quantitates on/off rates. Finally, the
transducer can be retracted to load the force on the attachment until bond
rupture occurs.

5.4 Adhesion and detachment of cells from surfaces: leukocytes


In the last 15 years, much work has been devoted to elucidate the mechanism
whereby leukocytes, in particular polymorphonuclear cells (neutrophils) and
lymphocytes, are able to be arrested on the endothelium of post-capillary
venules and subsequently to migrate out into the extravascular space. The
process appears to occur in three stages: "rolling" of leukocytes along the vessel
wall, followed by firm arrest of the cells, and finally transmigration through
the vessel wall [ 141 ].
41

5.4.1 Rolling vs firm adhesion of leukocytes


A series complex interactions involving fluid mechanical forces and the
formation and breakage of receptor-ligand bonds are involved when circulating
leukocytes adhere to the venular vessel wall [141]. Blood cells flow through
post-capillary venules of diameter from 15-25 lam at mean velocities ranging
from 0.3-1 mm s -~. Wall shear stresses are said to be of the order of 0.2 N m 2.
Normally, leukocytes flowing in close proximity to the wall do not adhere to
the endothelial cells, i.e., no bonds are formed between the leukocyte and an
endothelial cell. However, during inflammation, stimulation of the endothelium
and/or the leukocyte together with the associated low flow state results in the
appearance of more peripherally located white cells [56,84]. Many of these
cells appear to be "rolling" along the vessel wall with translational velocities
from 10-40 ~m s -~ [142-145], markedly lower than predicted for rigid spheres
very close to a plane wall [146]. Rolling is not a smooth process" the
translational velocity varies continually and the cell is frequently arrested.
Some of the rolling and arrested cells rejoin the mainstream, then flowing at
much higher velocities; others become firmly adherent, likely because of
activating stimuli received while in contact with the endothelium. These cells
can later proceed to enter the interendothelial cell junctions and migrate out
[147]. Each of the 3 steps of the cell-wall interaction process is mediated by
different sets of receptor-ligand pairs (adhesion molecules)with different
kinetics of bond formation and rupture.
To elucidate the separate roles of the various adhesion molecules and the
effect of fluid mechanical stress on adhesion, recourse was had to in vitro
experiments in which leukocytes were allowed to roll on, and adhere to
surfaces bearing adhesion molecules in a parallel plate flow chambers [148].
Except near the side walls of the chamber, there is a Poiseuille velocity
distribution between upper and lower surfaces, and the wall shear rate Gw =
3Q/2wh 2, where w is the width and h the height of the chamber.

5.4.2 Leukocyte rolling: selectin mediated adhesion


Since leukocytes, unimpeded by interactions with endothelial cells, likely
flow past the vessel wall at velocities of the order of 10 venule diameters s ~, a
rapid rate of bond formation is necessary to arrest such a cell. As regards
rolling, which involves the continual making and breaking of bonds, the
kinetics of bond formation and dissociation are clearly more important than the
so called bond affinity, i.e., the equilibrium association constant K A = K'o,/n'oy,
where too, (moles s ~) and 1r (s -~) are the on and off rate constants. The ability
of the bonds to withstand high strain before rupture (high "tensile strength";
low value of xt~) will also be important in initial adhesion and rolling, an issue
addressed in two mathematical models of the rolling process [131,149].
The initial slowing down of leukocytes in the bloodstream involves a class of
cell transmembrane adhesion molecules known as selectins (L-selectin on the
42

leukocyte and E - and P-selectin on the endothelial cells). The ligand binding
sites on selectins are calcium-dependent lectin-like domains, carbohydrate
structures which recognize fucose containing oligosaccharide moieties known as
sialyl-Lewis x and sialyl-Lewis ~ on the leukocyte or endothelial cell. In the case
of neutrophils, L-selectin is concentrated on the tips of the microvillus-like
projection of the ruffled cell membrane, favouring the formation of the
selectin-carbohydrate bond with the endothelial cell.
Attempts have been made to determine the parameters of Bell's model from
experiments with rolling cells as well as receptor-linked latex spheres. Alon et
al. [ 150], using neutrophils rolling along lipid bilayers containing P-selectin on
the lower surface of a parallel plate flow chamber, measured the force
dependence of ~;oZ for the P-selectin-carbohydrate ligand bond using the
distribution of times during which cells were arrested. They found xt~ = 0.05
nm, an extremely small value, suggesting the linkage to be close to an "ideal"
bond, the lifetime of which varies little with applied force [131,151 ]. The work
also showed that both too, and tCog~ were fast compared to other known
macromolecular interactions [150]. This may be compared with xa = 0.40 nm,
obtained from computer simulation of the observed force dependence of
rupture of a protein-protein bond between doublets of latex spheres [152].

5.4.3 Leukocyte adhesion mediated by integrins


After a variable period of rolling, if activation of the leukocyte has occurred,
bonds between the activated leukocyte 13/-integrins and their counter-receptors
on the endothelial cell belonging to the immunoglobulin superfamily, can form
after transient cell arrest and eventually induce firm adhesion. Integrins are
heterodimeric cell-surface proteins consisting of one of several o~-subunits and
one of several [3-subunits bound non-covalently. The ~2-integrins, Mac-l,
LFA-1 and VLA-4, are the known leukocyte integrins. The immunoglobulin
superfamily of receptors is defined by the presence of the immunoglobulin
domain, which is composed of 70 amino acids arranged in a well characterized
structure [141]. The immunoglobulin adhesion molecules implicated in
leukocyte-endothelial cell interactions are cellular adhesion molecules ICAM-1
and VCAM-1 and mucosal addressing cell adhesion molecule, MAdCAM-1.
The effect of wall shear stress on leukocyte-endothelial cell adhesion has been
extensively studied. Selectin-mediated rolling followed by integrin-mediated
cell adherence is highest in the range of venular wall shear stress, ~:w, between
0.1 and 0.4 N m -z, but decreases rapidly at higher Tw. By contrast, integrin-
mediated adhesion, in the absence of rolling, leads to cell adhesion only at Vw <
0.1 N m 2 [145,148], emphasizing the necessity of the selectins for successful
cell arrest to occur. The physiological importance of the selectins is underlined
by the recently described leukocyte adherence deficiency 2 (LAD-2) syndrome,
characterized by impaired immune function as well as developmental anomalies
traced to a defective fucose metabolism. The other leukocyte adherence
43

deficiency syndrome, LAD-l, characterized by life-threatening bacterial and


fungal infections, is due to a structural defect in the leukocyte integrins, making
the leukocyte unable migrate through the interendothelial cleft.

5.5 Force dependence of detachment of cells from surfaces


Measuring the detachment of cells from a substrate to which they specifically
bind is the most commonly used technique for evaluating the strength of
adhesion. In the presence of an applied force f per bond, such as generated in
shear flow over the substrate surface, the lifetime tb of a cell-surface bond is
given by Equation (25) [128], and the value of f = fa for instantaneous bond
rupture can be usefully identified with the "strength" of a single bond. A
measure of the adhesion strength can also be inferred from the force
dependence of the time course of the detachment of a population of cells.
Measurements of particle adhesion and detachment have been made using
stagnation point flow chambers - as in the "impinging jet" [153] and "radial
flow detachment assay" [154] methods, in which a narrow confined stream of
suspension exits from an orifice and impinges on a plane surface where it is
observed under a microscope. One can then study the role of hydrodynamic,
colloidal and other forces on deposition and detachment of particles on surfaces
under well-defined fluid flow and mass transfer conditions. A stagnation point
flow chamber was first used with native whole blood to study platelet and
leukocyte deposition in order to test the thrombogenicity of surfaces (155,156).

5.5.1 Impinging jet technique


Dabros and van de Ven [157] first obtained solutions for the flow field and
mass transfer equations in the region of the stagnation point flow. Studies have
since been carried out on the deposition and detachment of latex spheres in
aqueous electrolytes [157], live and fixed E. coli bacteria on glass surfaces
[158], and SSRC on glass covered with monoclonal antibody [138,159].
The coordinates of the stagnation point flow field are shown below in Figure
19. A jet of suspension of radius R impinges on a surface (glass or plastic
coverslip coated with receptor) a distance h away, and is contained between the
two surfaces flowing out through both sides. The flow field is described by:
U~=~z ; Uz=-Yz 2 [35]
where U~ and U~ are the radial and normal components of the velocity field, r
and z being the radial and axial distance from the stagnation point, S, and ~,
defines the strength of the flow field, for which expressions have been given as
a function of Re (153). The wall shear rate on the cover slip, Gw, is given by:
Gw = ~Ur/~z = 'yr; [36]
44

f
I~ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ lI I S ~ \ \ \ \ " ~ \ \ \ \ \ \ \ \ \ ~
i

I I

t . , 4 I 2R - - ~

Figure 19. Stagnation point flow chamber: A jet of suspension of


radius R impinges on a surface a distance h away and is contained
between the two surfaces, flowing out through both sides.

Thus, Gw is zero at the stagnation point and increases linearly with radial
distance. The region where y is c o n s t a n t - a pure stagnation point flow,
extends out to r = 0.1h.. For greater r, corrections to the flow field can be
used. The particles are observed in epi-normal or interference contrast
illumination and events recorded on videotape.
Suspensions of cells or derivatized microspheres beating ligand flow onto the
surface, and the number of particles per unit surface area expressed as:
n, = n= [1 -exp(-t/cr)] [37]
where n is the number at t - oo and cr is given by the relation:
l/or = 1/cr,~ + l/crbt [38]
cr~ being the characteristic escape time of the cells and crb~that of slowing down
of the deposition by the presence of deposited particles [160]. Measurements of
n~ with time in various regions of the surface yield values of the initial adhesion
rate and the above parameters. The escape time (average lifetime of a cell on
the surface) is a measure of the cell-surface bond strength, and can be
compared for the different receptor-ligand systems.
Detachment experiments are carried out with the surface covered with
spheres, and then exposed to a stream of suspending medium of known viscosity
at various flow rates. The tangential hydrodynamic force, F~, exerted on a
particle of radius b adhering to a surface is given by [146]:
F~ = 1.7x6mS;wb 2 = 1.7x6n:yrb 2 [39]
can be related to the surface bond strength through the force dependence of the
time course of detachment, which is dependent on the surface density of ligand.
45

An example is shown in Figure 20 for the detachment of populations of SSRC


from surface-bound antibody using the impinging jet system, under forces from
20 to 100 pN. The maximum number of cross-bridges per cell was estimated
to be 7, and the average rupture shear force per bond to be 17 pN [138].

100~"
~ 50
=
r.~ 80 "~ 40
= E 30
o
= 60 ~ 20
"" "~ 10

E 40
r 1~ 0 20 40 60 80 100 120
" pN

0
0 10 20 30 40 50 60
Time, minutes
Figure 20. Shear-induced detachment of SSRC on a glass surface
having covalently-bound monoclonal IgM. Plot of the normalized
surface density against time. The average tangential hydrodynamic
force was 75 pN. Inset: plot of the surface density vs the average
hydrodynamic force. From [138], with permission.

5.5.2 Radial flow detachment assay


Similar stagnation point flow chambers have been used to study receptor-
mediated cell attachment and detachment in a radial-flow detachment assay
[161,162]. The assay has been used to study the relationship between the
detachment force per bond and the bond affinity, K a [163]. In this device,
detachment was observed at large distance from the stagnation point flow,
where Gw decreases with increasing r. Antigen-coated latex spheres were
allowed to adhere to antibody-coated surfaces under static conditions, and then
exposed to steady shear flow. Spheres detached from the surface up to a
critical radial distance from the stagnation point, corresponding to a "force
required to break the bonds" beyond which the force was too low to detach the
particles. As expected, given the stochastic nature of bond rupture, the contour
of the ring separating adherent from non-adherent spheres, was not a sharp
one. The shear stress to detach 10 ~m latex spheres covalently coated with
rabbit anti-IgG on a surface covalently coated with rabbit IgG corresponded to
46

a force of 3 nN [161], 30x greater than that found in the SSRC-IgM antibody
impinging jet experiments [138]. However, the surface density of rabbit IgG
antibody was 1-3 orders of magnitude greater and the number of bonds, rib, was
thought to be very high.

6. B L O O D C E L L I N T E R A C T I O N S IN R E G I O N S OF D I S T U R B E D
FLOW

Fluid mechanical factors play an important role in the localization of sites of


atherosclerosis, the focal deposition of platelets resulting in thrombosis, and the
formation of aneurysms in the human circulation. The localization is confined
mainly to regions of geometrical irregularity where vessels branch, curve and
change diameter and where blood is subjected to sudden changes in velocity
and/or direction. In such regions, flow is disturbed and separation of
streamlines from the wall with formation of eddies are likely to occur. In this
Section the flow patterns and fluid mechanical stresses at these sites, their effect
on cell and cell-wall interactions are described, and their possible involvement
in the genesis of the above mentioned vascular diseases are considered.

6.1 Model System- Blood cells in an annular vortex


The flow behaviour and interactions of red cells and platelets were studied in
the annular vortex downstream of a sudden tubular expansion of a 150 ktm into
a 500 l.tm diameter glass tube, serving as a model of a region of flow separation
and an arterial stenosis [ 164 - 166].

6.1.1 R e d blood cells


As predicted by fluid mechanical theory [167], when dilute suspensions of
cells were subjected to steady flow through the model stenosis, a captive
annular ring vortex was formed downstream of the expansion. Figure 20
shows paths and orientations of the red cells in the median plane of the tube at
inflow tube Reynolds number, Re0, based on upstream mean fluid velocity,
<U>, and diameter. During a single orbit, the measured particle paths and
velocities, as well as the locations of the vortex center and reattachment point,
were in good agreement with those predicted by the theory applicable to the
fluid [164]. Over longer periods, however, single cells and small aggregates <
20 ktm in diameter migrated outward across the closed streamlines, describing a
series of spiral orbits of continually increasing diameter until they rejoined the
mainstream. In contrast, aggregates of cells > 30 ktm in diameter remained
trapped within the vortex, assuming equilibrium orbits or staying at the center.
In pulsatile flow (a sinusoidal oscillatory flow superimposed in parallel with the
steady flow), the observed phenomena were qualitatively similar to those
described in steady flow. The vortex varied periodically in size and intensity;
47

Re = 37.8 ~

< U > = 0 . 2 3 m s "1

I
i
Q

Figure 21: Single orbits and orientations of glutaraldehyde-hardened


human RBC in the median plane of the annular vortex formed
downstream of the expansion of a 151 gm into a 504 gm diameter tube
at Re0 = 37.8. Arrows indicate location of the reattachment point;
particle velocities were 2.4 (O), 3.6 (A), 3.0 (B), 6.9 (C), 94 (D), 0.84
(E), 0.95 (F) and 272 mm s-1 at (G). From [164], with permission.

the axial location of the vortex center and reattachment point oscillated in phase
with the upstream fluid velocity between maximum and minimum positions
about a mean which corresponded to that in steady flow. At hematocrits from
15 to 45%, migration of cells still occurred and resulted in a lowering of the
vortex hematocrit.
In part, particle migration in the vortex was likely due to the dilution effect
of cell-poor plasma taken into the vortex from the fluid layer adjacent to the
vessel wall upstream of the expansion. The mechanism for the trapping of
large aggregates in the vortex was qualitatively explained by using existing
fluid mechanical theory [57] for lateral particle migration near a tube wall due
to inertia of the fluid, and by the operation of a mechanical wall effect [ 168].

6.1.2 Platelet aggregation in the vortex


The flow behaviour and interactions of human platelets in the annular vortex
were studied in platelet-rich plasma (PRP) and in washed platelet suspensions.
The vortex provided favourable conditions for the spontaneous aggregation of
normal human platelets through repeated shear-induced collisions of cells while
circulating in its orbits [165]. In a given suspension, the formation and growth
of platelet aggregates was observed in a narrow range of Re0. In PRP
48

anticoagulated with heparin, containing significant numbers of platelets with


pseudopods (Figure 12) and microaggregates of 2-6 cells, the rate and extent of
aggregation were greatest, with aggregates > 100 ~m in length forming in <1
minute in the range 4.5 < Re0 < 17. In PRP anticoagulated with sodium citrate,
and in washed platelet suspensions, containing very few microaggregates, the
aggregation was much reduced unless prior to flow, the platelets were activated
with ADP at concentrations below those required to form aggregates in tube
flow [116]. Then, the large aggregates seen in heparin-PRP reappeared. In
pulsatile flow, the number and size of the aggregates in the above suspensions
decreased markedly, presumably due to the continuously changing orbits during
the alternate expansion and contraction of the vortex, which shortened the
particle residence times, as well as to the large variation in the shear rate in
each cycle, beyond the range favourable for platelet aggregation.
Calculations of the two-body collision frequency in various regions of the
vortex for which the sheafing rate of strain was known [165,167], showed that,
at Reo = 12, the total 2-body collision frequency [= JNo/2; Equation (18)] in the
vortex periphery alone would, on average, result in -- 5.4 collisions s -~.
Thus, even for collision efficiencies as low as 10 -2, it can be seen that
appreciable aggregation would occur within a minute of flow.

6.1.3 Wall adhesion of platelets in the vortex


The effects of disturbed flow on initial platelet adhesion to the vessel wall
were studied using an expansion tube whose inner wall was coated with collagen
fibers (to which platelets strongly adhere), and suspensions of washed human
platelets containing washed red cells at 0-50% hematocrit [166]. As illustrated
in Figure 22, platelet adhesion was localized within the vortex and downstream
on either side of the reattachment point with a local minimum at the
reattachment point itself. Furthermore, platelet adhesion increased, and both
adhesion peaks became more pronounced, as the hematocrit increased.
Surprisingly though, the adhesion peak in the vortex decreased and flattened out
as the Reynolds number increased. These results are inconsistent with a
diffusion-controlled platelet adhesion (when the rate determining step in
adhesion is the rate at which cells are brought to the vessel surface), which
should show an increase in adhesion number density with increasing shear rate
[169]. It appears that the particular flow pattern within the vortex is
responsible for this localization. Thus, as illustrated in the upper panel of
Figure 22, only those cells carried by the curved streamlines to within one
particle radius of the surface (black spheres) interact with the vessel wall and
adhere to it on both sides of the reattachment point, which is also a stagnation
point [166,170]. It follows from this that platelet adhesion onto the vessel wall,
whether it be the natural endothelium or an artificial surface, will be localized
wherever there is a stagnation point (or a reattachrnent point if it is a result of
flow separation) where blood cells are carried by the flow toward the vessel
49

9 -:tF ~

'~ 1.0
B Re=
_ 37.5
U= 4 3 . 0 m m sec -!
o 0.8 ! ...... L.. ,... 5 x 1 0 5 p l a t e l e t s ~ e "1
\
,4-,
_.e 0.6
_
-q~ v - OI
o
"~ 0.4 Steady flow
o .......... Pulsatile flow
9 0.2 , AV=3~t
.Q f = 1.5 cycles sec -!
E
Lmin. IL Lmax.
l I; I I I I i i i i i i
Z 0
0 2 4 6 8 10 12
Distance from expansion, mm

Figure 22. Flow patterns and platelet adhesion in the annular vortex.
A: Streamlines in the expansion tube showing convective transport of
spheres on radially directed curved paths on either side of the
reattachment point (dash-dotted line). B: Plot of platelet adhesion
density on the collagen-coated tube in steady and pulsatile flow
(amplitude AV = 3 ~1) at the same mean Re0, with the reattachment point
oscillating between L,,z, and L,,~,. From [166], with permission.

wall along curved streamlines having a pronounced radial velocity component.


If this mechanism operates in the circulation, a relatively higher adhesion of
platelets, and hence a higher risk of thrombus formation is predictable, not only
in regions of disturbed flow (adhesion peaks on either side of the reattachment
point) such as downstream of aortic or venous valves, mural thrombi, and
stenoses, but also in all the branching arteries at the flow divider where there is
a stagnation point.

6.2 In vivo example of an expansion flow: venous valve


The above described studies have been extended to natural blood vessels by
Karino who developed a method for observing flow patterns in transparent
arterial and venous segments from dogs, and arterial segments from humans,
postmortem [171]. The results obtained in an isolated transparent dog
saphenous vein containing a bileaflet valve [172] is shown in Figure 23 with the
detailed flow patterns as observed along the common median plane of vein and
valve. Here, there is an expansion flow with flow separation occurring at the
50

Re=42.1
D=2.03mm
d =0.81 mm
0=53.3mm s-' R

Figure 23. Streamlines in the common median plane of the


valve leaflets in a 2 mm dog saphenous vein showing
formation of a spiral vortex in each valve pocket. Solid
lines are paths of 50 gm polystyrene (PS) tracer spheres
close to, dashed lines are paths far from the median plane; R
is the reattachment point. From [172], with permission.

edge of the valve leaflet which, at physiologically representative flow rates,


resulted in the formation of large paired vortices in each valve pocket, located
symmetrically on both sides of the bisector plane of the valve leaflets. Particles
continually entered the valve pockets from the mainstream, spent up to 10 s
describing a series of spiral orbits of decreasing diameter as they moved away
from the bisector plane, and eventually left the vortex. It was also shown that
another smaller counter-rotating secondary vortex, driven by the large primary
vortex, existed deep in each valve pocket (blank area of Figure 23) where
venous thrombi are believed to originate [173,174], and where it was observed
that cell concentrations were appreciably lower than in the primary vortex. In
such stagnant regions, fluid velocities were extremely low allowing red cells to
aggregate. The results suggest that in some pathological states, the valve-pocket
vortices could act as automatic traps and generators of thrombi in a fashion
similar to that described above in the annular vortex formed downstream from
a sudden tubular expansion.

6.3 A Model System: Flow Patterns at T-Bifurcations


The flow patterns and distributions of fluid velocity and shear rate in glass
models of T-junctions have been described using cinemicrography and frame
by frame analysis of the motions of 50-gm diameter latex spheres [175,176].
51

Branching angles varied from 30 to 150 ~, and the effect of side-to-main-tube


diameter ratio, the branch-to-parent-tube flow ratio, Q~/Qo, and inflow tube
Reynolds numbers, Reo, on flow separation were studied. Figure 24 shows the
flow pattern obtained by in the common median plane of a 90 ~ uniform
diameter T-junction when the main branch was partially occluded so that 80%
of the flow left through the side branch (unlikely to occur in vivo). A large
recirculation zone formed in the main tube consisting of a pair of spiral
secondary flows located symmetrically on both sides of the common median
plane of the T-junction (due to the sudden deceleration of fluid velocity as a
portion of the flow is drawn off into the side tube). A small, secondary
recirculation zone was formed in the side tube (due to the sudden change in
direction of the fluid which continues to move to the outside of the 90 ~ corner,
toward the flow divider). Particles entering the large recirculation zone
described complex orbits; some of them rejoined the flow through the main
tube, others entered the side branch in a paired, spiral secondary flow with
pronounced radial components, and some of these circulated through the
secondary recirculation zone. When the degree of occlusion of the main tube
was gradually reduced, the large recirculation zone became smaller and
eventually disappeared as the flow rate ratio was reversed, while that in the side
branch grew in size.
The critical Reo for the formation of the main recirculation zone was lowest
at 90 ~ for all Q~/Qo, whereas for the side recirculation zone it decreased as the
branching angle increased from 45 to 135 ~. With decreasing side tube
diameter, the main recirculation zone became smaller and was confined to a
thin layer adjacent to the tube wall, wrapped around the mainstream [176].
The effect of radius of curvature of the walls at the junction was studied by
comparing the critical Reo and the sizes of the recirculation zones in the square
T-junction (Figure 24; radii of curvature < 2% of tube radius) with that in a
rounded T-junction (radii of curvature -~ tube radius) [175]. The recirculation
zone in the side tube formed at a much lower Re0 in the square than the
rounded junction, and at a given Re 0 and Q1/Qo, a larger main recirculation
zone existed in the rounded junction. It appears that the formation of the side
recirculation zone is largely affected by curvature of the wall at the bend
opposite to the flow divider while that of the main recirculation zone is largely
affected by the curvature at the flow divider.

6.4 In Vivo Example of T-Bifurcations- Aortic T-Junctions


The flow patterns in transparent segments of a dog abdominal aorta
containing branches of the celiac, superior-mesenteric and right and left renal
arteries have been described [177] The flow patterns illustrated in Figure 25
for the aorto-celiac junction resemble those observed in the model T-junctions,
but the degree of flow disturbance, even at Reo as high as 609, is much smaller.
Instead of a large main standing recirculation zone as in the glass models, there
52

Q1

Qo_~_.
~_~_~,i
9 ~_ ~ ~ ~ / ' ~ f ~ , ~ - - - . . ~ _ _~----------------~__~ ------------___.__~~

.... 4, '~'~"_..,,S~_ ~ ,,.o ,o Q 2

S R

Figure 24. Flow patterns (paths of tracer 50 l.tm PS spheres) in the common
median plane of a 3 mm diameter glass T-junction, made b_y fitting and glueing
2 pieces of tubing. Flow enters at left with mean velocity U o ; 80% left through
the side tube; points show sphere positions at 22 ms intervals, numbers the
velocities in mm s -~. From [175], with permission.

Q2

d/D: 4.0/7.2 Cel~rC~ery


Uo = 2 1 5 mms-' / ~llEtf",tllll
Q,,Qo-O,~ / ~l~', X~tll
/ ~::~,J!,'
/'~
F':t ~ 11
", '~! ]~ I I /
/C,oo~,- t
mesenteric/
,:"ikl, artery
~,,,,r~ ' ........
............................
1.....

Abdominal aorta

Figure 25. Plot, as in Figure 24 of the flow patterns (paths of 200 l-tm PS
spheres) at the aortoceliac artery junction of a dog abdominal aorta, with a
considerably smaller flow disturbance (thin-layered recirculation zone) than in
Figure 24, as only 28% of flow entered the branch, curvature of wall opposite
flow divider is high, that at the divider is low. From [ 177], with permission.
53

was only a pair of recirculation zones, confined to a thin layer close to the wall
surrounding the mainstream. There was no side recirculation zone, no doubt
due to the gentle curvature of the bend opposite the flow divider. This
characteristic was shared by all the aortic T-junctions studied, as was the very
sharp curvature of the bend at the flow divider. From the results described
above in the glass model T-junctions, this represents the optimum condition for
minimizing the size of both regions of separated flow. Nevertheless, the curved
streamlines of the recirculation zone and secondary flows will bring blood cells
towards the vessel wall in a zone around the flow divider and in the side branch
on the outer wall.

6.5 The Human Carotid Bifurcation


The human carotid bifurcation, unlike other Y-bifurcations, exhibits a
marked flow disturbance associated with a large recirculation zone located in
the carotid s i n u s - the vessel expansion at the entry of the internal carotid
artery from the common carotid artery. Figure 26 shows the flow patterns in
the carotid sinus of a transparent arterial segment prepared from a human
subject post-mortem [178]. A standing recirculation zone consisting of a pair
of complex spiral secondary flows, located symmetrically on both sides of the
common median plane of the bifurcation existed over a wide range of Re0, and
Q1/Qo(internal/common carotid). Particles were deflected at the flow divider
and travelled laterally and very slowly along the wall above and below the
common median plane, almost at fight angles to, and encircling the mainstream.
They then changed direction, moving back along the outer wall of the internal
carotid artery at the sinus, describing spiral orbits in the recirculation zone
before rejoining the mainstream. Downstream from the stagnation point (R), a
strong counter-rotating double helicoidal flow developed. The formation and
the size of the recirculation zone were largely dependent on Q~/Qoand on Re0.
The velocity profiles were strongly skewed towards the inner walls of the
bifurcation, creating a high shear field along the vessel wall downstream from
the flow divider. Due to the paired standing recirculation zones in the carotid
sinus, the wall shear rate, and hence the wall shear stress, changes sign and
becomes negative at the separation point (S); it becomes positive again
downstream of the stagnation (reattachment) point (R). Thus, in the carotid
sinus, there is a region where the vessel wall is stretched in opposite directions
by the counter-directed wall shear stresses.
The results suggest that, under physiological conditions a standing, although
pulsating recirculation zone exists in the carotid sinus, thereby affecting local
mass transfer and interactions of blood cells with the vessel wall which may
lead to the incidence of thrombosis and atherosclerosis in this region.
54

Reo--592
U0=266 mm s-' Q2
Q~/Q. =0.613

Qo

i
2a3 :

QI

Figure 26. Flow pattern in the human carotid bifurcation showing


formation of a large recirculation zone and counterrotating double
helicoidal flow in the sinus region of the internal (lower) carotid artery.
Solid lines are paths in or close to, and dashed lines paths far from the
common median plane. Arrows at S and R denote respective locations of
separation and reattachment points. From [178], with permission.

6.6 Flow patterns and atherosclerotic wall thickening


It appears that local flow patterns are involved in the localization of
atherosclerosis, as is illustrated by studies of the middle cerebral bifurcation of
the human circle of Willis at the base of the brain [179]. In each of the five
vessels studied, atherosclerotic thickening of the vessel wall was localized
around the hips of the bifurcation. The detailed flow patterns revealed that a
standing recirculation zone, very similar to that observed in the carotid artery
bifurcation, formed along the outer wall of one or both daughter vessels
(depending on the Re in the parent vessel and Q~/Q2) at the exact locations
where the atherosclerotic thickening of the vessel wall occurred. Furthermore,
under the normal physiological range of flow rates and flow ratios tested, there
was a strong correlation between the longitudinal length of the regions of
disturbed flow and that of the atherosclerotic wall thickening. Figure 27 shows
the detailed flow patterns observed in steady flow in one of the bifurcations
having an almost perfectly symmetric structure and spatial arrangement of the
daughter vessels. As is evident from the figure, even when Q 1 / Q 2 "" 1.0, the
region of disturbed flow was much longer in the right side branch where the
region of atherosclerotic wall thickening was also longer than that in the left
side branch where the wall thickening was confined to only a very narrow area.
55

In pulsatile flow, the complex spiral secondary flows and the recirculation
zones oscillated in phase with the pulsatile flow velocity, and the locations of
the stagnation and separation points moved back and forth along the vessel wall.
However, the general flow patterns remained the same as those observed in
steady flow.

i il I~/
i [i

D, = 3.09 mm
Q.= 147 ml/min

tlillli !I ll|iJ!//iHtli|
~o

Figure 27. Flow patterns at a human middle cerebral artery


bifurcation in the circle of Willis with flow distributed
49%:51% in the branches. The location and lengths of the
secondary flows and recirculation zones at the outer walls of
the bifurcation match those of the atherosclerotic wall
thickening. S and P denote separation and stagnation points.
From [179], with permission.

7. CONCLUDING REMARKS

The rheology of blood, whether at the macroscopic, microscopic or


submicroscopic (molecular) level is indeed a complex subject, as so
graphically and forcibly suggested by G.I. Taylor, years ago. Yet, much of
the work reported in this chapter has contributed significantly towards an
understanding of the rheology at all 3 levels, and this in large measure
through theoretical and experimental studies of the mechanics of the motions
and interactions of blood cells. It should be noted though, that the subject is
much larger than that given here in a chapter whose scope was limited to
aspects of blood cell and blood cell-vessel wall wall interactions. Thus, some
56

recent work had unfortunately to be omitted or only briefly reported.


Examples which come to mind include the motion and wall adhesion of
tumour cells in relation to cancer metastasis [180], the molecular and
biophysical mechanisms of shear stress-induced signal transduction in
circulating and endothelial cells, and flow-induced effects on cell morphology
and function [181,182], the mechanics of cell locomotion [182], and the
mechanics of the motion of RBC in microcirculatory networks, including
physical interactions with the endothelial cell glycocalyx in the capillaries
[24,183]. These are some of the areas of investigation at the cutting edge of
present day research into the biophysics and biorheology of the circulation,
using the latest state of the art technology, a long way from the work of the
pioneers mentioned in the first pages of the chapter.

REFERENCES

0 L.E. Bayliss, In: A. Frey-Wyssling (ed.), Deformation and Flow in


Biological Systems, Interscience, New York, 1952.
. L.E. Bayliss, J. Physiol. (London), 179 (1965) 1.
3. J.R. Pappenheimer, In: Handbook of Physiology, The Cardiovascular
System, sect. 2, Vol. IV, American Physiological Society, Bethesda, MD,
1984.
. S.P. Sutera and R. Skalak, Ann. Rev. Fluid Mech., 25 (1993) 1.
5. R. F~hraeus and T. Lindqvist, Am. J. Physiol., 96 (1931) 562.
6. H.L. Goldsmith, G.R. Cokelet, and P. Gaehtgens, Am. J. Physiol, 89
(1989) H1005.
. L. Dintenfass, Angiology, 8 (1962) 333.
8. E.A. Evans, Biophys. J., 13 (1973) 926; ibid, idem (1973) 941..
9. R.M. Hochmuth, In: R. Skalak and S. Chien (eds.), Handbook of
Bioengineering, McGraw Hill, New York, 1987.
10. G.V.F. Seaman, In: D. MacN. Surgenor (ed.), The Red Blood Cell, II,
Academic, New York, 1975.
11. S. Levine, M. Levine, K.A. Sharp and D.E. Brooks, Biophys. J., 42
(1983) 127.
12. B.V. Derjaguin and L.D. Landau, Acta Physiochemica URSS, 14 (1941)
633.
13. E.G. Verwey and Th.G. Overbeek, Theory of the Stability of Lyophobic
Colloids, Elsevier Scientific, Amsterdam, 1948.
14. D.E. Brooks, R.G. Greig and J. Janzen, In: H. J. Meiselman and D.E.
Brooks (eds.), Erythrocyte Mechanics and Blood Flow, Alan R. Liss, New
York, 1980.
15. D.E. Brooks, J. Colloid Interface Sci., 43 (1973) 700.
16. S. Chien and K.-M. Jan, Microvasc. Res., 5 (1973) 155.
57

17. E.A. Evans and D. Needham, Macromolecules, 21 (1988) 1822.


18. W.R. Milnor, Hemodynamics, 2nd ed., Williams & Wilkins, Baltimore,
1989.
19. W.W. Nichols and F.M. O'Rourke (eds.), McDonald's Blood Flow in
Arteries, Lea and Febiger, Philadelphia, 1990.
20. M. Anliker, M. Casty, P. Friedli, R. Kubli and H. Keller, In: N.H.C.
Hwang and N.A. Nomaann (eds.), Cardiovascular Flow Dynamics and
Measurements, University Park, Baltimore, 1977.
21. J.W. Pothero and A. C. Burton, Biophys. J., 1 (1961) 565.
22. U. Bagge, P.-I. Brhnemark, R. Karlsson and R. Skalak, Blood Cells, 6
(1980) 231.
23. P. Gaehtgens, C. Duehrssen and K.H. Albrecht, Blood Cells, 6 (1980)
799.
24. T.W Secomb, In: C.P. Ellington and T.J. Pedley (eds.), Biological Fluid
Dynamics, Company of Biologists, Cambridge, U.K., 1995.
25. U. Bagge and P.-I. Brgmemark, Adv. Microcirc., 7 (1977) 1.
26. G.W. Schmid-Schtinbein, S. Usami, R. Skalak and S. Chien., Microvasc.
Res., 19 (1980) 45.
27. T. Karino, M. Motomiya and H.L. Goldsmith, In: J.C. Stanley (ed.),
Biologic and Synthetic Vascular Prostheses, Grune & Stratton, New York,
1982.
28. P.D. Stein and H.N. Sabbah, Circ. Res., 398 (1976) 58.
29. S. Chien, Science, 168 (1970) 977.
30. P.L. Easthope and D.E. Brooks, Biorheology, 17 (1980) 235.
31. F.J. Walbum and D.J. Schneck, Biorheology, 13 (1976) 201.
32. N. Casson, In: C.C. Mills (ed.), Rheology of Disperse Systems, Pergamon,
New York, 1959.
33. G.W. Scott-Blair, Nature, 183 (1959) 613.
34. G.R. Cokelet, In: Biomechanics: Its Foundation and Objectives, Prentice
Hall, Englewood Cliffs, NJ, 1972.
35. D. Quemada, Rheol. Acta, 16 (1978) 82.
36. D. Quemada, Rheol. Acta, 17 (1978) 632.
37. G.R. Cokelet, In: R. Skalak and S. Chien (eds.), Handbook of
Bioengineering, McGraw Hill, New York, 1987.
38. S. Chien, In: D. MacN. Surgenor (ed.), The Red Blood Cell, II,
Academic, New York, 1975.
39. W. Reinke, P. Gaehtgens and P.C. Johnson, Am. J. Physiol., 253 (1987)
H540.
40. G.R. Cokelet and H.L. Goldsmith, Circ. Res., 68 (1991) 1.
41. E.B. Vadas, H.L. Goldsmith, and S.G. Mason, J. Colloid Interface Sci., 43
(1973) 630.
42. S.P. Tha, J. Shuster, and H.L. Goldsmith, B iophys. J., 50 (1986) 1117.
58

43. H.L. Goldsmith and J. Marlow, Proc. Roy. Soc. (London), B 181 (1972)
351.
44. H.L. Goldsmith and S.G. Mason, J. Fluid Mech., 12 (1962) 88.
45. H.L. Goldsmith and S.G. Mason, J. Colloid Sci., 17 (1962) 448.
46. G.B. Jeffery, Proc. Roy. Soc. (London), A102 (1922) 162.
47. H.L. Goldsmith and S.G. Mason, In: F.R. Eirich (ed.) Rheology: Theory
and Applications, Vol. IV, Academic, New York, 1967.
48. M.M. Frojmovic, M. Newton and H.L. Goldsmith, Microvasc. Res., 11
(1976) 203.
49. E.B. Vadas, R.G. Cox, H.L. Goldsmith and S.G. Mason, J. Colloid
Interface Sci., 57(1976) 308.
50. F.D. Rumscheidt and S.G. Mason, J. Colloid Sci., 6 (1961) 238.
51. T.M. Fischer, M. Strhr-Liesen and H. Schmid-Schrnbein, Science, 202
(1978) 894.
52. S.R. Keller, and R. Skalak, J. Fluid Mech., 120 (1982) 27.
53. H. Brenner, In: T.B. Drew, J.W. Hoopes and T. Vermeulen (eds.),
Advances in Chemical Engineering, Vol. 6, Academic, New York, 1966.
54. L.G. Leal, Ann. Rev. Fluid Mech., 12 (1980) 435.
55. H.L. Goldsmith, Fed. Proc., 30 (1971) 1578.
56. H.L. Goldsmith and S. Spain, Microvasc. Res., 27 (1984) 204
57. R.G. Cox and S.K. Hsu, Int. J. Multiphase Flow, 3 (1977) 201.
58. S. Segr6 and A. Silberberg, J. Fluid Mech., 14 (1962) 136.
59. H.L. Goldsmith, and J.C. Marlow, J. Colloid Interface Sci., 71 (1979)
383.
60. H.L. Goldsmith, Fed. Proc., 26 (1967) 1813.
61. H.L. Goldsmith and R. Skalak, Ann. Rev. Fluid. Mech., 7 (1975) 213.
62. A. Karnis, H.L. Goldsmith, and S.G. Mason, J. Colloid Interface Sci., 22
(1966) 531.
63. J.C. Slattery, J. Fluid Mech., 19 (1964) 625.
64. R. Skalak, Biorheology, 29 (1992) 479.
65. D. Leighton and A. Acrivos, J. Fluid Mech., 181 (1987) 415.
66. R.J. Phillips, R.C. Armstrong, R.A. Brown, A.L. Graham and J.R.A.
Abbott, Phys. Fluids, 4 (1992) 30.
67. P.R. Nott and J.F. Brady, J. Fluid Mech., 275 (1994) 155.
68. C.J. Koh, P. Hookham and G. Leal, J. Fluid Mech., 266 (1994) 1.
69. N. Phan-Tien, A.L. Graham, S.A. Altobelli, J.R. Abbott and L.A. Mondy,
Ind. Eng. Chem. Res., 34 (1995) 3187.
70. H.L. Goldsmith, In: T.H. Spaet (ed.), Progress in Hemostasis and
Thrombosis, Vol. 1, Grune & Stratton, New York, 1972.
71. R.L. Replogle, H.J. Meiselman and E.W. Merrill, Circulation, 36 (1967)
148.
72. F.P. Gauthier, H.L. Goldsmith and S.G. Mason, Biorheology, 9 (1972)
205.
59

73. E.B. Vadas, H.L. Goldsmith and S.G. Mason, Trans. Soc. Rheol., 20
(1976) 373.
74. H.L. Goldsmith and T. Karino, Ann. N.Y. Acad. Sci., 283 (1977) 241.
75. H. Brenner, Int. J. Multiphase Flow, 1 (1974) 195.
76. V.T. Turitto, A.M. Benis and E.F. Leonard, Ind. Eng. Chem. Fund., 11
(1972) 216.
77. E.F. Grabowski, L.I. Friedman and E.F. Leonard, Ind. Eng. Chem. Fund.
11 (1972) 224.
78. I.A. Feuerstein, B.M. B rophy and J.L. Brash, Trans. Am. Soc. Artif.
Intern. Org., 21 (1975) 21.
79. V.T. Turitto and H. Baumgartner, Microvasc. Res., 9 (1975) 335.
80. V.T. Turitto and H.J. Weiss, Science, 207 (1980) 541.
81. P.A. Aarts, K.S. Bolhuis, R.M. Sakariassen, R.M. Heethar and J.J. Sixma,
Blood, 62 (1983) 214.
82. A.L. Zydney and C.K. Colton, Physicochem. Hydrodyn., 10 (1988) 77.
83. R. Fhhraeus, Physiol. Rev., 9 (1929) 241.
84. U. Nobis, A.R. Pries and P. Gaehtgens, In: U.Bagge, G.V.R. Born and P.
Gaehtgens (eds.) White Blood Cells: Morphology and Rheology as Related
to Function, Martinu Nijhoff, The Hague/Boston, 1982.
85. H.L. Goldsmith and S. Spain, In: H.J. Meiselman and M. Lichtman, (eds.),
White Cell Mechanics: Basic Science and Clinical Aspects, Raven, New
York, 1984.
86. A. Karnis, H.L. Goldsmith and S.G. Mason, Can. J. Chem. Eng., 44
(1966) 181.
87. G.J. Tangelder, H.C. Teirlinck, D.W. Slaaf and R.S. Reneman, Am. J.
Physiol., 248 (1985) H318.
88. P.A. Aarts, S.A.T. van den Broek, G.W. Prins, G.D.C. Kuiken, J.J.
Sixma and R.M. Heethar, Arteriosclerosis, 8 (1988) 819.
89. A.W. Tilles and E.C. Eckstein, Microvasc. Res., 33 (1987) 211.
90. E.C. Eckstein, A.W. Tilles and F.J. Millero, Microvasc. Res., 36 (1989)
31.
91. E.C. Eckstein and F. Belgacem, Biophys. J., 60 (1990) 53.
92. F. Seitz, Phys. Rev., 73 (1948) 1513.
93. D.F.J. Tees, O. Coenen and H.L. Goldsmith, B iophys. J., 65 (1993) 1318.
94. H.L. Goldsmith and S.G. Mason, Proc. Roy. Soc. (London), A 282 (1964)
569.
95. T.G.M. van de Ven and S.G. Mason, J. Colloid Interface Sci., 57 (1976)
505.
96. K. Takamura, H.L. Goldsmith and S.G. Mason, J. Colloid Interface Sci.,
82 (1981) 175.
97. P.A. Arp and S.G. Mason, J. Colloid Interface Sci., 61 (1977) 21.
98 C.Y. Lin, K.Y. Lee and N.F. Sather, J. Fluid Mech., 56 (1972) 375.
99. G.K. Batchelor and J.T. Green, J. Fluid. Mech., 56 (1972) 375.
60

100. P.M. Adler, J. Colloid Interface Sci., 84 (1981) 461.


101. T.G.M. van de Ven, Colloidal Hydrodynamics, Academic, San Diego,
1989.
102. M. von Smoluchowski, Z. Phys. Chem., 92 (1917) 129.
103. P. Arp and S.G. Mason, Can. J. Chem., 54 (1976) 3769.
104. T.G.M. van de Ven and S.G. Mason, Colloid Polym. Sci., 225 (1977) 468.
105. P.M. Adler, J. Colloid. Interface Sci., 81 (1981) 489.
106. K. Takamura, H.L. Goldsmith and S.G. Mason, J. Colloid Interface Sci.,
72 (1979) 385.
107. K. Takamura, H.L. Goldsmith and S.G. Mason, J. Colloid Interface Sci.,
82 (1981) 190.
108. P. Davies, Thromb. Haemost., 70 (1993) 128.
109. S.R.P. Gudi and J.A. Frangos, In: V.C. Mow, F. Guilak, R. Tran-Son-Tay
and R.M. Hochmuth (eds.), Cell Mechanics and Cellular Engineering,
Springer, New York, 1994.
110. D.E. Ingber, In: V.C. Mow, F. Guilak, R. Tran-Son-Tay and R.M.
Hochmuth (eds.), Cell Mechanics and Cellular Engineering, Springer,
New York, 1994.
111. R. Skalak, B.A. Skierczynski, S. Usarni and S. Chien, In: V.C. Mow, F.
Guilak, R. Tran-Son-Tay and R.M. Hochmuth (eds.), Cell Mechanics and
Cellular Engineering, Springer, New York, 1994.
112. D.N. Bell, S., Spain and H.L. Goldsmith, Biophys. J., 56 (1989) 817.
113. D.R. Phillips, J.F. Charo, L.V. Parise and L.A. Fitzgerald, Blood, 55
(1988) 831.
114. P.L. Sung, M.M. Frojmovic, T.E. O'Toole, E.F. Plow, C. Zhu, M.
Ginsberg and S. Chien, Blood, 81 (1993) 419.
115. M. Frojmovic, T. Wong and Th. van de Ven, Biophys. J., 59 (1991) 815.
116. D.N. Bell, S., Spain and H.L. Goldsmith, Biophys. J., 56 (1989) 829.
117. M.M. Frojmovic, K.A. Longmire and T.G.M. van de Ven. Biophys. J., 58
(1990) 309.
118. D.N. Bell, S., Spain and H.L. Goldsmith, Thromb. Haemost., 63 (1990)
112.
119. H.L. Goldsmith, D.N. Bell, S. Braovac, A. Steinberg and F. McIntosh,
Biophys. J., 69 (1584) 1995.
120. T.K. Belval and J.D. Hellums, Biophys. J., 50 (1986) 479.
121. R.L. Drake, In: G.M. Hidy and J.R. Brock (eds.), Topics in Current
Aerosol Research, part 2, Pergamon, Oxford, 1972.
122. A.D. Taylor, S. Neelameghan, J.D. Hellums, C.W. Smith and S.I. Simon,
Biophys. J., 71 (1996) 3488.
123. S. Neelameghan, A.D.Taylor, J.D. Hellums, M. Dembo, C.W. Smith and
S.I. Simon, Biophys. J., 72 (1997) 1527.
124. E.-L. Florin, V.T. Moy and H.E. Gaub, Science, 254 (1994) 415.
125. G.U. Lee, D.A. Kidwell and R.J. Colton, Langmuir, 10 (1994) 354.
61

126. E. Evans, K. Ritchie and R. Merkel, B iophys. J., 68 (1995) 2580.


127. S.C. Kuo and M.P. Sheetz, Science, 260 (1993) 232.
128. G.I. Bell, Science, 200 (1978) 618.
129. S.N. Zhurkov, Int. J. Fract. Mech., 1 (1965) 311.
130. H.A. Kramers, Physica., 7 (1941) 284.
131. D.A. Hammer and S.M. Apte, Biophys. J., 63 (1992) 35.
132. E. Evans, D. Berk and A. Leung, Biophys. J., 59 (1991) 838.
133. E. Evans, In: R. Lipowsky and E. Sackmann (eds.), Handbook of Physics
of Biological Systems, Vol. 1, Elsevier Science, Amsterdam, 1995.
134. E. Evans and K. Ritchie, Biophys. J., 72 (1997): 1541.
135. H.L. Goldsmith, O. Lichtarge, M. Tessier-Lavigne and S. Spain,
Biorheology, 18 (1981) 531.
136. H. Brenner and M.E. O'Neill, Chem. Eng. Sci., 27 (1972) 1421.
137. S. Tha and H.L. Goldsmith, Biophys. J., 50 (1987) 1109.
138. Z. Xia, H.L. Goldsmith and T.G.M. van de Ven, Biophys. J., 66 (1994)
1222.
139. D.F.J. Tees and H.L. Goldsmith, Biophys. J., 71 (1996) 1102.
140. E. Evans, R. Merkel, R. Ritchie, S. Tha and A. Zilker, In: P. Bongrand,
P.M. Claesson and A.S.G. Curtis (eds.), Studying Cell Adhesion,
Springer, Berlin, Heidelberg, New York, 1994.
141. T.A. Springer, Cell 76 (1994) 301.
142. D.A. Jones, C.W. Smith and L.V. Mclntire, Biomaterials, 17 (1996) 337.
143. A. Atherton and G.V.R. Born, J. Physiol. (Lond.), 233 (1973) 157.
144. G.W. Schmid-Schrnbein, R. Skalak, S.I. Simon and R.L. Engler. Ann.
N.Y. Acad. Sci., 516 (1987) 348.
145. M.B. Lawrence and T.A. Springer, Cell, 65 (1991) 859.
146. A.J. Goldman, R.G. Cox and H. Brenner, Chem. Eng. Sci., 22 (1967) 22.
147. C.W. Smith, Can. J. Physiol. Pharmacol., 71 (1993) 76.
148. M.B. Lawrence, L.V. Mclntire and S.G. Eskin, Blood, 70 (1987) 1284.
149. A. Tozeren and K. Ley, Biophys. J., 63 (1992) 700.
150. R. Alon, D.A. Hammer and T.A. Springer, Nature (1995) 539.
151. M. Dembo, D.C. Torney, K. Saxman and D. Hammer, Proc. Roy. Soc.
Lond. B. Biol. Sci. 234 (1994) 55.
152. D. Kwong, D.F.J. Tees and H.L. Goldsmith, Biophys. J., 71 (1996) 1115.
153. T. Dabros, T. and T.G.M. van de Ven, Colloid Polymer Sci., 261 (1983)
694.
154. C. Cozens-Roberts, D.A. Lauffenburger and J.A. Quinn, Biophys. J., 58
(1990) 107.
155. H.E. Petschek, D. Adamis and A.R. Kantrowitz, A.R, Trans. Am. Soc.
Artif. Int. Organs, 14 (1968) 256.
156. E. Nyilas, W.A. Morton, D.M. Lederman, T.-H. Chiu. and R.D.
Cumming, Trans. Am. Soc. Artif. Int. Organs, 21 (1975) 55.
62

157. T. Dabros and T.G.M. van de Ven, PhysicoChem. Hydrodyn., 8 (1987)


161.
158. Z. Xia, L. Woo and T.G.M. van de Ven, Biorheology, 26 (1989) 359.
159. Z. Xia, H.L. Goldsmith and T.G.M. van de Ven, Biophys. J., 65 (1993)
1073.
160. T. Dabros and T.G.M. van de Ven, J. Colloid Interface Sci., 93 (1983)
576.
161. C. Cozens-Roberts, D.A. Lauffenburger and J.A. Quinn, Biophys. J., 58
(1990) 857.
162. A. Saterback, S.C. Kuo and D.A. Lauffenburger, Biophys. J., 65 (1993)
243.
163. S.C. Kuo and D.A. Lauffenburger, Biophys. J., 65 (1993) 2191.
164. T. Karino and H.L. Goldsmith, Phil. Trans. Roy. Soc. Lond. Biol. Sci.,
279 (1977) 413.
165. T. Karino and H.L. Goldsmith, Microvasc. Res., 17 (1979) 217.
166. T. Karino and H.L. Goldsmith, Microvasc. Res., 17 (1979) 238.
167. E.O. Macagno and T.K. Hung, J. Fluid Mech., 28 (1967) 43.
168. A. Kamis and S.G. Mason, J. Colloid Interface Sci., 23 (1967) 120.
169. V.T. Turitto, In: T.H. Spaet (ed.), Progress in Hemostasis and
Thrombosis, Vol. 6, Grune & Stratton, New York, 1982.
170. T. Karino and H.L. Goldsmith, Biorheology, 21 (1984) 587.
171. T. Karino and M. Motomiya, Biorheology, 20 (1983) 119.
172. T. Karino and M. Motomiya, Thromb. Res., 36 (1984) 245.
173. L. Diener, J.L.E. Ericsson and F. Lund, In: T. Shimamoto and F.
Numano (eds.), Atherogenesis, Excerpta Medica, Amsterdam, 1969.
174. S. Sevitt, In: J.J. Bergan and J.S.T. Yao, (eds.), Venous Problems, Year
Book Medical, Chicago, 1978.
175. T. Karino, H.H.M. Kwong and H.L. Goldsmith, Biorheology, 16 (1979)
231.
176. T. Karino and H.L. Goldsmith, Biorheology, 22 (1985) 87.
177. T. Karino, M. Motomiya and H.L. Goldsmith, J. Biomech., 23 (1990)
537.
178. M. Motomiya and T. Karino, Stroke, 15 (1984) 50.
179 T. Karino, Intern. Angiology, 5 (1986) 297.
180. A.F. Chambers, I. MacDonald, E. Schmidt, S. Koop, V.L. Morris, R.
Khokha and A.C. Groom, Cancer Metastasis Rev., 14 (1995) 279.
181. J.A. Frangos (ed.), Physical Forces and the Mammalian Cell, Academic,
San Diego, 1993.
182. V.C. Mow, F. Guilak, R. Tran-Son-Tay and R.M. Hochmuth (eds.),
Cellular Mechanics and Cellular Engineering, New York, Springer, 1994.
183. A.R. Pries, T.W. Secomb and P. Gaehtgens, Cardiovasc. Res., 32 (1996)
654.
63

S H E A R T H I C K E N I N G AND FLOW I N D U C E D S T R U C T U R E S
IN FOODS AND B I O P O L Y M E R SYSTEMS

E. B. Bagley" and F. R. Dintzis b

aDept, of Food Science, University of Illinois, Urbana, IL;


address for mail." 756 S. Columbus, Morton, IL 61550-2428

bUSDA, ARS, National Center for Agricultural Utilization Research,


Peoria, IL 61604

1. INTRODUCTION

The handling of materials such as foods and food components, synthetic


polymers and biopolymers, and the combining and/or transformation of these
materials into useful products, normally involve complex industrial processes
such as mixing, flow into and through pipes, extrusion, injection molding. To
predict material behavior relevant to processing and to quantify or model
behavior, rheological data are vital. Experience over the years with synthetic
polymers has shown that such rheological data can also relate to molecular
factors such as molecular weight, molecular weight distribution, and chain
branching, and to the final finished product properties (Graessley [1]; Bagley
and Schreiber [2]). Useful relationships among processing history, molecular
structure, molecular aggregation phenomena, and end use applications, have
been developed over the years for synthetic high polymers such as
polyethylene and polystyrene.
Attention should be drawn to two recent books, one of which emphasizes
flow-induced structure in polymer systems (Nakatani and Dadmun [3]) and the
other deals with the rheo-physics of multiphase polymer systems (Sondergaard
and Lyngaae-Jorgensen [4]). There is no doubt that many biopolymers and
foods can be considered as multiphase polymer systems and some of the
chapters in Sondergaard and Lyngaae-Jorgensen dealing with phase transitions
in shear flow, on flow-induced dissolution and crystallization in polymer
systems, and flow-induced phase changes in polymer blends can be especially
relevant to the phenomenon of shear thickening in biopolymer and food
systems.
64

The simplest rheological behavior is that shown by relatively small


molecules for which the viscosity is independent of shear rate. The next level
of complexity is evident in the flow behavior of many synthetic polymers,
both as melts and in solution, where at low shear rate the materials show
Newtonian behavior, with this Newtonian viscosity level depending on the 3.4
power of the weight average molecular weight, above a certain critical
molecular weight, of the polymer (for example, see discussion by Graessley
[1]). As the shear rate is increased a shear rate is reached beyond which the
steady state viscosity can begin to decrease. The details of the flow curve,
such as the shear rate at which this shear thinning begins, and the degree or
extent of the shear thinning, depend on molecular factors such as molecular
weight, molecular weight distribution, and on the amount and type of chain
branching. For solutions, concentration plays a significant role in the
transition from Newtonian to shear thinning behavior as illustrated in Figure
1, taken from Launay and McKenna [5].

E
10 D

12_

10 -1

10-2

I I I I I I
r

10 -~ 1 10 10 2 10 3 10 4
?(S -1 )

Figure 1. Flow curves at 25 ~ C of locust bean gum solutions of 0.25%, 0.5%,


0.8%, 1.5%, and 1.8% concentrations (A to E, respectively). The flow is
Newtonian at the lower concentrations, curves A, B, and C, at the lower shear
rates. (From Launay and McKenna [5], by permission).
65

In addition to relating results to molecular factors the rheological results can


be used in analyzing behavior during processing. For this it is essential to
have the rheological data available for the shear rate ranges encountered
during processing. Note also that these non-Newtonian materials are usually
viscoelastic which give rise to important processing effects such as die swell
during extrusion through pipes. The die swell is given (for cylindrical
extrudates) as the ratio of the extrudate diameter to the die diameter and easily
can be, for polyethylene, of the order of 3 to 5! The die swell is often
correlated with commercially significant properties such as product gloss or
matte. In addition, the relaxation times associated with these viscoelastic
materials can give rise to significant time effects such as the complex
transients observed as stress overshoots. Time effects can play an important
role in the shear thickening phenomenon.
Food and biopolymer systems can show all of the rheological phenomena
described in the above paragraph. However, whereas many uncharged
synthetic polymers show little intermolecular interactions beyond the chain
entanglements often associated with materials of high molecular weight, food
and biopolymer systems have three additional sources of rheological
complexity. First, biopolymers are often of very high molecular weight
indeed. Amylopectin, the branched component of starch, can have molecular
weights above a hundred million, corresponding to perhaps 600,000 or more
anhydroglucose units in the molecule. In terms of degree of polymerization
this is far beyond the extent of polymerization of most common synthetic
polymers such as commercial polyethylenes or polystyrenes. The massive size
of this starch molecule component makes it particularly susceptible to shear
degradation which gives rise to a number of experimental results not normally
seen in polymer systems.
Second, food systems and biopolymers in general contain a variety of
molecular groups such as hydrogen bonds and ionic groups, which foster inter-
and intra-molecular interactions which can lead to a variety of complex
phenomena influencing both processability and final product properties. The
effect of these interacting units can be modified or amplified by additives of
various kinds which again can contribute to unusual but significant effects, of
value and interest both in practical and scientific terms.
Finally, and this aspect of biopolymer systems has not been extensively
discussed, it seems reasonable to expect that in the biosynthesis of these
polymers, where the polymer is "laid down" as a substrate in biosynthesis
during plant growth, the molecules may be intimately entangled, more so than
would be expected for synthetic polymers. This might well result in
entanglements of an especially effective kind, particularly when the molecular
weight is enormous.
66

These food and biopolymer systems, in addition to showing shear thinning


effects, can also, under specific circumstances, show shear thickening
behavior. Shear thickening behavior seems to be relatively rare and the
interpretation of apparent thickening can be complicated by a variety of
effects. Thickening is often associated with shear induced structure formation,
but often the details of, and the mechanism for, the structure formation are
unclear. Nevertheless, the processes leading to these shear thickening effects
are of great interest and are significant and need to be carefully considered in
terms of processing and process design as well as in interpretation of data
used for material characterization.
Shear thickening occurs for what might superficially be considered
comparatively simple systems such as dispersions of rigid spherical particles.
Shear thickening in such systems is often considered a reflection of volume
increase under shear, or dilatancy. As Tanner [6] discusses, this concept of
dilatancy goes back to the analysis of the flow of concentrated suspensions of
solids by Osborne Reynolds, but Tanner emphasizes that in recent years "The
term dilatant has come to be used for all fluids which exhibit the property of
increasing viscosity with increasing rates of shear." This is an unfortunate
development since it implies a mechanism for shear thickening which may
well not be correct for many, if not most, shear thickening systems. That is,
shear thickening can occur without a volume change, so that this misleading
usage of the term dilatant should be discontinued unless a volume change is
proved.
Shear thickening is thus a phenomenon associated with complicated
interacting systems. These interactions may arise from specific molecular
factors, such as hydrogen bonding, or physical factors such as chain
entanglement or particle crowding (in particulate suspensions or dispersions).
These effects can be exacerbated during flow because of the formation of
shear induced structures of various types. When shear thickening occurs in
systems such as suspensions of rigid spheres, where geometrical considerations
come into play, the shear thickening occurs concomitantly with system volume
expansion, leading to the concept of dilatant flow. These systems can also be
considered to produce shear induced structures.
The objective of this chapter is to review first of all rheological
measurements and theory related to shear thickening effects. Then a variety
of shear thickening systems will be discussed with particular emphasis on food
and biopolymer related systems. Special attention will be given to the need
for a careful distinction between shear thickening alone and shear thickening
accompanied by dilatancy. We wish to point out an unexpected problem in
using on line computer searches to review literature relevant to this chapter.
Use of the search term, "biopolymer," in DIALOG searches of Chemical
67

Abstracts (American Chemical Society) did NOT result in listing of our own
publications pertaining to starch, although use of the search term, "starch," did.
This indexing situation was most surprising to us.

2. R H E O L O G I C A L M E A S U R E M E N T S AND R E L A T E D T H E O R Y

In the simplest sense a shear-thickening fluid would be considered as one in


which the viscosity increases with increasing shear rate. For such a fluid a
plot of shear stress versus shear rate would curve upward as shown in Figure
2a. Such plots would typically be obtained using a viscometer such as a
Couette bob-in-cup instrument or a cone-and-plate viscometer. (Rheological
instrumentation has been reviewed in numerous publications (Dealy [7]) and
there are available useful commercial instrumentation brochures from
companies such as Haake, Physica, and others.) When shear thickening is
observed either in plots such as viscosity versus shear rate (Figure 2b) or shear
stress versus shear rate (Figure 2a) there is an implicit assumption that the

b
Shear thickening Shear Shear
thickening
thinning
J
/ / / Newtonian

~ thinning
Newtonian

? ?

Figure 2. a. Schematic representation of flow behaviors on a shear stress (z)


versus shear rate (?) plot.
b. Schematic representation of viscosity behaviors on a viscosity (q)
versus shear rate (?) plot.

plotted values of shear stress, shear rate and viscosity are steady state values.
This may not be the case and the appearance of shear thickening may be a
result of transient effects. Such effects are particularly likely with
instrumentation in which shear sweeps are performed, for example in a
Couette rheometer in which the shear rate may be cycled over a certain range
in a certain time. In a recent study of effects of processing on starch solution
properties, shear rate sweeps from zero to 750 reciprocal seconds and back
68

down to zero were carried out over a four minute period (Dimzis and Bagley
[8]). In such experiments the possibility of transient effects cannot be ignored.
Such a flow history is sometimes referred to as the thixotropic loop or T loop
(see Greener and Connelly [9]).
Other instrumental factors can be significant in the phenomena of shear
thickening. For example, a recent contribution by Chow and Zukoski [10]
deals with the effect of gap size in shear thickening of a suspension. The
possibility of such experimental artifacts must be kept firmly in mind in
interpreting complex rheological data.
Fluids for which viscosity depends not only on shear rate but also on time
can be subdivided into two classes, thixotropic and rheopectic (Tanner [6]).
Thixotropic fluids are generally regarded as ones in which there is a
breakdown of structure by shear while rheopectic fluids are ones (Tanner [6])
in which there is formation of structure due to shear. The history of the term
"thixotropy" has been reviewed briefly by Cheng and Evans [ 11 ] but they use
the additional term "antithixotropic" because in their view there are three
distinct types of structural changes brought about by shear: l) structural
breakdown under shear with recovery at rest (thixotropy); 2) structure buildup
under shear, disintegrating at rest (anti-thixotropy or negative thixotropy); 3)
structure breakdown at moderate and high rates of shear with recovery
accelerated at low shear rates, the recovered structure being stable at rest
(rheopexy).
The difficulty with the introduction of new terminology to describe particular
types of behavior is that nature keeps providing us with new phenomena
which do not fit into the accepted categories of the times. Thus, starch
systems (as will be discussed in more detail below) can show structure buildup
under shear but the structure so formed appears to be quite stable, both at rest
and under different shear fields (Dintzis and Bagley [8]). This type of
behavior is not covered by the three classifications proposed by Cheng and
Evans. Further, when the classifications become too numerous it often
happens that different workers use the terminology in different ways. A good
example of such confusion was given by Cheng and Evans, who noted in
their 1965 paper "At the present time thixotropy is still used by certain
rheologists to denote what is more commonly known as pseudoplasticity (that
is, the decrease in viscosity with increase in shear rate) without reference to
time dependence."
Barnes has recently reviewed shear thickening effects for nonaggregating
solid particle dispersions in Newtonian liquids (Barnes [12]). He summarizes
his views in the abstract of his article as follows" "The actual nature of the
shear thickening will depend on the parameters of the suspended phase: phase
volume, particle size and distribution, particle shape, as well as those of the
69

suspending phase (viscosity and the details of the deformation, i.e., shear or
extensional flow, steady or transient, time and rate of deformation)." Barnes
notes that shear thickening is defined in the British Standard Rheological
Nomenclature as "the increase in viscosity with increase in shear rate, and is
to be distinguished from rheopexy which is the increase of viscosity with time
at constant shear rate." Barnes also makes some interesting comments on, and
gives some references to, the practical implications of shear thickening.

60

.,..:::...:'.. : .........
50 , | ,,.. m" i~i =

40 ................................... :.,'" : . ~ ................................


(~ W a x y m a ! z e s t a r c.=.h ! " '= ' -" =iP ram'.

CO
L~ ="" m""
30 . . . . . . . . . . . . . . . . . . . . . . . . ~..-':-'-'"- ;,.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

(1)
(-" "
,~, ~"",:
~r 9 ,

co 20 ......... :::: ................ : .................................................


"":' 1 .'A
" B:uffalo starch : ~..~,~,,,,~
- .." i .,? i i
1 0 - ; . . . . . . . !;, . . . . . . . ~. . . . . ~ , ~ . ~ ~ ~ ' ~ - - ..... ! ......... ! ......... ! .........
. 9 m m= , ~ . . . . .

,-" ,~0~,,,~
~ : : : : : :
n,~'.~., i.., i,.. i.., i... i,,, i,,, i . , ~
vO 1oo 200 300 400 500 600 700 800
Shear Rate[I/s]

Figure 3. First cycle of a shear stress/stain rate response of 2.0% (wt/vol)


(solid symbols) waxy maize and (open symbols) normal maize or "Buffalo"
starch solutions in 0.2N KOH at 30 ~ C. The loop designated 1 is
anticlockwise; loop 2 is clockwise. This normal maize starch solution did not
exhibit shear thickening behavior; the waxy maize starch solution did. (From
Dintzis and Bagley [8]).

Rheological measurements in any event require careful analysis to


understand the processes involved in the shearing of complex fluids or
concentrated suspensions/dispersions, particularly when non-steady state data
are obtained as in the use of the thixotropic loop experiments in which shear
rate is increased and then decreased over a certain range and in a specified
70

time period. Results can be a reflection of the relaxation times of the


materials being studied. For example, Professor I. Krieger emphasized in a
recent meeting that linear viscoelastic materials will always yield
anticlockwise loops in plots of shear rate (abscissa) versus shear stress
(ordinate), and subsequently in a private communication he demonstrated this
for both the Kelvin (Voigt) and Maxwell models (Krieger [13]). Greener and
Connelly [9] showed that the application of a T-Loop ramping of shear rate
versus time can lead, for the Wagner model, to anticlockwise hysteresis loops
or even to figure 8 plots such as are shown in Figure 3.
When shear thickening effects are observed it is thus important to determine
whether the observations are simply a reflection of the relaxation time
characteristics of the material being investigated or whether the thickening is
a reflection of structure buildup during shear. When structure buildup occurs
it then becomes important to know whether this structure is stable or reverts
with time (with or without further shear treatment) to its initial state.
Shear thickening has also stimulated work on non-equilibrium molecular
dynamic (NEMD) methods as tools for investigating the origins of shear
dilatancy and shear thickening phenomena (Woodcock [14]). These NEMD
calculations show that shear thickening and volumetric changes can occur even
in simple classical dense fluids as kinetic manifestations of a transition to
athermal particle dynamics at shear rates approaching a critical characteristic
response frequency. These computations are perhaps relevant to colloidal
dispersions, corresponding to suspensions of particles, but the application of
the methodology to polymer systems is probably not feasible at this time.

3. SUSPENSIONS

Numerous examples of shear thickening going back to the original studies


by Reynolds of dilatancy of sand systems could be cited. A particularly
interesting example and one that might be considered to be related to food
systems is given in the work by Williamson and Heckert [15] who worked
with corn starch dispersions. They demonstrated the increase of apparent
viscosity with increase in shearing stress using a modified Stormer viscometer
and discussed results in terms of geometrical factors (starch granule packing)
and related concerns including the question of starch granule swelling in
water. A more recent investigation of starch systems was included in the
extensive work by Hoffman dealing with discontinuous and dilatant viscosity
behavior in concentrated suspensions. His studies (Hoffman [ 16,17]) included
a variety of suspensions including monodisperse suspensions of polymeric
materials. In later studies (Hoffman [18]) the behavior of unfractionated
71

potato starch suspended in glycerol was examined using a parallel plate


rheometer with various gap widths and volume fractions at which both shear
thickening (volume fractions of 0.52 and 0.56 respectively) and discontinuous
flow (volume fraction of 0.60) curves were obtained (Figure 4). These
volume fraction levels are in the range where one might expect shear

10 4
LL

o 10 3 -
(n ro
0 .on
o >

.o LL
(/) t-
r c) 10 2 -

03 r
or)

10 ........ i ........ I ........ i ........ I .......


10 -2 10 -1 1 10 10 2 10 3
TIFf', ( S u s p e n d i n g Fluid Viscosity) x ( S h e a r Rate); ( d y n e s / c m 2)

Figure 4. Influence of the volume fraction of unfractionated potato starch


granules on the flow behavior of potato starch granules/glycerol suspensions;
torsional flow at about 23 ~ C; O Volume fraction starch (VF) = 0.52; rn VF
- 0.56; V VF - 0.60. (From Hoffman [18] by permission).

thickening and dilatancy due to particle crowding and ordering into layers, as
discussed by Hoffman. Although not explicitly stated the viscosity data given
by Hoffman appear to be steady state values.
Starch particles in general, and certainly potato starch in particular, are not
uniform spheres and as Hoffman discusses one needs to consider that particle
size distribution and particle shape can both influence the thickening effects.
As Williamson and Heckert noted, and as was no doubt also true for the work
of Hoffman, the starch granules are little affected by solvent (water, glycerol)
at room temperature. In many actual food processing applications, though,
temperatures would be in the gelatinization range, high enough that swelling
of the starch may occur, or higher, where continued exposure to high enough
temperatures, pressures and shear would lead to starch granule disruption and
solubilization.
72

Further studies of the flow of potato starch suspensions have been reported
by Bang et al. [ 19]. They report dilatant behavior but caution that "Not all the
shear thickening is associated with a volume increase, but the term dilatancy
is more broadly used for the materials which increase viscosity with increase
of shear rate." However, the work of Bang et al. [19], while showing shear
thickening of potato starch suspensions at concentration levels of 14 to 30 wt
% at temperatures in the range of 20 ~ C-40 ~ C does not show the flow
discontinuity which Hoffman reports at concentration levels such as 50% or
greater. Bang et al. give a theoretical explanation that involves swelling of the
outer amylopectin molecules of the starch granule for the "dilatancy" in their
systems. This hypothetical swelling with the extension of the outer granule
amylopectin molecules into water is proposed as the mechanism for the
formation of a "three-dimensional scaffold structure" which results in the shear
thickening. More data on the partial swelling of the starch granules at these
low temperatures (below 40 ~ C) would be needed in establishing the validity
of the Bang et al. explanation for the thickening.
Rheological investigations of the "cooking" of starch granules have been
made (Bagley and Christianson [20]); Christianson and Bagley [21]). Figure
5 taken from the Christianson and Bagley [21] reference illustrates the type
of behavior observed. The dispersion was 25% com starch in water, heated
at 65 ~ C (on the lower end of the gelatinization temperature range for this
starch) for various times. After heating, a sample was removed from the
heating vessel and the viscosity measured in a Haake Rotovisco Couette
rheometer at 60 ~ C. The shear rate was raised from the low end (around 3
reciprocal seconds) to the high shear rate range in 3 to 5 minutes depending
on the upper shear rate reached (see Figure 5).
The initial uncooked dispersion had a very low viscosity which could not be
measured on the equipment used for the cooked samples. After 15 minutes
heating at 65 ~ C the starch granules had swollen enough that the viscosity
level was now well within the measurable range for the Rotovisco. As can be
seen from Figure 5 the viscosity initially decreases with increasing shear rate
but a region is reached just above 10 reciprocal seconds where the viscosity
begins to increase with increasing shear rate. This shear thickening is
interpreted as follows. After 15 minutes at 65 ~ C the effective volume fraction
of the water-swollen starch granules is now high enough to have a viscosity
in the range 500 to 1000 cp. However, the granules still have a relatively
rigid core and act as more or less rigid particles giving rise to the shear
thickening shown in the lower curve of Figure 5. After 30 minutes the
particles have swollen still more so that the effective volume fraction is such
that viscosities half a decade higher than the 15 minute cook are observed.
The particles, however, are more deformable, being much less rigid, more
73

100,000
At 60~ after cooking at 65~
| 15 min o 45 min
~e~ A 30 min 9 75 min
e~ 25% (db)corn starch in water

10,000 --0~0~ ~o

..=....
~
r
.....,.
Q~
"~'---A-- A.----A--A ~ A .

1,000

" E ) ~ E ) _ E).___E)._E)~ E ) ~ E)--" E)~ E ) ~ " E)..

100 1 , I
1 10 100 1000
Shear Rate (sec-1)
Figure 5. Viscosity (cp) vs shear rate (sec l) for a 25% dispersion of normal
maize starch in water, measured at 60 ~ C after cooking in a Corn Industries
Viscometer at 65 ~ C for 15 (Q); 30 (A); 45 (n); and 75 ( e ) rain. (From
Christianson and Bagley [21].)

plasticizing water having penetrated the granules. The evidence of shear


thickening is minimal. At still longer cook times, 45 and 75 minutes, the
granules occupy still greater volume fractions, are now very deformable with
little rigidity, and the high viscosity dispersions show typical non-Newtonian
shear thinning behavior. Such shear-thinning behavior for concentrated
dispersions of deformable, swollen particles has been seen earlier (Taylor and
Bagley [22]). It is particularly interesting that for highly swollen and readily
deformable gel particles this shear thinning behavior even continues at
dispersion concentrations corresponding to volume fractions of unity, that is
for dispersions of close-packed deformable particles.
The question must be considered as to whether the shear thickening is a
reflection of the fact that the structural relaxation time is comparable to the
74

time required to ramp the shear rate from the low initial level to the upper
shear rate. Unfortunately, if the ramp time is varied to resolve this question
it means that the dispersion must be kept at 60 ~ C for longer periods of time
and the experimental result will be clouded by the effects of a longer cooking
period. However, the shear thickening observed for measurements made
around room temperature where the ratio of ramp time to relaxation times is
quite different from that at 60 ~ C suggests that the shear thickening observed
is not a result of time effects. More extensive investigation of the point would
be desirable. Nevertheless, the observation of shear thickening in these studies
certainly seems to provide information on the swelling process and the
changes in the deformability of swollen starch granules during the cooking
process.

4. S H E A R T H I C K E N I N G EFFECTS IN S O L U T I O N S / D I S P E R S I O N S

It is rather remarkable that moderately concentrated and dilute


dispersions/solutions have been shown to exhibit flow behavior that has been
termed shear-thickening. For example, such systems include: the copper salt
of cetyl phenyl ether sulfonic acid, at concentrations as low as 0.02%,
(Hartley [23]); isoionic DNA (Kuznetsov et al. [24]); 5% rubber in toluene
(Crane and Schiffer [25]); 5% polymethacrylic acid (Eliassaf et al. [26]); and
dilute systems of polyacrylamide in glycerol-water mixtures (Dupuis et al.
[27]). All exhibit antithixotropic and/or dilatant behavior. In some cases,
particularly where there are strongly interacting ionic, polar or hydrogen
bonding groups, mechanisms are conceivable which might lead to the structure
formation implied by shear thickening effects. In other cases of flexible
polymer molecules the mechanism by which shear thickening could occur are
not so evident. This is in contrast to the shear thickening seen in dispersions
of rigid particles where one can see intuitively how geometric and crowding
effects could lead to shear thickening and dilatancy and detailed mechanisms
have been proposed to describe the effects quantitatively (e.g., Hoffman
[16,17]).
It has recently been shown that relatively dilute starch systems can also show
shear thickening behavior. Starch is a particularly interesting molecule to
consider in this regard. First of all, it is of practical importance having
extensive use both as a food and in industrial applications such as paper
coating. Second, it is a biopolymer which exists both in linear (amylose) and
branched (amylopectin) forms, and hence is of particular scientific interest.
Both the scientific and technological interest in starch are magnified because
the molecular weights of the starch, especially the amylopectin component,
75

can be extremely large, of the order of one hundred million or more. This
means that starch, in terms of the number of monomer units per polymer
molecule, is much larger than most normal commercial thermoplastic polymers
such as polyethylene and polystyrene.
Starch is also interesting and challenging because of its solubility
characteristics. Common practice in the use of starch has been to "paste" the
material by heating with mechanical stirring at temperatures up to about
95 ~ C (Whistler et al. [28]). While well beyond the temperature range at
which gelatinization occurs (i.e., where starch crystalline melting occurs and
significant swelling due to water sorption takes place) the starch granules are
not by any means fully solubilized. Amylose may be largely in solution but
relatively little amylopectin can be fully solubilized, and starch granule
fragments may persist even to 100 ~ C and above. The gelatinization range
depends on numerous factors such as the type of starch (corn, rice, wheat,
etc.), the variety (e.g., waxy corn starch consisting primarily of amylopectin)
and various agronomic factors. For common starches this gelatinization range
is 60o-80 ~ C.
With so many hydrogen bonding sites it seems surprising that starch is not
more readily soluble in water. In fact, it is necessary to heat starch in water
well above 100 ~ C (to, say, 120 ~ to 140 ~ C in an autoclave, for example) to
approach polymer solubilization. Heating starch, with stirring, to such
temperatures produces "solutions" that are quite different in properties from
dispersion/solutions that have only been "pasted" by heating to around the
95 ~ C level (Christianson et al. [29]). Alternate routes to solubilizing starch
include jet cooking (in which a starch slurry is passed through a nozzle with
high pressure steam, Christianson et al. [30]) and the use of solvent systems
such as dimethyl sulfoxide (DMSO), water/NaOH, water/KOH and other
alkaline aqueous systems.
Starch is heated above 100 ~ C in aseptic food processing. This can be a
continuous sterilization process for viscous liquid foods in which high
temperature is applied for a short time. Dail and Steffe [31,32], in studying
this aseptic processing, found that solutions of cross-linked waxy maize starch
showed shear thickening effects and that this was probably due to dilatancy
instead of time-dependent thickening due to on-going pasting.
In investigations of the effect of thermo-mechanical processing on starch
structure and size it was found that the intrinsic viscosity of starch decreased
significantly as processing severity increased (Dintzis and Bagley [33]). At
the same time dilatancy was evident in samples that were not severely treated
(Dintzis and Bagley [8]). Since the viscosities were obtained with a shear
sweep the observed thickening could have been a result of time effects as
discussed by Krieger [13]. However, the thickening in these starch solutions
76

was of an unusual type, as indicated in Figure 6 taken from Dintzis et al. [34].
In a shear sweep from rest to 750 reciprocal seconds over a period of 2
minutes, and then without pause decelerating to zero shear rate in another 2
minutes, with several subsequent repeats of this cycle, results as shown in
Figure 6 were obtained. The solution freshly placed in the fixture was initially
shear thinning but just beyond 100 s~ shear thickening was observed. After
an increase of nearly half a decade in viscosity, the solution became again
shear thinning. Subsequent lowering of the shear rate gave a viscosity/shear
rate plot following a power law all the way down to the lower shear rate limits
of the machine. Subsequent cycling gave viscosity/shear rate curves which to

101

o o . Fresh c o o k
60 ~C F l o w
.x. 3 hrs. old, stored in D e w a r

f/)
C 0
rl
v
x o
>, 10 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
o,.~. x i
or) x
x
0
o
O x
x
o
x
o o
~
O x x ,,, :"Ooc b
x ~,,,, "o

>

o o o
o c>! o 0.~. om~ (~

1 0 -1 . . . . . . . . , . . . . . . . . i . . . . . . . .
10 0 10 ~ 10 2 10 3

Shear Rate (s -1)


Figure 6. Viscosity/shear rate behavior in Couette flow at 60 ~ C of a 10%
(wt/vol) sample of waxy maize starch autoclaved with stirring at 140 ~ C for
15 min at pH - 7.0. Because the initial low stresses in these aqueous systems
are below the operating range of the sensor, the first three data points on each
of the curves AB are unreliable and probably not valid. O = freshly cooked
sample; X = sample stored in Dewar flask for 3 hr prior to flow measurement.
(From Dintzis et al. [34]).

a first approximation followed the initial downcurve BC, from approximately


750 to about 9 s ~ of Figure 6 though the viscosity level did drop slowly on
77

successive recyclings. The result was shown not to be an artifact of the


instrument (see Figure 2 in Dintzis et al. [34]). These results have been
extended to waxy rice and waxy barley starches and it has been established
that it is the amylopectin component that causes these effects (Dintzis et al.
[35]). With careful sample preparation one can detect a very attenuated form
of the shear thickening and structure formation effects in regular maize starch.
This very sizeable shear thickening effect, in which the viscosity increased
by almost half a decade, occurred only with freshly prepared solutions. Once
the higher viscosity level had been generated, the starch solution showed, on
subsequent shear rate cycling, shear thinning behavior only. Structure
formation during the shear thickening portion of the initial curve was
suspected and phase microscopy pictures confirmed this suspicion as shown
in Figure 7. The fresh, unsheared samples of waxy maize, dull waxy maize
and normal maize starches showed little evidence of any structure. Shearing
of the regular or normal starch solutions generated little, if any, structure.
However, shear-generated "structure", as indicated by refractive index
inhomogeneities in the micrographs, is quite evident in the dull waxy maize
sample and predominates in the case of the waxy maize starch. The later
sample appeared quite clear, water-white, but after shearing through the shear
thickening region the solution developed an obvious opalescence.
Results in these studies are very sensitive to sample preparation procedures.
This can be illustrated with reference to time effects observed in some
solutions. These effects can be seen (Figure 7 of Dintzis et al. [34]) where
shear thickening was observed occurring over surprisingly long times at
constant shear rate for a gently dispersed 3% waxy maize sample in
90%DMSO-H20. However, vigorous stirring changed this time behavior
significantly, eliminating the shear thickening effect. Normal maize solutions
generally appear to be simple shear thinning materials.
A more detailed study of 3 g waxy maize starch in 100 g of 90% DMSO
showed three different types of behavior. When the dispersion/solution in a
flask was stirred with a stirring bar for 1 hour at 100 rpm, one initially
observed an opaque dispersion of low viscosity that soon became transparent
as the viscosity increased during stirring. Finally, the fluid became viscous,
elastic, and water white, but with "fish eyes" (visible regions of refractive
index inhomogeneity) readily apparent. This fluid, when examined using a
Rotovisco R rheometer with shear rates held at 10, 50, 100 and then again 10
s1 for 20 minutes at each rate, showed no shear-thickening behavior. This
starch stock fluid was subjected to further stirring and then tested again in the
Rotovisco R. After sufficient stirring of the stock fluid, shear-thickening
behavior was observed in the rheometer. When the stock fluid was again
subjected to additional stirring, the shear-thickening effect and the "fish eyes"
78

disappeared and the solution became shear thinning.

. .i~

....

9,:<~,~,,, ,,

i:~(,"':
C ~
~,~ ~,

" 7"'

. ...
9~"" ,~,,;;

.~

E. . . . ..,.,.
~ ' . F .........

. . . . . . .~ :,~. . .

... ....

... :

~ ~7"'~' ' .~"

lOpm ..,,,,,..,,
.:

Figure 7. Phase Contrast micrographs of sheared (in cone and plate) vs not
sheared starches in 90% DMSO-H20. A, 2.0% waxy maize starch sheared;
B, 2.0% waxy maize starch not sheared; C, 2.0% dull waxy maize starch
sheared; D, 2.0% dull waxy maize starch not sheared; E, 2.6% normal maize
starch sheared; F, 2.6% normal maize starch not sheared. Starch
concentrations chosen to provide approximately equal amylopectin amounts
in each fluid. (From Dintzis et al. [34]).

5. W A T E R SOLUBLE CHITOSAN DERIVATIVES

Clearly, complex processes are at work in these shear-thickening starch


79

systems but other biopolymer systems can also show surprising behavior.
Thus, Yalpani et al. [36] reported on the complex rheological behavior of a
branched, water-soluble chitosan derivative, 1-deoxylactit-l-yl chitosan, a
comb-like polymer. The complexities reported included shear thickening seen
in viscosity versus shear rate plots for 1 and 2% (wt/wt) concentration levels
of this chitosan derivative in water. The authors emphasized that steady state
values of shear stress were obtained in their Brabender Rheotron viscometer
and used in calculation of viscosities. The onset of the observed viscosity
increase occurred at a lower shear rate for the 2% than for the 1% solution.
Lower concentrations (0.5% and 0.1%) showed approximately Newtonian
behavior over the shear rate range of 10 to 3000 s -~.
Rising temperature reduces the shear thickening exhibited by this chitosan
derivative. For a 1% solution an approximately five-fold viscosity increase
occurs in the shear thickening region at 25 ~ C. The shear rate at which
thickening occurs increases with increasing temperature over the range 25 ~
35 ~ C. Specifically, the viscosity peak occurs at 40 s -~ at 25 ~ C and at 35 ~ C
at about 350 s -1. The magnitude of the viscosity increase in the shear
thickening region decreases with increasing temperature. At 43 ~ and 50 ~ C
evidence of shear thickening has pretty well disappeared and the solutions are
only slightly shear thinning in character, or in the authors' words " .... at
43 ~ C, essentially Newtonian flow is again observed."
Yalpani et al. also report some unusual time effects for these chitosan
derivative solutions, in particular showing oscillatory "shear stress-time
dependency at steady shear." To illustrate, for the 2% solution, at a shear rate
beyond the peak of the viscosity-shear rate plot, oscillations in shear stress at
constant shear rate is seen. The corresponding viscosity variation is quite
large, from 800 to 2000 Pa.s, "without any appreciable damping being
apparent over the periods (few minutes) .... investigated." The phenomenon
was not observed with the 1% solution.

6. SHEAR THICKENING IN PROTEIN SYSTEMS

The effect of shear on structure, phase separation, mixing of polymers and


multiphase polymer systems has been noted briefly earlier and recent books
serve to spotlight the current interest in these effects (Nakatani and Dadmun
[3]; Sondergaard and Lyngaae-Jorgensen [4]). The importance of these effects
in food systems, particularly in protein-based foods, has been emphasized by
Suchkov et al. [37]. They note that: "The problem of structure-property
relationships for protein solutions under shear flow conditions is of great
applied importance for modelling food protein recovery and processing. There
80

are several reasons for this. Food systems are always subjected to shear field
(mechanical) treatments during processing. This occurs particularly during
mixing, stirring, shaking, pouring, homogenization, pumping, extruding and
cutting. Normally, the larger scale and greater efficiency of food industrial
processes are related to the more intense shear forces. The mechanism of
shear effect on protein recovery from dilute solutions is of significance for the
improvement of technologies for protein isolation from different raw materials,
for scaling up and for wider use of continuous precipitation." The same
authors also note that proteins together with polysaccharides are the main
components responsible for the properties and quality of foods. Mechanical
treatment is an important part of the processing of these components, alone
and in combination, and the rheological characteristics are critical aspects of
protein/polysaccharide-based foods. Clearly, structure formation during
shearing, as indicated by shear thickening effects, can be critical in influencing
the final physical and organoleptic properties of processed foods.

~ o o o-1
T
0-2
2O

E
00
~ 15 -i
/
7

10 I I I ... I
0 10 C* 20 30 40
Cpr, % wt.

Figure 8. Phase diagram of the legumin-NaCl-water system" 0, turbidity


measurements; O, phase component analysis. (From Suchkov et al., private
communication and [37]).
81

The work of Suchkov et al. is of significance in the food area because they
examined both the rheological and phase behavior of an important protein
system, specifically salt solutions of the 11S globulin protein from broad beans
(legumin). Stable single phase salt solutions of this protein "can undergo a
reversible transition from the single phase to the two phase state .......... either
on cooling or diluting the protein solution with water." A concentrated
"mesophase" and a dilute phase are formed so that in effect the water solution
changes into a water-in-water emulsion.
Figure 8 (Figure 3 of Suchkov et al. [37]) shows the phase diagram for 11S
broad bean globulin-NaCl-water system. Viscosity measurements of a 19.6
wt % system were made at temperatures from 11 ~ C to 29 ~ C in 1.5 ~ C
intervals, and the results for temperatures 14.0 ~ to 21.5 ~ are shown in Figure
9 (Figure 2 of Suchkov et al.). The authors note that above the binodal the
single phase legumin/salt solutions showed Newtonian behavior while within
a range of temperatures below the binodal the system exhibited non-
Newtonian behavior and specifically showed shear thickening effects. It is
evident from Figures 8 and 9 that the onset of apparent shear thickening does
not coincide very exactly with the development of two phases. It is not until
17.0 ~ C, well below the maximum of the phase diagram plot, that the shear
thickening is seen. The author's comment (V. Ya. Grinberg, personal

0,2 --

6 (14~
0,1
oo -5
n
4 (17~
E 3
2 (20~
~" 0,02 1

I ...... I ! I
0,01
0 600 1200 1800 2400
~/, S-1

Figure 9. Flow curves at different temperatures of the legumin-NaCl-water


system for a protein concentration of 19.6 wt%. (From Suchkov et al., private
communication and [37]).
82

communication) that the shear sweep cycle time was about 40 rain. and that
each of the coexisting phases in the protein system was a Newtonian liquid
suggests that the observed shear thickening was not the result of time effects.
It is also interesting to observe in Figure 9 that viscosities can be lower, at
the lower shear rates, for the low temperature measurements than at the
higher temperatures; for example at 14.0 ~ C the viscosity below about 750
sl is lower than the Newtonian viscosities at 17 to 21.5 ~ C. Such a reversal
could be indicative of a shear dissolving process. Certainly, the results are
extremely interesting and potentially valuable for process design applications.
Considerably more work and extensive investigation of protein and food
systems is needed along lines discussed in the literature, for example, by Wolf
and Horst [38] and Tolstoguzov [39].

7. H O N E Y

Scott Blair [40] in Chapter VI, The Rheology of Honey, from his well-
known book on "Foodstuffs, Their Plasticity, Fluidity and Consistency" notes
that the term "dilatancy" had been expanded "of recent years" to include
"any system whose viscosity increases with increasing rate of shear beyond a
certain minimum value". He noted that "the system need not contain any solid
particles but can be a true solution such as gum arabic in water to which a
small amount of borax has been added." He then went on to reference the
work of Pryce-Jones [41 ], who, using a twin-Couette rheometer, demonstrated
that a large number of honeys showed the property of shear thickening.
Specific honeys, such as honey from Eucalyptus ficifolia, also possessed the
property of being easily drawn into long strings, the property of
"Spinnbarkeit". Scott Blair further emphasized that dilatant honey systems
displayed the Weissenberg effect, as evidenced by the fact that when a glass
rod is immersed in honey and rotated at certain speeds the honey will climb
up the rod. This certainly implies that the system is exhibiting considerable
elasticity so that one evidently is dealing with viscoelastic fluids for these
dilatant honeys. Honey dilatancy is also emphasized in more recent texts, for
example in Sone [42].
Honeys are essentially concentrated sugar solutions. However, for the 15
dilatant honeys investigated by Pryce-Jones [43] all contained dextran
(C6H1005) n with n of the order of 8000. This would mean the dextran is a
relatively high molecular weight material, approximately 1.3X106 Daltons.
The concentrations of dextran are relatively low, quoted by Scott Blair as
7.2% dextran for the Eucalyptus ficifolia and 6.4% for Eucalyptus eugeniodes
honeys, respectively. Of course, for molecules of molecular weights over a
83

million one should perhaps regard 6 and 7% solutions as really quite


concentrated. The viscosity increase for the 7.2% solution occurs over a
narrow shear rate range. At shear rates of 50 reciprocal seconds the viscosity
is about 85 poise, rising to 90 poise at 75 reciprocal seconds and 130 poise at
shear rates around 110 s~.
That the dilatancy is due to the presence of the dextran is shown by the
observation (Scott Blair [40]) that the removal of the dextran by precipitation
with acetone and reconstituting the honey to its original concentration yields
a product which behaves like a true Newtonian fluid. Scott Blair also notes
that the addition of dextran to clover or sainfoin honeys, which do not show
dilatancy, stimulates "The peculiar rheological effect" of (presumably) shear
thickening and rod climbing. Although, as Sone notes, such "observations
indicate that dextran is the cause of dilatancy in honey" the exact mechanism
is obscure. No doubt, as in the case of starch fluids discussed above, the large
number of hydrogen bonding sites accessible in the dextran, as in starch,
contributes to the formation of some structure under shear which leads to a
shear-thickening effect. There has been very compelling rheological evidence
confirming the part played by dextran in these shear thickening effects
(Sabati~ and Choplin [44]; Choplin and Sabati~ [45]; Sabati6 et al. [46]).
Sabati6 and Choplin used a carefully prepared dextran and found that structure
formation did occur and that the viscosities of the materials before and after
structure formation differed by at least a decade, the structure formation being
irreversible [44]. The shear thickening and structure formation behavior of
dextran systems are indeed very similar to the behavior of amylopectin
solutions showing features such as anticlockwise "thixotropic loops"
(Figure 6, [45]). Choplin and Sabati~ also conclude that the structure formed
in dextran, irreversible provided sufficient strain is applied, may be due (in
part) to the ability of the molecule to interact by hydrogen bonding. However,
this would not address the problem of the reported differences among
amylopectins in the magnitude of the effects. Some interesting earlier
references are provided by Sabati~ et al. [46] in reviewing the rheology of
native dextran solutions and subsequently in this paper the authors conclude
that, in addition to hydrogen bonding effects, "chain-chain" packing that
depends strongly on molecular conformation contributes to the dextran shear
thickening effects. Equivalent conformational differences no doubt lie at the
root of the reported differences among amylopectins [33,34,35] but it is not
at all evident how these conformational differences could be more clearly
illuminated.
84

8. RECENT MODELS

Dr. Craig Carriere kindly called our attention to some recent papers
describing theoretical models for the development of shear thickening. There
is the phenomenological model of Hess and Hess [47]. Wang examined a
model treating the dynamics of coexistence between a physically cross-linked
network with temporary junctions and unbound chains [48]. Hatzekiriakos et
al. [49] modeled polymer macromolecules as Hookean elastic dumbbells
deforming affinely during flow. While not exhaustive, these references are of
fundamental interest and computational investigation of the applicability of
these approaches to specific shear thickening systems is clearly warranted. It
must be noted, though, that these models imply structural reversibility which
is not observed in the behavior of amylopectin.

9. A C K N O W L E D G E M E N T

We very much appreciated seeing a pre-publication copy of the manuscript,


"Effect of shear on the phase behavior of the legumin-salt-water system.
Modelling protein recovery.," by V.V Suchkov, I.A. Popello, Ya. Valerji and
V.B. Tolstoguzov. We thank these authors for their kind permission to
include Figures 8 and 9 in this review and for their promptness in responding
to our questions. Thanks are also due to Dr. Craig Carriere for helpful
discussions.

REFERENCES

~ W.W. Graessley, Viscoelasticity and Flow in Polymer Melts and


Concentrated Solutions, in "Physical Properties of Polymers," Am.
Chem. Soc., Washington, D.C. (1984).
. E.B. Bagley and H.P. Schreiber, Elasticity Effects in Polymer Extrusion,
in "Rheology, Theory and Applications, Vol. V," F.R. Eirich (ed.),
Academic Press, N.Y. (1969).
. A.I. Nakatani and M.D. Dadmun, Flow-Induced Structure in Polymers,
ACS Symp. Series 597, Am. Chem. Soc., Washington, D.C. (1995).
, K. Sondergaard and J. Lyngaae-Jorgensen, Rheo-Physics of Multiphase
Polymer Systems, Technomic Pub. Co. Inc., Lancaster, Pennsylvania
and Basel, Switzerland, 1995.
~ B. Launay, and B. McKenna, Physical Properties of Foods, R. Jowitt,
F. Escher, B. Hallstrom, H.F.Th. Meffert, W.E.L. Spiess and G. Vos,
(eds.), Appl. Sci. Pub., Ltd., UK (1983).
85

6. R.I. Tanner, Engineering Rheology, Clarendon Press, Oxford, 1985.


7. J.M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold
Co., New York, 1982.
8. F.R. Dintzis and E.B. Bagley, J. Appl. Polym. Sci. 56 (1995) 637.
9. J. Greener and R.W. Connelly, J. Rheol. 30(2) (1986) 285.
10. M.K. Chow and C.F. Zukoski, J. Rheol. 39(1) (1995) 15.
11. D.C-H. Cheng, and F. Evans, Brit. J. Appl. Physiol. 16 (1965) 1599.
12. H.A. Barnes, J. Rheol. 33(2) (1989) 329.
13. I.M. Krieger, Comments made at the Rheol. Symp., Pittcon '96 Conf.,
Chicago and in private communication (1996).
14. L.V. Woodcock, Chem. Physics Letters, 111 (No. 4,5) (1984) 455.
15. R.V. Williamson and W.W. Heckert, Ind. Eng. Chem. 23(6) (1931) 667.
16. R.L. Hoffman, Trans. Soc. Rheol. 16(1) (1972) 155.
17. R.L. Hoffman, J. Colloid and Interface Sci. 46(3) (1974) 491.
18. R.L. Hoffman, Advances in Colloid and Interface Sci. 17 (1982) 161.
19. J-H. Bang, E-R. Kim, S-J. Hahn and T. Ree, 1983, "Flow Mechanism
Of Dilatant Systems. (1) Starch Suspensions in Water," Bulletin of
Korean Chemical Sot., Vol. 4, No. 5212-217, 1983.
20. E.B. Bagley and D.D. Christianson, J. Texture Stud. 13 (1982) 115.
21. D.D. Christianson and E.B. Bagley, Cereal Chem., 60(2)(1983) 116.
22. N.W. Taylor and E.B. Bagley, J. Polym. Sci., Polym. Phys. Edition 13
(1975) 1133.
23. G.S. Hartley, Nature 142, (1938) 161.
24. I.A. Kuznetsov, V.F. Popenko and S.M. Filippov, Biofizika 23 (1978)
539.
25. J. Crane and D. Schiffer, J. Polym. Sci., 23 (1957) 93.
26. J. Eliassaf, A. Silberberg and A. Katchalsky, Nature 176, (1955) 1119.
27. D. Dupuis, F.Y. Lewandowski, P. Steiert and C. Wolff, J. Non-
Newtonian Fluid Mechanics 54 (1994) 11.
28. R.L. Whistler, J.N. BeMiller and E.F. Paschall, Starch: Chemistry and
Technology, 2nd ed., Academic Press, Orlando, Florida, 1984.
29. D.D. Christianson, E.M. Casiraghi and E.B. Bagley, Carbohydr. Polym.
6 (1986) 335.
30. D.D. Christianson, G.F. Fanta and E.B. Bagley, Carbohydr. Polym. 17
(1992) 221.
31. R.V. Dail and J.F. Steffe, J. Food Sci., 55(6) (1990) 1660.
32. R.V. Dail and J.F. Steffe, J. Food Sci., 55(6) (1990) 1764.
33. F.R. Dintzis and E.B. Bagley, J. Rheol. 39(6) (1995) 1483.
34. F.R. Dintzis, E.B. Bagley and F.C. Felker, J. Rheol. 39(6) (1995) 1399.
35. F.R. Dintzis, M.A. Berhow, E.B. Bagley, Y.V. Wu and F.C. Felker,
Cereal Chem. 73 (1996) 638.
86

36. M. Yalpani, L.D. Hall, M.A. Tung and D.E. Brooks, Nature, 302 (1983)
812.
37. V.V. Suchkov, I.A. Popello, V.Y. Grinberg and V.B. Tolstoguzov, Food
Hydrocolloids 11(2) (1997) 135.
38. B.A. Wolf and R. Horst, "Flow Induced Dissolution/Demixing in
Polymer Systems - A Predictive Scheme" in Rheo-Physics of
Multiphase Polymer Systems, K. Sondergaard and J. Lyngaae-Jorgensen
(eds.), Technomic Pub. Co. Inc., Lancaster, Pennsylvania, 1995.
39. V. Tolstoguzov, Gums and Stabilizers for the Food Industry, G.O.
Phillips, P.A. Williams and D.J. Wedlock (eds.), "Applications of phase
separated biopolymer systems," IRL Press, Oxford 1996.
40. J.W. Scott Blair, Foodstuffs, Their Plasticity, Fluidity and Consistency,
Interscience Pub., Inc., New York (1953).
41. J. Pryce-Jones, J. Sci. Inst., 18 (1941) 39.
42. T. Sone, in "Consistency of Foodstuffs," D. Reidel Co,, Dordrecht,
Holland, 1972.
43. J. Pryce-Jones, Bee World 33(9) (1952) 147.
44. J. Sabati6 and L. Choplin, "Shear Induced Structure in Dextran
Solutions" in Proceedings of the Symposium on Recent Developments
in Structured Continua, D. DeKee and P.N Kaloni (eds.), Longman
Scientific and Technical, (1985) 165.
45. L. Choplin and J. Sabati6, Rheol. Acta 25 (1986) 570.
46. J. Sabati6, L. Choplin, F. Paul and P. Monsan, Rheol. Acta 25 (1986)
287.
47. Hess, Ortwin and Hess, Siegfried, Physics A., 207 (1994) 517.
48. Wang, Shi-Qing, Macromolecules 25 (1992) 7003.
49. Hatzekiriakos, G. Savvas and D. Vlassopoulos, Rheol. Acta 35 (1996)
274.
87

R H E O L O G Y OF FOOD EMULSIONS

C. Gallegos and J.M. Franco

Departamento de Ingenieria Quimica, Universidad de Huelva, Escuela


Polit~cnica Superior, La R6bida, 21819, Palos de la Ftra. (Huelva), Spain.

1. INTRODUCTION

An emulsion may be defined as a thermodynamically unstable heterogeneous


system formed by at least two liquids that are at best only slightly soluble. The
internal phase is dispersed in the other in the form of small droplets, with
diameters higher than 0.1 ~tm [1-3]. Food emulsions are governed by the same
principles as are other emulsion systems. The only specific requirements for these
emulsions are that they must posses long-term stability (several months or years)
and that they must contain only ingredients which are acceptable for human
consumption [4].
Most of food emulsions are oil-in-water (o/w) emulsions, e.g. mayonnaise,
salad dressing, coffee cream, cream liqueurs, soft drinks, ice creams,... [5, 6].
They are typically stabilized by an adsorbed layer of protein at the oil-water
interface, although many food colloids contain a mixture of macromolecular and
low-molecular-weight emulsifiers [7-10]. Different types of surface active
molecules may interact with each other in different ways [11], for example
inducing protein displacement from the interface [12].
On the other hand, water-in-oil (w/o) emulsions, such as butter, margarine or
spreads, are more dependent on the cristallinity of the oil phase, the presence of
rigid surfactants on the interface and the presence of agents which increase the
viscosity of the aqueous droplets [4].
In some cases, it is possible to obtain a multiple emulsion, for instance water-
in-oil-in-water (w/o/w). wl/o/w2 emulsions are systems where a w~/o simple
emulsion is dispersed in a second aqueous phase w2. The oil phase exists as a thin
layer, stabilized by a molecular monolayer of emulsifying agents at both oil-water
interfaces [13, 14]. In many occasions, this type of emulsions are presented as
88

reduced-calorie products since this particular microstructure yields a drastic


reduction of fat and energy content [15].
This chapter summarises the state-of-the-art of the bulk rheology of oil-in-water
food emulsions, as well as its relationship with the emulsion microstructure and
its resulting stability.

2. EMULSION STABILITY

2.1. Mechanism of emulsion destabUization.


As has been previously mentioned, stability is the most important factor to be
considered in emulsion technology. An emulsion is stable when there is no
change in the size distribution or the spatial arrangement of droplets over the
experimental time-scale. The three main mechanisms of destabilization are the
flocculation, creaming and especially coalescence of oil droplets.
Creaming results from the action of gravity on oil droplets, of lower density
than the continuous medium, yielding a vertical concentration gradient of droplets
but no change in the overall size distribution. The rupture of the emulsion may be
detected by the appearance of a separated aqueous phase at the bottom. The
aggregation of the droplets without breakdown of the emulsifier layer is called
flocculation. Flocculation may frequently be an intermediate step to the rupture of
the emulsion by coalescence, which implies the irreversible process of the
droplets becoming larger yielding a flee oil phase at the top of the sample.
The dominant mechanism of instability will be determined by the interparticle
and hydrodynamic interactions. Whether or not aggregation or coalescence
occurs, depends on the composition and surface viscoelasticity of the adsorbed
surface layer [16-23] and may be predicted by the classical DLVO theory [24,
25] or, more usually, combined with the steric stabilization theory [26].
All these mechanism are interrelated among them. Thus, flocculation generally
favours creaming since the aggregates of oil droplets have a settling velocity
higher than that of the individual droplets or, on the contrary, even when
creaming takes place without flocculation the final state is a flocculated emulsion
which may favour coalescence.

2.2. Factors affecting emulsion stability.


In general, emulsion stability depends on droplet size distribution, rheology of
the continuous phase and interparticle interactions [27, 28], and all of them affect
the bulk theology of the emulsion. As Figure 1 schematically shows, structural
parameters, bulk rheology and emulsion stability are interrelated among them. In
addition to this, all of them are decisively influenced by processing conditions.
89

EMULSION
STABILITY

STRUCTURAL PARAMETERS

* rheology of the continuous phase

RHEOLOGY ]~. * n a t u r e of particles 9

r droplet size distribution


-

- deformability
- internal viscosity
- surface viscoelasticity
- concentration

* interparticle interactions

PROCESSING CONDITIONS

* temperature
* residence time
* agitation speed
* etc.

Figure 1. Relationship among microstructure, processing, rheology and stability


of food emulsions.
90

Consequently, emulsion rheology can not be understood without considering the


structural parameters of the emulsion. At the same time, a rheological
Characterization of these materials provides useful information in order to predict
the stability of emulsions.
One of the most important factors affecting emulsion stability is the droplet size
distribution. It is a well-known fact that fine emulsions with narrow distributions
yield a higher stability than coarse emulsions with broad distributions [28, 29].
Lower droplet size favours stability against creaming since settling velocity is a
function of droplet diameter.
Another crucial factor influencing emulsion stability is the strength and nature
of interparticle interactions. If the attractive forces exceed the repulsive
electrostatic and/or steric interactions the aggregation of oil droplets occurs and
coalescence may take place depending on the surface viscoelasticity of the
surface layer. Nevertheless, if the primary minimum of the energy between the
approaching droplets is not deep the flocculation is easily reversible and
temporary [4, 27]. As has been previously mentioned, flocculation may frequently
be an intermediate step to the rupture of the emulsion, either after a creaming
process or after the coalescence of oil droplets. However, in concentrated
emulsions an extensive flocculation process may favour emulsion stability by
forming a weak gel-like particulate network [8, 30] where the continuous phase is
immobilized within the interstices [31, 32]. Classical DLVO theory indicates that
an increase in ionic strength should lead to instability. However, in emulsions
containing polymers such as proteins, the presence of electrolytes has been
shown to modify the charge stabilization among oil droplets [32, 33]. Thus, salt
affects the tridimensional conformation of ionic polymers and, subsequently
emulsion stability [34]. In the same direction, stabilization may be achieved by
modifying pH [35].
The strength and nature of the interactions among droplets depend on the type
and concentration of the emulsifiers used, which determine if the flocculation is
reversible or irreversible. The extension of this aggregation process influences the
bulk theology of emulsions [36, 37].
Further stabilization may arise from the presence of low concentrations of
water-soluble polysaccharides, hydrocolloids, which act as thickening or
structuring agents in the continuous phase [38-40]. The main function of these
hydrocolloids is to reduce the mobility of oil droplets, decreasing the extension of
both flocculation and creaming processes. Traditionally, this type of stabilization
has been related to the existence of a finite yield stress in the thickening solution
in which emulsion droplets do not cream since the gravitational lift on the
droplets is less than the yield stress. Recently, it has been suggested that the
91

stabilization takes place by the formation of a particulate network being the yield
stress of this structural network the decisive factor to stabilize the emulsion [41 ].
In relation to this, another important aspect to be taken into account is the
thermodynamic compatibility between the hydrocolloid and the emulsifier used
[5, 38, 42]. Thus, added biopolymer may have a positive or negative effect on
stability depending on its concentration and the nature of the droplet-biopolymer
interactions.

0 M I C R O S T R U C T U R E AND R H E O L O G Y OF FOOD EMULSIONS

3.1. Steady-state flow behaviour


The flow properties of an emulsion are among its more important physical
characteristics. From a technical point of view, the unit operations related to the
manufacture of an emulsion (mixing, pumping, filling, etc.) require the knowledge
of its flow behaviour to assess mixing efficiency, power consumption, etc. [43-
45]. In addition to this, knowledge of the viscous behaviour of emulsions is basic
in relation to their final use, that is to understand their response to pouring or
extrusion from packs, draining, etc. [45].
Much work has been carried out on the effect of several structural parameters,
such as disperse phase volume fraction and particle size distribution, on the flow
behaviour of o/w emulsions. Thus, many efforts have been dedicated to
understand the relationship between the viscosity and the disperse phase volume
fraction, ~ [37, 43, 45, 46].
Very dilute emulsions (~) < 0.01) show a Newtonian behaviour, being the
relationship between the relative viscosity and 4) defined by the equation
proposed by Einstein [47]:

n 1 + 2,5 (1)
fir - r l c -
where r/~ is the viscosity of the continuous phase.
The classical Taylor's treatment [48] takes into account the internal circulation
of the dispersed fluid:

(1 + 0.4(nc / n j ) )
qrel -1+2.5~ l+(qc/qd) JOO (2)

where rk is the continuous phase viscosity and r/a is the dispersed phase
viscosity.
92

As the volume fraction increases, the relative viscosity becomes a more


complex function of ~):

rl r - 1 + 2,5~) + b ~)2 + cff 3 +... (3)

where b is a coefficient that accounts for hydrodynamic interactions between the


droplets [49]. This term is usually sufficient to describe the viscosity of
dispersions up to ~ = 0.2.
Other equations widely quoted in the literature for Newtonian emulsions are
Mooney's equation [50]:

( 25~ "~
fir = exP/1 +am,J (4)

Eilers' equation [51, 52]:

( 1.25~) )
fir-~1 + 1_ aEd~) (5)

or Roscoe's equation [53]:

- (6)

In addition to these, equations proposed for solid suspensions have been also
used to describe the influence of ~ on the relative viscosity, with the following
form [54-57]:

qr = qr ($ / d~m ) (7)

where ~, is the maximum packing fraction. In suspensions, the viscosity


becomes infinite for ~ = ~m, but in emulsions ~m is a not well-defined parameter
due to the deformability of the droplets. As examples of this type of equation
could be cited:
F (~/~.,),,,3 l
Frankel and Acrivos rl~ = c L 1 - ($ / $,,),,3 J, c- 9/8 (8)
93

-[nl*m
Krieger-Dougherty
F ,
qr = [1 - ~ m
(9)

[rl] --- intrinsic viscosity

1- * 7-2
Maron-Pierce-Kitano f i r - ~mJ (10)

However, concentrated emulsions usually show non-Newtonian behaviour. This


behaviour has been related to either droplet flocculation or the non-Newtonian
behaviour of the dispersed phase [58]. Concerning the evolution of viscosity with
shear rate, the general picture shows three different regions, a constant viscosity,
rio, at low shear rates, a power-law decrease in viscosity, and finally a constant
viscosity, rl~, at high shear rates, characteristic of the unflocculated system.
Wilkinson or Carreau models have been widely used to describe this behaviour
[9, 58], although depending on the oil volume fraction all these regions may not
be observed [36, 59, 60].
A difference of several decades between 11o and rl~ is usually found in a flow
curve of concentrated food emulsions obtained using a controlled-stress
rheometer [61]. These flow curves are the result of a dramatic structural
breakdown which may be attributed to both an irreversible process (droplet
coalescence) and a reversible one (deflocculation, droplet alignment, droplet
deformation).
Consequently, the relationship between viscosity and volume fraction should
include the influence of shear rate. Some authors [46] have used the above
mentioned equations with non-Newtonian emulsions submitted at very high shear
rates, when a high-shear-rate-limiting viscosity has been attained, due to a
complete deflocculation of disperse phase droplets. In these conditions, the
viscosity of concentrated emulsions can be satisfactorily described by the
relation:

11oo _ exp 2 , 5 , ) (1 1)
qc 1- k~)

where k depends on the hydrodynamic interactions among droplets.


94

Pal and Rhodes [62] have proposed an empirical equation to correlate results
obtained with both Newtonian and non-Newtonian emulsions:

fir - fir [~ / (~)qr=lO0] (12)

where (~b),~,.=;00is the disperse phase volume fraction at which fir- 100.
The same authors [36] have developed a theoretical equation for non-
Newtonian emulsions. They explain the non-Newtonian behaviour as a
consequence of the formation of flocs, yielding the immobilization of an
important amount of continuous phase, and, consequently, increasing the effective
concentration of disperse phase. In addition to this, the effective concentration
increases due to the hydration effect, that causes an association between
continuous phase and emulsifier molecules. The equation results:

f i r -- 1 - k o k F (3)),1-2'5' (13)

where ko is the hydration factor, and ky is the flocculation factor, that depends on
shear rate (being unity at high shear rates).
Another method to describe the viscosity dependence with both shear rate and
oil volume fraction has been proposed by Partal et al. [60]. This method is based
on the superposition of the flow curves of emulsions with different oil volume
fraction, using an empirical shift factor. The relationship between the shift factor,
a(~), and the oil volume fraction is described by a Frankel-Acrivos type equation:

(14)

where 4)o is the reference oil volume fraction. The master curve shown in Figure 2
also contains the flow curves of emulsions having different emulsifier
concentration. In this case the superposition was carried out by using another
shift factor, which varies exponentially with the emulsifier concentration.
When considering the influence of ~ on emulsion viscosity it is also necessary
to know that there is a simultaneous influence of droplet size. In fact, rlrel
increases as the mean droplet size decreases. Thus, for example, k in equation
95

(11) increases as droplet size decreases [63]. In addition to this, an emulsion with
a broad distribution of droplet sizes will have a lower viscosity than those with a
narrower distribution of particle size [29]. In this case, a model that relates the
relative viscosity and ~) for emulsions containing an i-model size distribution has
been proposed [46].

~rel(~; ) = .L_Lexp
i=l 1 - ki~ i (15)

being +i the oil volume fraction for each diameter.

100

~ Carreau modd I

10 9

il-

0.1 ........ ' ........ ~ ........ ~ ........ ~ ........ ~ ........


0.01 0.1 1 10 100 1000 10000

a+asE+ (s-1)
Figure 2. Master flow curve for emulsions containing 1-10% wt. sucrose ester
and 55-85% wt. oil. Reproduced, with permission, from ref. 60.

This fact has been confirmed by different authors [9, 28, 32]. Thus, Figure 3
shows that the viscosity of salad dressing emulsions containing a mixture of
emulsifiers increases as mean droplet size and polydispersity decreases. This has
been attributed to the fact that narrower droplet size distribution and lower mean
diameter yield stronger interparticle interactions [28]. On the other hand, Partal et
al. [60] have found a linear correlation between some rheological parameters
(high-shear-rate-limiting viscosity and flow index) and polydispersity. However,
in other cases this correlation fails, because, for instance, the emulsifier that is in
96

excess may form a gel-like continuous phase. This structuration may be favoured
by increasing the droplet size, and consequently the emulsifier concentration in
the continuous phase [64].

1
ra~

,,..a

9
~ ().1

[ ~Carreaulnodel ]

0.01 . . . . . . . . I . . . . . . . . ~ .......... i . . . . . . .
0.1 1 10 100 1000

shear rate (s "l)

30 ' I ' I ' I ' I ' I '

25

20

15

0 5 10 15 20 25 3O
diameter (~.m)

Figure 3. Steady-state viscosity (a) and droplet size distribution (b) of salad
dressing emulsions (emulsification time: r'l 3 min, O 5 min, A 10 min). Adapted,
with permission, from ref. 8.
97
A large majority of food emulsions may show slip effects under steady-state
shear [61, 65-68]. It has been demonstrated that multiphase systems show slip
effects because of a displacement of the disperse phase away form the solid walls
of the sensor system in a rheometer or the walls of pipes or tubes [69]. In the case
of emulsions, this effect is even more dramatic due to the deformability of the
droplets and is also enhanced when creaming process becomes important. Figure
4 shows a typical example of a flow curve showing slip phenomena. As Barnes
[69] has pointed out we normally expect a single drop from a lower-shear rate
Newtonian plateau through a shear thinning region towards the high-shear-rate-
Newtonian plateau. However, if slip occurs, an intermediate false pseudo-
Newtonian region is observed as Figure 4 indicates.

9 lip flowcuwe

flowcurve ~ith slip effects "',9~ ~

log (stress)

Figure 4. Squematic flow curves obtained in a plate-plate geomet~, with smooth


(slip curve) and roughened (non-slip curve) plates.

As has been reported [61], the extension of slip is clearly dependent on the
emulsion composition due to the existence of different interparticle interactions
that influence emulsion microstructure. Thus, an emulsion containing a very
soluble low-molecular-weight emulsifier shows wall slip (determined by
comparison of the flow curves obtained with sensor systems having smooth and
rough surfaces respectively), in contrast to an emulsion containing an only slightly
soluble emulsifier which practically does not show slip effects. This difference
98

must be related to the aqueous phase behaviour. Thus, in the first case the
droplets are suspended in a low viscosity continuous phase favouring their
mobility, but in the other emulsion a very structured continuous phase (gel-like)
takes place, apparently reducing the slip problem.
However, the differences found in the steady-state flow curves carried out with
different types of surfaces are much higher when the viscosity is plotted versus
shear stress than those found when viscosity is plotted versus shear rate. In fact,
in many cases the differences found in the viscosity-shear rate plots are not
significant [61 ] and slip effects may be not taking into account if the experimental
flow curves are obtained using a controlled-shear rate rheometer. This can be
observed in Table I where the Carreau model parameters are compared for an
emulsion measured with rough and smooth surface sensor systems respectively.

Table I. Parameters of the Carreau model for a food emulsion.

TIo (Pa s) ~ (s -1) s


0.46
P P 2 O ( - r o u g h s u f f ac-e) .......................i [ 4 i 6 "g............................1,7 ....i0 "'-'~..............................................................
C35/4 (smooth surface) 1.5 10s 0.8 10.4 0.46
Confidence limit +0.6 105 +1.0 10 .4 +0.02

3.2. L i n e a r viscoelasticity
The linear viscoelastic properties of food emulsions are usually confined to that
range of strains below any perceivable visual movement during measurement, and
though not generally applicable to large strain and stress technical applications,
are nevertheless useful in assesing the microstructure and even possible the long-
term stability of an emulsion [45].
The first studies on the response of concentrated emulsions in the linear region
were done by creep-recovery experiments [46]. The creep behaviour of a
concentrated emulsion, as a function of disperse phase fraction, often looks as
despicted in Figure 5. Consequently, food emulsions behave as viscoelastic
liquids, even highly concentrated emulsions such as mayonnaise [28]. This
response has been modelled using different analogycal models, being the simplest
combination one Maxwell element combined with one Voigt element in series,
that is "one retardation process" model:

t e(_t/~)
.! ( t ) - do + + J1 (16)
qo
99

This model has been used by Kiosseoglou and Sherman [18] to describe the
response of an egg-yolk film at a groundnut oil-water interface and by Matsumoto
and Sherman [70] for nonionic surfactant films at an olive oil-water interface.

2.0x102

1.5x10"2

,f--,
~ l.OxlO_2 m

5.0x10-3

0.0 I , I ~ I , I ,

50 100 150 200 250


time(s)

Figure 5. Creep-recovery behavior of oil-in-water emulsions stabilized with a


nonionic surfactant.

However most of food emulsions show a more complex rheological behaviour.


Thus, Gladwell et al., [71] used a "two retardation process" Kelvin-Voigt model
to fit the creep compliance response of soya oil-water emulsions emulsified with
egg yolk and stabilized with xanthan gum.

(-t/z 1) (-t/z 2)
J (t)= Jo + J1 e + J2 e + t /rio (17)

This model has been used to describe the creep behaviour of a wide variety of
emulsions [72-76].
On the contrary, oscillatory shear tests have not been extensively used for
emulsion rheology [28]. As a consequence, only very little data were available in
the literature until a few years ago [77], when several authors started to use this
technique on different types of emulsions [37, 45, 78]. A general overview of
these studies demonstrates that the frequency dependence of the storage and loss
100

moduli is dramatically dependent on the concentration of the emulsion,


processing conditions and nature of the emulsifier used.

103 ....... '1 ' ' ...... I ........ I ' ....... I .......

0121O

OO D O0 DDOODO0 000oO Oo D OooO0


102 0
AAA &A

AAA AA A AAAAAAAAA A A A A A A A A9 l o l IlI I I I IO O


OO OO II II
O O 00 IIII
101 II I 9 t l i i i t l I I I I I
I I 9 iiii I I I I II I I
I I I I

lO 0 ........ I , . , ..... I , . , , ,,,.I , ~ ...... I i i l l i ii

10-3 10-2 10q 10~ 10 ] 02

~ (rad/s)

Figure 6. Storage and loss moduli vs. frequency for model mayonnaise with
different oil contents. I"1 75% wt., O 77.5% wt., A 80% wt., ( G ' open symbols,
G " solid symbols). Adapted, with permission, from ref. 78.

Commercial food emulsions, such as mayonnaise, give the kind of behaviour


depicted in Figure 6, with the storage modulus higher than the loss modulus in a
frequency range comprised between 10-2 and 102 rad/s, and characterized by the
appearance of a minimum in G" at intermediate frequencies and a "plateau"
region in G' (slope values of around 0.1). This behaviour corresponds to highly
flocculated emulsions, resulting from the formation of physical entanglements
among protein molecules adsorbed at the oil/water interface of the oil droplets,
which leads to the formation of a structural network.
On the contrary, Tadros [37] has studied the dynamic viscoelastic response of
non-flocculated isoparaffinic oil-in-water emulsions in a relatively wide range of
disperse phase volume fraction (~ comprised between 0.48 and 0.60), stabilized
using a block copolymer (Synperonic PE). Even for the more concentrated
systems, the loss modulus shows higher values than the storage modulus up to
frequencies of around 10 rad/s. The terminal relaxation time increases
exponentially with the disperse phase volume fraction when the latter exceeds
101

0.54. The same tendency is observed in the evolution of G" with ~. These results
are explained on the basis that at ~ < 0.56, the droplet-droplet separation is
probably larger than twice the adsorbed layer thickness and hence the adsorbed
layers are not forced to overlap or compress. In this case, the repulsive interaction
between the adsorbed layer thickness is relatively weak and the emulsion shows
predominantly viscous response. However, when ~) > 0.56, the droplet-droplet
separation may become smaller than twice the adsorbed layer thickness and the
chains are forced to compress. This leads to strong steric repulsion and the
emulsion shows predominantly elastic response.
Similar behaviour has been found for vegetable oil-in-water emulsions (~)
comprised between 0.7 and 0.8) stabilized by a nonionic-low molecular-weight
elnu|sifier highly soluble in the aqueous phase, at very low concentration (1%
wt.) [791.

lO 1 ' ' ' ' ' ' " I ' ' ' ' ' ' " I ' ' ' ' ' ' " I ' ' . . . . . 'I ' . . . . . . . I ' ' ' ' ' ' "

10~
~Z

~z
10-1
GN~ (P a)
45 % 0 3590
50 % 0 4345
55 % 0 5300

10-e , , i .i,,,| . ! I ,.,.,l i i i ,,,,,i , i i iii.,| i , , ,,,,,| , , . iii,,

10-3 10-2 10q 10~ 101 102 103

c0 (rad/s)

Figure 7. Influence of oil concentration on the normalized viscoelasticity


functions for free-starch salad dressing emulsions, l"! 45% wt., O 50% wt., A
55% wt. Reproduced, with permission, from ref. 10.

However, when a high HLB sucrose ester (slightly soluble in water at low
temperature) is used as the only emulsifier [80], or in addition to a
macromolecular emulsifier [10], in concentrated food emulsions three
characteristic regions were observed: (i) a pseudoterminal region at low
102

frequencies that shows a tendency to a crossover of both viscoelastic functions,


(ii) an intermediate "plateau" region, and (iii) the beginning of the transition
region at high frequencies. Figure 7 shows the evolution of the storage and loss
moduli with frequency for an emulsion stabilized by a mixture of the above
mentioned emulsifiers. Franco et al. [8, 10] have used an empirical model,
adapted from the one given by De Rosa and Winter [81] for polymer melts, to
successfully describe the three experimental regions that appear in the linear
relaxation spectra of these emulsions:

-
EI/m I+)nl + for Xmin < ~ < XP (18)

H (~)- A for Xp < ~ < Xmax (19)

where ~ and ~ are the characteristic relaxation times for the onset of the plateau
and pseudo-terminal regions, respectively, Xm~ and )~m~ are the reciprocal of the
minimum and maximum experimental frequencies attained, m, n and c are the
power-law exponents for the three different regions, and A is an empirical
constant.
Madiedo and Gallegos [82] have proposed a different empirical model that
describes the three regions of the relaxation spectra of oil-in-water emulsions
stabilized by a mixture of two sucrose esters with different HLB values:

a x m +13xn
H ()~)- p (20)

In this model, m, n and p are, respectively, the slopes of the transition, plateau
and pseudo-terminal regions, ~p is the pseudo-terminal relaxation time, and the
parameters a and [3 are given by the following relationships

Homm (21)
a= (m-n) )~o
103

1 n Ho (22)
- m)

Ho is the minimum value of the relaxation spectrum, which appears between the
transition and the plateau regions, and ~o is the relaxation time that corresponds
to this minimum. This model allows a smooth transition between the plateau and
pseudo-terminal regions, in contrast to the BSW-CW equation.
Studies carried out on model mayonnaise (emulsions using egg yolk as the only
emulsifier) demonstrated that an increase in oil concentration (75-80% wt.)
produces a wider linear viscoelasticity range and larger values of the dynamic
linear viscoelasticity functions [78]. Only the transition and plateau regions of the
relaxation spectra appeared. The same influence was found when the emu|sion
was stabilized by a nonionic surfactant, with a high solubility in the aqueous
phase, for emulsifier concentrations higher than 1% wt. [79]. In this case, a
dramatic decrease in the loss tangent as oil concentration increases (60-80% wt.)
was noticed in the low frequency range. This is due to the fact that a decrease in
oil concentration favours the development of a pseudo-terminal region in the
relaxation spectra of these emulsions, as previously mentioned.
Guerrero et al. [80] have studied model emulsions stabilized by a sucrose
palmitate. The tendency showed by the linear viscoelasticity functions was
practically the same in the whole oil concentration range studied (60-80% wt.),
although their relaxation spectra always displayed three different regions.
Furthermore, the values of the loss tangent were practically independent of oil
content. This fact allowed to superpose the linear viscoelasticity functions using
the plateau modulus as normalization factor and a horizontal shift factor only for
the less concentrated systems. The superposition obtained was good enough,
although some scatter in the plateau region was shown.
Similar results have been obtained by Franco et al. [10] studying salad dressing
emulsions (35%-55% wt. oil) stabilized by a mixture of egg yolk and a high HLB
sucrose ester. Thus, the linear viscoelasticity functions at different concentrations
can be superposed using the plateau modulus as normalization factor, which
increases with oil concentration. However, there is a slight, but significant
increase in the slope of the plateau region as oil concentration increases, which
may be related to the development of the entanglement network. On the contrary,
there is no significant influence of the oil concentration on the relaxation time that
defines the onset of the pseudo-terminal region of the relaxation spectra. The
addition of 2% wt. starch, although favouring the emulsion stabilization at lower
oil concentrations (down to 35% wt.), dampens the influence of oil content on the
104

linear viscoelasticity functions of these systems. Figure 8 shows the normalized


linear viscoelasticity functions and the values of the plateau modulus as a function
of oil concentration for starch-containing (2% wt.) emulsions. The BSW-CW
model fails to describe the behaviour of these last emulsions. Thus, there is a
sharp increase in the slope of the spectra at the highest relaxation times for the
less concentrated systems (35 and 40% wt.), associated with their poorer elastic
characteristics.
In relation to the influence of emulsifier concentration on the linear
viscoelasticity functions of concentrated food emulsions, an increase in low-
molecular-weight emulsifier concentration favours the formation of a
tridimensional structural network, that is the development of the plateau region of
the relaxation spectrum, yielding a significant decrease in the crossover frequency
at which G ' - G". The linear viscoelasticity functions for emulsions having low
emulsifier concentration cannot be superposed to the curves of the most
concentrated emulsions due to the above mentioned fact, which coincides with
negative values of the slope of the plateau region. These results have been found
for emulsions stabilized by a nonionic surfactant as the only emulsifier [79] and
for emulsions stabilized by a mixture of macromolecular and low-molecular-
weight emulsifiers [ 10].
It is worth pointing out that, contrary to the above-mentioned effects, Franco et
al. [10] found that an increase in the macromolecular emulsifier concentration
may produce a decrease in the values of the linear viscoelasticity functions.
Similar results were obtained by other authors, analyzing the surface
viscoelasticity of egg yolk films at the oil/water interface [18]. Although the
addition of macromolecular emulsifiers has been demonstrated to favour the
formation of an extensive structural network, it has been suggested that optimum
viscoelasticity was achieved at monolayer saturation. At higher protein
concentrations the interfacial viscoelasticity decreases because slip planes may
develop at the interface [83]. The protein displacement from the interface by the
low-molecular weight emulsifier may contribute to these results [84].
Although some methods have been developed to correct wall-slip on oscillatory
shear measurements [85] there is no evidence of this phenomena in food
emulsions, even when wall-slip caused by large velocity gradients was observed
[86].

3.3. Transient flow: Non-linear viscoelasticity


Different authors have studied the transient flow behaviour of concentrated
food emulsions [87-90]. All of them report the appearance of a stress overshoot
followed by stress decay to an equilibrium shear stress value. This behaviour has
105

been explained from two different points of view: thixotropy and non-linear
viscoelasticity.

101 . . . . . . . . n . . . . . . . . I . . . . . . . . I . . . . . . . . I . . . . . . . . t | i | ||111

10~

10_1
o o00 t~

10-2 . . . . . . . . . . n
10-3 10-2 10 -1 10 0 101 10 2 10 3

c0(rad/s)

3500 -

3400
f
3300 J

320(I I
J
J
3100 J
o
Z
O 3000

2900 I J

2800

3'5 ' 4'0 4'5 ' 5'0 ' 5'5

Oil content (% wt.)

Figure 8. a) Influence of oil concentration on the normalized viscoelasticity


functions for salad dressing emulsions (1"3 35% wt., O 40% wt., A 45% wt., V
50% wt., X 55% wt.). b) Evolution of the plateau modulus with oil content.
Adapted, with permission, from ref. 10.
106

Figoni and Shoemaker [87] and Gallegos et al. [89] tried to describe the stress
decay after the overshoot by using a kinetic model, sum of two first order kinetic
functions, as follows:

cy- cye - (%1 - eye1) exp (-klt) + (%2 - eye2) exp (-k2t) (23)

The kinetic constants of this model have been related to different shear-induced
processes: deflocculation and coalescence of oil droplets.
However, others authors [88, 90-92] have tried to describe the above
mentioned behaviour using a non-linear viscoelasticity model, the Wagner model
[93], a class of strain-dependent K-BKZ-type constitutive equation [94]. Mackley
et al. [95] and Madiedo et al. [96] have also used this model to predict the
steady-state flow of oil-in-water food emulsions.
Considering simple shear, and assuming that the memory function can be
separated into time and strain-dependent components (Figure 9), the Wagner
model would be reduced to the following equation:

t d G ( t t'
cy (t) - -~ - ) h (~,) y (t,t') dt' (24)
-oo dt' "
where G (t-t') is the linear relaxation modulus and h (y) is the damping function:

h (Y) = exp ( - k 7) (25)

being k an empirical parameter, which quantifies the level on non-linearity in the


material.
Nevertheless, Gallegos et al. [78, 90, 91] have also used a damping function
described by the Soskey-Winter model, which fits fairly well the experimental
results obtained with different types of emulsions:

1
h ( 7 ) - l + a 7b (26)

In the previous integral equation, the relaxation modulus may be described in


terms of a discrete set of Maxwell elements [91, 95].
In this case, the Wagner model results:

t ~gi -(t-t')/~.
cy (t) - - "f =~1~ e 1 h (Y) Y(t,t') dt' (27)
-ooi l
107

Introducing equation (25), the apparent viscosity for a shear rate, +, can be
calculated as

n gi Xi
= z (28)
i= 1 (l+kXi~;) 2

lO 4 ........ i ........ i ........ I ........ :

10 3

10 2

~" 101 El

i D~ [] nclc~tan_,.,_~
o
~ lOo

10 -]

10 ~ ........ i ........ i ........ i ........ i


10 -] 100 101 102 103

time(s)

Figure 9. Linear and non-linear relaxation moduli for a salad dressing emulsion,
( i G(t) from H(X), X G(t) from the Ninomiya-Ferry approximation).

In other cases [78, 90] a power-law equation was used, yielding the following
integral equation:

cy(t) - I m C ( t - t') - m - 1 h (y) ~,(t,t') dt' (29)


--o0

Figure 10 shows the experimental and calculated values of R(t) (ratio between
the shear stress after instantaneous imposition of shear rate, ~(t), and the steady-
state stress, r~(oo)) for a transient flow test carried out at 0.6 s-~ As can be
observed, the fit of the model is acceptable at low shear rates. Higher shear rates
produce increasing differences between experimental and calculated values.
108

Other authors [96] have used a continuous relaxation spectrum, calculated from
the experimental values of the dynamic moduli, using the Tikhonov regularization
method [97, 98], to predict the steady-state flow behaviour of different emulsions.
In this case the Wagner model results

t oo e_(t_t,)/Z (30)
cr (t) - - ~ ~ H (X) h (~,) qt(t,t') dlnkdt'
--o0 --o0

and the steady-state viscosity may be calculated from the following integral
equation:

oo X H(:X)
11 ( f ) - [ d lnX (31)
-oo (1 + kXf)2

2~F
~ = 0 . 6 s -1

1.5

1.O~- III~ X~ - Xll X--M - - X T X - - J - - X

0.5

0.0 , I ~ I ~ I ~ I ~ I ~ I

0 25 50 75 100 125 150

time (s)

Figure 10. Relationship between transient and steady-state stress values as a


function of shearing time for a salad dressing emulsion, (11 experimental values,
X Wagner model prediction).

In addition to the rheological tests, optical observations of the sheared materials


were also reported using a purpose built optical shearing cell. Wall-slip, micro-
109

domain movement, chaining and changes in droplet size distribution, were all
observed under different shear conditions and depending on the nature of the
emulsifier used.
The authors conclude [96] that the Wagner model gives acceptable results, in a
relatively wide shear rate range for vegetable protein-stabilized emulsions. The
overestimation of the apparent viscosity at low shear rates is thought to be due to
sample slipping between the rheometer plates. On the contrary, the model largely
fails for a whole egg-stabilized commercial mayonnaise. This difference is
explained on the basis that the shear rate is inducing significant changes in the
droplet size distribution of mayonnaise throughout practically the entire
rheological range studied. On the contrary, the vegetable protein-stabilized
emulsion only shows a significant change in droplet size distribution at very high
shear rates. The authors remark this fact may indicate that the applicability of the
Wagner model is closely related to shear-induced structural changes in emulsion
microstructure.

4. INFLUENCE OF PROCESSING ON FOOD EMULSION


RHEOLOGY

Emulsification is a complex unit operation in which many variables influence


the processing and the final rheological characteristics of the product. The
manufacture of emulsions is an energy-intensive and highly dynamic process,
which usually requires the application of mechanical energy. The two critical
steps are the consecutive disruption of droplets and their coalescence, both of
which are favoured by an intense agitation. Consequently, the improvement of the
emulsification process requires the measurement of the droplet size of the
dispersed phase and its polydispersity, as well as knowledge of its rheological
properties [28, 32].
The emulsification process may be greatly affected by the viscous and
viscoelastic properties of the continuous phase at which the disperse phase is
added. Thus, Gallegos et al. [64] have studied the influences that temperature,
agitation speed and emulsification time cause on the droplet size distribution and
viscoelasticity of vegetable oil-in-water emulsions stabilized by a well-known
polyoxyethylene nonionic surfactant, which produces a low-viscosity micellar
continuous phase, and manufactured using an anchor impeller.
Figure 11 shows the droplet size distribution (DSD) curves as a function of the
agitation speed of the anchor impeller (N). A well-pronounced maximum at
droplet diameters larger than 5 ~tm and a very smooth maximum at lower sizes
are noticed, although the curves move to lower sizes as N increases. Only one
110

maximum was found when the emulsion was prepared with a rotor-stator turbine
(4000 rpm). These DSD curves determine the linear viscoelastic response of the
emulsions. As can be observed in Figure 12, the systems prepared with the
anchor impeller show higher values of the loss modulus at low frequencies,
althought G' and G" curves crossover at a frequency O3c.Above this characteristic
frequency, a plateau region in G' develops as agitation speed increases. Similar
behaviour was also found by Tadros [37] for weakly flocculated emulsions. On
the contrary, the emulsion prepared with the rotor-stator turbine only shows a
plateau region, similar behaviour to that found for highly concentrated o/w
emulsions stabilized by a macromolecular emulsifier [78]. This behaviour
corresponds to an emulsion having much smaller sizes and narrower droplet
distribution than those prepared with the anchor impeller.
An increase in temperature favours the coalescence process, leading to larger
droplet sizes and broader distributions. In the same way, the plateau region in G'
tends to vanish and the crossover frequency increases. On the contrary, an
increase in the emulsification time enhances the development of the plateau
region.

l0

O. ~ ~

Diameter (~m)
Figure 11. Droplet size distributions for emulsions containing NPE-PEG-10,
prepared with different process conditions, (m 1oo rpm, 9 150 rpm, A 200 rpm, V
250 rpm, I"! 300 rpm, 4000rpm). Reproduced, with permission, from ref. 64.
111

Different results are obtained if a sucrose ester nonionic surfactant, which


forms a gel-like structure in the continuous phase for a wide range of
concentrations and temperatures [98], is used as emulsifier. An increase in
agitation speed or emulsification time also produces a decrease in droplet size
and polydispersity [64]. However, an increase in the agitation speed produces a
decrease in the values of the dynamic viscoelasticity functions, because the gel-
like structure tends to vanish and the continuous phase of the emulsion becomes
less viscoelastic. In fact, the viscoelastic properties of these emulsions depend on
the balance between the formation of a larger interfacial surface and the
breakdown of the gel-like structure of the continuous phase during processing.

1000

100

IIIIQIll
10 ~ o ~ O O Q o ~ ~ ~~
m o
u oo
oo
oQ
~oooooo~
Oo O DB
0
Oo []
[]
9
9
O~ o 9
mm
m

~ ........ o.o ' ........ o11 ........ ........ ........ oo' ......

Figure ]2. Evolution of the storage and loss modu]i with frequency, for an
emulsion containing NPE-PEG-] 0 as emulsifier, prepared with different process
conditions, (I-I ]00 rpm, O150 rpm, V 250 rpm, X 300 rpm, + 4000 rpm).
Reproduced, with permission, from ref. 64.

In relation to the manufacture of protein-stabilized emulsions, protein


denaturation by heat tends to improve emulsifying and foaming capacity by
enhancing macromolecular flexibility and surface hydrophobicity [5]. A large
number of researchers have studied the emulsifying capacity of whey proteins by
heating the solution prior to the addition of the oil phase [84, 99-105]. This
thennal denaturation favours emulsion stability and improves the rheological
112

properties, both effects being related to the formation of a gel structure. Similar
results have been obtained by Gallegos et al. [64], using vegetable protein as
emulsifier.

6000 , , , , , , , , , , , , , , 100

5000
o
9 m u m m n
80
oo 9 %

4000 9 00000 9
,,I 9
l
/ 9 ~ O9
3000 %
o'
o~
uNNN 9
9 9 []
",p
9 9 O O0 9 DO OOO 0 40
20OO 9 O go O 0
9 n []
o
9 [] 0 0 0 0 O0 0 O0
[]
1000
9
[] O O 0
[]
0
O0 0 20
9 [] ooO o @@@ @@@@ @@@ @
O00 .A@

0_ - - ~ B ~ . ? * *, * I , I , I , I
0 ~" ~ 2 ~ v ~. 40 60 80 100 120 140

time (min)

Figure 13. Effect of the lecithin/protein molar ratio (R) on developing storage
modulus (1 Hz) of whey protein emulsions. Lecithin/protein molar ratio (R): II
R=I6, 9 R=6, I"1 R=4, O R=2, 9 R=0. Dashed line shows temperature
programme. Adapted, with permission, from ref. 102.

Dickinson [106, 107] has also studied emulsion gelation by covalent cross-
linking of proteins using an enzyme. The resulting values of the storage modulus
are compared to those obtained by heat treatment of the emulsion. Furthermore,
data for some protein gels without emulsion droplets were also obtained. It is
concluded that, whereas the enzyme treatment produces stronger protein gels than
the heat treatment, the opposite is the case for emulsion gels. This is attributed to
topological constraints imposed by the permanent nature of the covalent
crosslinks in the enzyme-set system which restricts further reinforcement of the
network structure [106]. However, the G' values for enzyme-set emulsions gels
are slightly less frequency-dependent than those of equivalent heat-set emulsion
gels. This is explained by the fact that enzyme-set system theology is more
113

similar to that of a classical polymer gel with "chemical" bonds, whilst the heat-
set system rheology is typical of a "physical" gel with breakable bonds.

10 3 ' , , " ..... I . . . . . . . . ~ . . . . . . . . I , , ,- ..... I . . . . .

r 102

BSW-CW model

i01 , . . . . . ,,l , , ...... { , , , ,,,,,l . . . . . . . . ,,,{

10-2 1 0 -1 i0 ~ 101 10 2 10 3

10 a . . . . . . "1 . . . . . ' " I ' ' ' ' ' ' " 1 ' ' ' . . . . 'i ' ' ' ' ' ' " -

lO 3 i::1 rn

t~

io ~

:2

101
BSW-CW model

10 0 0 "z . . . . . . . . , . . . . . . . . I . . . . . . . . , . . . . . . . . , . . . . . . . .
1 10 -1 10 ~ 101 10 z 10 s

)~ (s)

Figure 14. Influence of the processing variables on the relaxation time spectra of
salad dressing emulsions, a) Influence of emulsification temperature: 1"I 20~ 9
without thermal control, A 50~ b) Influence of emulsification time and agitation
speed: [21 8000rpm-5min, O8000rpm-3min, A 5000rpm-5min, V 5000rpm-3min.
Reproduced, with permission, from ref. 8.
114

The same author has also investigated the effect of low-molecular weight
emulsifiers, added after emulsification but prior to thermal processing, on the
linear viscoelasticity functions of protein-stabilized emulsions [84, 108, 109]. The
results indicate that the rheology of the emulsion gel, produced by in situ heat
processing, is significantly influenced by the surfactant used. Thus, Figure 13
shows the evolution of the storage modulus during the thermal processing, as a
function of temperature, for different lecithin/protein molar ration, R. These
results are consistent with lecithin-protein complexation at the oil-water interface
and in the bulk aqueous phase.
Much more complicated behaviour was found when a polyoxyethylene sorbitan
monolaurate was used. In this case, the values of the storage modulus present a
maximum for R = 1, which is related to a partial displacement of protein from the
oil-water interface by the water-soluble surfactant [110].
The influence of processing on emulsions stabilized by a mixture of
macromolecular and low-molecular-weight emulsifiers has been also studied by
Franco et al. [8]. Figure 14 shows the linear relaxation spectra of emulsions,
prepared in a pilot-plant colloidal mill or with a rotor-stator turbine (lab-scale), as
a function of rotational speed, residence time and temperature of emulsification.
As can be observed the slope of the plateau region increases with the processing
variables, because of the development of an entanglement network. This is related
to a decrease in mean droplet size and polydispersity of the emulsion, yielding
stronger inter-droplet interactions [28].

5. CONCLUDING REMARKS

The rheology of food emulsions is mainly dependent on the strenght of interdroplet


interactions. Dilute emulsions (i.e. milk) have a low-viscosity Newtonian behaviour.
On the contrary, concentrated food emulsions show gel-like rheological
characteristics. This behaviour can be attained by increasing the disperse phase
volmne fraction, or by different flocculation mechanisms (heat denaturation of
proteins, covalent cross-linking of proteins, etc.). However, many other structural
parameters also influence the rheological (viscous and viscoelastic) response of
emulsions (i.e. droplet size, polydispersity, droplet deformability, etc.). In qualitative
terms, the influence of all of them on the rheology of emulsions is well understood, as
well as their dependence on the emulsion processing. However, because of the lack
of good model systems, i.e. with different degrees of polydispersity, the quantitative
understanding of the relationship among stability/microstructure/rheology/processmg
of food emulsions represents a very considerable challenge for the future.
115

REFERENCES

1. W. Clayton (ed.), Theory of emulsions and emulsification, Churchill, London,


1923.
2. P. Sherman, (ed.) Emulsion Science, Academic Press, London 1968.
3. P. Becher (ed.), Encyclopedia of emulsion technology, II. Marcel Dekker, New
York 1985.
4. D.G. Dalgleish, In Emulsions and emulsion stability, SjOblom, J. (ed.), Marcel
Dekker, New York 1996, 287.
5. E. Dickinson, G. Stainsby, Food Technol., 41 (1987), 75.
6. E. Dickinson, An Introduction to Food Colloids. Oxford University Press.
Oxford 1992.
7. J.M. Franco, A. Guerrero, C. Gallegos, Grasas y Aceites, 46 (1995), 108.
8. J.M. Franco, A. Guerrero, C. Gallegos, Rheol. Acta, 34 (1995), 513.
9. J.M. Franco, M. Berjano, A. Guerrero, J. Mufioz, C. Gallegos, Food Hydrocoll.,
9 (1995), 111.
10. J.M. Franco, M. Berjano, C. Gallegos, J. Agric. Food Chem., 45 (1997),
713.
11. D.C. Clark, P.J. Wilde, D.R. Wilson, R.C. Wunsteck, Food Hydrocoll., 6
(1992), 173.
12. E. Dickinson, In Interactions of Surfactants with Polymers and Proteins,
Goddard, E.D., Ananthapadmanabhan, K.P., (eds.) CRC Press, Boca Raton
1993, 295.
13. S. Matsumoto, In Macro and microemulsions: theory and applications, Shah,
D.O. (ed.) ASC Symposium Series, Washington 1985, 272.
! 4. S. Matsumoto, J Texture Stud., 17 (1986), 141.
15. B. De Cindio, D. Cacace, Int. J. Food Sci. Technol., 30 (1995), 505.
16. E.E. Uzgiris, H.P.M. Fromageot, Biopolym., 15 (1976), 257.
17. E. Dickinson, E.W. Robson, G. Stainsby, J. Chem. Soc. Faraday Trans., 79
(1983), 2939.
18. V.D. Kiosseoglou, P. Sherman, Colloid Polym. Sci., 26 (1983), 520.
19. E. Dickinson, S.E. Rolfe, D.G. Dalgleish, Food Hydrocoll., 2 (1988), 397.
20. D.G. Dalgleish, Colloids Surf., 46 (1990), 141.
21. E. Dickinson, J.A. Hunt, D.G. Dalgleish, Food Hydrocoll., 5 (1991), 403.
22. A. Kondo, K. Higashitani, J. Colloid Interface Sci., 150 (1992), 344.
23. J.L. Klemaszewski, K.P. Das, J.E. Kinsella, J. Food Sci., 57 (1992), 366.
24. B.V. Derjagum, L. Landau, Acta Physiochem., 14 (1941), 633.
25. E.J.W.Verwey, J.T.G. Overbeek, Theory of the Stability of Liophobic Colloids,
Elsevier, Amsterdam 1948.
116

26. W. Heller, T.L. Pugh, J. Chem. Phys., 22 (1954), 1778.


27. D.H. Melik, H.S. Fogler, In Encyclopedia of Emulsion Technology, III, Becher,
P. (ed.) Marcel Dekker, New York 1988, 3.
28. R.R. Rahalkar, In Viscoelastic Properties of Food, Rao, M.A. and Steffe J.F.
(eds.), Elsevier, London 1992, 317.
29. P. Sherman, J. Pharm. Pharmacol., 16 (1964), 1.
30 E. Dickinson, Colloids Surf., 42 (1989), 191.
31 P. Sherman, J. Colloid Interface Sci., 24 (1967), 67.
32 A.R. Carrillo, J.L. Kokini, J. Food Sci., 53 (1988), 1352.
33 M. van den Yempel, J. Colloid Sci., 13 (1958), 125.
34 E. Vernon-Carter, P. Sherman, J. Dispersion Sci. Technol., 2 (1981), 399.
35 J.A. Hunt, D.G. Dalgleish, J. Agric. Food Chem., 42 (1994), 2131.
36 R. Pal., E. Rhodes, J. Rheol., 33 (1989), 1021.
37. Th.F. Tadros, In First World Congress on Emulsion, vol. 4, Paris 1993,237.
38 E. Dickinson, In Gums and Stabilisers for the Food Industry, vol. 4, Wedlock,
D.J., Williams, P.A. (eds.). I R Press, Oxford 1998, 244.
39. Y. Cao, E. Dickinson, D.J. Wedlock, D.J., Food Hydrocoll., 5 (1991), 443.
40. Dickinson, In Progress and Trends in Rheology, IV, Gallegos, C. (ed.).
Steinkopff, Darmstadt 1994, 227.
41. A. Parker, P.A. Gunning, K. Ng, M. Robins, Food Hydrocoll., 9 (1995),
333.
42. R.L. Scott, J. Chem. Phys., 17 (1949), 279.
43. R. Pal, Int. J. Multiphase Flow, 15 (1989), 1011.
44. R. Pal, AIChE J., 39 (1993), 1754.
45. H.A. Barnes, Colloids Surf. A, 91 (1994), 89.
46. P. Sherman, in Encyclopedia of Emulsion Technology, Becher, P. (ed.),
Marcel Dekker, New York 1983, 405.
47. A. Einstein, Investigations on the Theory of the Brownian Movement, Dover,
New York 1906.
48. G.I. Taylor, Proc. Royal Soc., A138 (1932), 41.
49. G.K. Batcherlor, J. Fluid Mech., 83 (1977), 97.
50. M. Mooney, J. Colloid Sci., 6 (1951), 162.
51 H. Eilers, Kolloid-Z, 97 (1941), 313.
52 H. Eilers, Kolloid-Z, 102 (1943), 154.
53 R. Roscoe, J. Appl. Phys., 3 (1952), 267.
54 S.H. Maron, P.E. Pierce, J. Colloid Sci., 11 (1956), 80.
55 I.M. Krieger, T.J. Dougherty, Trans. Soc. Rheol., 3 (1959), 137.
56 N.A. Frankel, A. Acrivos, Chem. Eng. Sci., 22 (1967), 847.
57. J.S. Chong, E.B. Christiansen, A.D. Baer, J. Appl. Polym. Sci., 15 (1971),
117

2007.
58. R. Pal, J. Rheol., 36 (1992), 1245.
59.Y. Otsubo, R.K. Prud'homme, Rheol. Acta, 33 (1994), 29.
60. P. Partal, A. Guerrero, M. Berjano, J. Mufioz, C. Gallegos, J. Texture Stud.,
25 (1994), 331.
61. J.M. Franco, C. Gallegos, H.A. Barnes, J. Food Eng., submitted for
publication.
62. R. Pal, E. Rhodes, J. Colloid Interface Sci., 107 (1985), 301.
63. F.L. Saunders, J. Colloid Sci., 16 (1961), 13.
64. C. Gallegos, M.C. Sfinchez, A. Guerrero, J.M. Franco, In Rheology and Fluid
Mechanics of Nonlinear Materials, Siginer, D.A. and Advani, S.G. (eds.),
ASME, New York 1996, 177.
65 A. Yoshimura, R.K. Prud'homme, J. Rheol., 32 (1988), 53.
66 J. Carnali, H.A. Barnes, J. Rheol., 34 (1990), 841.
67 R. Pal, Chem. Eng. Commun., 98 (1990), 211.
68 R. Pal, Chem. Eng. Sci., 52 (1997), 1177.
69 H.A. Barnes, J. Non-Newtonian Fluid. Mech., 56 (1995), 221.
70 S. Matsumoto, P. Sherman, J. Texture Stud., 12 (1981), 243.
71 N. Gladwell, R.R. Rahalkar, P. Richmond, J. Food Sci., 50 (1985), 1477.
72 H.J. Rivas, P. Sherman, J. Texture Stud., 14 (1983), 251.
73 H.J. Rivas, P. Sherman, J. Texture Stud., 14 (1983), 267.
74. V.D. Kiosseoglou, P. Sherman, J. Texture Stud., 14 (1983), 397.
75 N. Gladwell, R.R. Rahalkar, P. Richmond, Rheol. Acta, 25 (1986), 55.
76 J.M. Madiedo, J. Mufioz, M. Berjano, C. Gallegos, In Progress and Trends
in Rheology IV, Gallegos C. (ed.) Steinkopff, Darmstadt 1994, 281.
77. T. van Vliet, J. Lyklema, M. van den Tempel, J. Colloid Interface Sci., 65
(1978), 505.
78. C. Gallegos, M. Berjano, L. Choplin, J. Rheol. 36 (1992), 465.
79. M.C. Sfinchez, M. Berjano, C. Gallegos, Afinidad, submitted for publication.
80. A. Guerrero, P. Partal, M. Berjano, C. Gallegos, Prog. Colloid Polym.
Sci., 100 (1996), 246.
81 M.E. de Rosa, H.H. Winter, Rheol. Acta, 33 (1994), 220.
82 J.M. Madiedo, C. Gallegos, Applied Rheol. 7 (1997), 161.
83 D.E. Graham, M.C. Phillips, J. Colloid Interface Sci., 76 (1980), 240.
84 E. Dickinson, S.T. Hong, J. Agric. Food Chem., 43 (1995), 2560.
85 A. Yoshimura, R.K. Prud'homme, J. Rheol., 32 (1988), 575.
86 L. Ma, G.V. Barbosa-Cfinovas, J. Food Eng., 25 (1995) 397.
87 P.I. Figoni, C.F. Shoemaker, J. Texture Stud., 14 (1983), 431.
88 O.H. Campanella, M. Peleg, J. Rheol., 31 (1987), 439.
118

89. C. Gallegos, M. Berjano, F.P. Garcia, J. Mufioz, V. Flores, Grasas


Aceites, 39 (1988), 254.
90. C. Gallegos, M. Berjano, A. Guerrero, J. Mufioz, V. Flores, J. Texture
Stud., 23 (1992), 153.
91. C. Gallegos, J.M. Franco, Les Cahiers de Rheologie, 14 (1995), 107.
92. P. Partal, Ph.D. Thesis, University of Seville, Seville, 1995.
93. M.H. Wagner, Rheol. Acta, 15 (1976), 136.
94. R.G. Larson, Constitutive Equations for Polymer Melts and Solutions,
Butterworths, Boston 1988.
95. M.R. Mackley, R.T.J. Marshall, J.B.A.F. Smeulders, F.D. Zhao, Chem.
Eng. Sci., 49 (1994), 2251.
96. J.M. Madiedo, C. Bower, M.R. Mackley, C. Gallegos, Grasas Aceites,
submitted for publication.
97. J.M. Madiedo, Ph. D. Thesis, University of Seville, Seville, 1996.
98. J.M. Madiedo, J. Mufioz, C. Gallegos, In Rheology and Fluid Mechanics of
Nonlinear Materials, Siginer, D.A. and Advani, S.G. (eds.), ASME, New York
1996, 151.
99. N.K.D. Kella, J.E. Kinsella, Biochem. J., 255 (1988), 113.
100 J.E. Kinsella, D.M. Whitehead, Adv. Food Nutr. Res., 33 (1989), 343.
101 R. Jost, F. Dannenberg, J. Rosset, Food Microstructure, 8 (1989), 23.
102. Y.L. Xiong, J.M. Aguilere, J.E. Kinsella, J. Food Sci., 56 (1991), 920.
103 R.A. Yost, J.E. Kinsella, J. Food Sci., 57 (1992), 892.
104. R.A. Yost, J.E. Kinsella, J. Food Sci., 58 (1993), 158.
105. D.J. McClements, F.J. Monahan, J.E. Kinsella, J. Texture Stud., 24 (1993), 411.
106. E. Dickinson, Y. Yamamoto, J. Agric. Food Chem., 44 (1966), 1371.
107. E. Dickinson, In Proceedings of 1st Intern. Symposium on Food Rheology and
Structure, Zurich 1997, 50.
108. E. Dickinson, Y. Yamamoto, Food Hydrocoll., 10 (1996), 301.
109. E. Dickinson, Y. Yamamoto, J. Food Sci., 61 (1996), 811.
110. J. Chen, E. Dickinson, J. Sci. Food Agric., 62 (1993), 283.
119

WORMLIKE MICELLAR SURFACTANT SOLUTIONS:


RHEOLOGICAL AND FLUID MECHANICAL ODDITIES

R. Steger a a n d P.O. B r u n n b

~Rheotest Medingen GmbH, ROdertalstrafle 1, D-01458 Medingen b. Dresden,


Germany

bUniversitgit Erlangen-Niirnberg, Lehrstuhlfiir Strfmungsmechanik, Cauerstr. 4,


D-91058 Erlangen, Germany

1. INTRODUCTION
In aqueous surfactant solutions the surfactant molecules form aggregates
(termed micelles) at some critical concentration cmc (= critical micelle
concentration). Its exact value depends on the size of the hydrophobic part of the
molecule. The hydrophilic group plays also a role, in particular whether it is ionic
or nonionic. At concentrations above cmc surfactants are present in the form of
micelles and monomers. Micelles are fragile dynamic objects which are
constantly formed and destroyed by the addition and loss of monomers. This
permanem exchange of material is in real thermodynamic equilibrium.
Usually the miceUes formed above cmc are of globular shape and remain so up
to rather high surfactant concentrations. Yet, some systems exist for which, even
at high dilution, a second critical concentration, termed ct. (transition
concentration), ct > cmc, exists at which rod- (or worm-) like micelles are being
formed. This usually happens when additives like oppositly charged surfactants,
organic counterions or uncharged components like aromatic hydrocarbons are
added [1,2].
Thus, above c, the rod- (or worm-) like micelles are thermodynamically more
stable than globular micelles. The dynamic nature of their creation and breakup
implies that micellar solutions cannot be regarded as suspensions of well-def'med
particles with a given size. Some of the kinetics, which govern these processes, is
well understood (e.g.[3]).
As soon as rod like micelles are present in the solution, viscoelastic behavior is
encountered in viscometric measurements [4,5]. In addition, the solutions then
120

show the ability of drag reduction in turbulent pipe flow [2,6,7] as well as
enhanced resistance in porous medium flow [8,9].
The system studied in this paper is the surfactant N-cetyl-N,N,N-
trimethylammoniumbromide (C~6TMA-Br) with sodiumsalicylate (Na-Sal) as
counterion (equimolar solution). Most of the results reported will refer to a
concentration of 1000 weight-ppm (___-2.3 mmol), a concentration well above ct
(at least in the temperature range tested by us, 25~ _< T _<50~ [ 10].

2. RHEOLOGY

2.1. Experimental
The viscometric results reported have all been obtained by the commercial
Couette viscometer Haake CV 100 - RV20. It is of the CR type (constant rate),
the outer cylinder being the rotating one. Details about the measuring systems
employed (dimensions) have been given before [ 10].

10 -~
Surfactant o 4 h
I000 p p m 9 20 rain
T=35"(3 0 4 min
DA45 x 1 min
EiO000D~%
o ~ i
1 0 -2

-,==4

0
r
o o 0
-,,...4

-3 XXXXX x
10 i .v 9 9 i~ 9 in|
.~ . v . . . .v' I ~

1
." . r 9 9 9 '~l

10
= ' " 9 9 ~

10 z
l
lb a

Shear Rate in 1/s


Figure 1. Influence of the sweep time (varying between 1 minute up to 4 hours)
on the apparent viscosity as a function of apparent shear rate (R / R~ = 1.078) o
121

Using the software provided by the manufacturer a sweep test reveals time
dependent behavior up to a sweep time of four hours (Sweep times longer than
this do not change anything). Starting essentially from the base line of the solvent
viscosity, 11~, the apparent shear viscosity 11 shows sudden shear thickening at a
critical apparent shear rate ~, ~, from 11 ~ 11, up to some value rl p. The apparent
viscosity will then stay essentially at that plateau value 1"1p up to an apparent
shear rate, termed ~,,.
Increasing the shear rate fin-ther leads to shear thinning behavior (see figure 1).
On a double-logarithmic plot the slope for ~ > ~,, is minus one, which in terms of
shear stress x means that there exists a critical shear stress ~ ~ ( = 11p ~,~), at
which rl ,,jumps" from its plateau value 1"1p to some lower value (see figure 2).

10-1 . . . . . . . . .

- Surfactant o 4h
. 1000 ppm 2 0 min
. T=35 C r 4 min
13., DA45 x 1 min

~ r-.,I

1 0 "2 },
r,/3 r
0
0
~
o 0 x
>. - ~ r

xXXXX X X
'w- r w: i : w T C I
IO- - .... . . ...-.,,= , ...... . , , . , . , ....... , - , - . . . . . ,
I 0-3 10 2 10-I I lO

Shear Stress T in Pa

Figure 2. The results of figure 1 as a function of shear stress


122

It is of interest to note that while the quantities ~ , rl p and ~ are time


dependent (as far as the sweep time is concerned) ~ r is not. Thus, the rheological
characteristics of the solution seem to be i) time dependend behavior (which has
been shown to be rheopectic [10], ii) a sudden increase of rl at some (time
dependend) apparent critical shear rate ~ and iii) a sudden decrease of 11at
some ( time independent ) shear stress ~ . In order to avoid a sweep time
dependence all viscometric results to follow have been obtained with a sweep
time of more than four hours.
For completeness one should note that the sudden increase in flat some
apparent critical shear rate ~ is found for many surfactant solutions in the
concentration range where rod-like micelles dominate [4,5]. For that reason it is
not astonishing to see that it has received considerable attention in the literature.
The picture that at "~ a new phase is suddenly ,,created", the so called s_hear-
induced-s_tate (SIS) is intriguing since it allows one to consider the q , ~ rl o

Surfactant a DB45
co 1000 ppm * ME4-5
T=35 C A ME46
t3-, 10-1

9r-',,I o000~
o o~o
t~ -2
10
" O
~ & & A A A A A
Q A
9 r-,,,ll D
0 o
0 -3 kAk
r 10 - dk' **~0~ I~II AAA A A
r~
9r,-,I
b,.

-4
10 11 - - It "l-- 9 : I I-i n1011
! .... Iv '" Iv-~-I --IV -II ~11 i ~ 2 - -W~ '' W-- 9 ~ W W W31~01

Shear Rate ~ in I/s

Figure 3. Influence of the relative gap width ~con the apparent flow curve

A: K=1.037, *" K=1.078; ~ K=1.176


123

jump on the basis of a phase transition (like a sol-gel phase transition). Readers
interested in this subject are referred to the excellent papers of H. Hoffmann and,
repectively, H. Rehage [5,11,12].
The viscometric results of our solution show a dependence of 1"1upon the"type
and the size of the measuring system used. For example, by using three systems
of different relative size (indicated by the fact that K:, the ratio of the outer to
inner cylinder, 1<-R o /R] > 1, differs) we find that )~ increases while "iqp
simultaneously decreases with decreasing 1< (figure 3). If for 1<~ 1 there is any
jump (rl p > 1"1s) at all then this jump is bound to be rather small.
Interesting is another, and totally unexpected, fact namely that there is also a
dependence upon the actual gap width. Figure 4 shows that three systems with
the same K - 1.078 furnish (rather unsystematic) results. Two Mooney Evert
systems (termed ME) imply an increase of 11with increasing gap width, the
double gap device (DA) contradicting that somewhat, since the gap widths
(1.375mm and 1.63 ram) are between the others. We conclude from these results
that the apparent flow curves depend upon the details of the measuring system
used.

10-1 _

Surfactant o DA45
1000 oppm * ME45
T=35 e ME30
ooo
12., . oooooo O
%'*******~o ~
17~ 10 -2- Oo
B I,,,,-f
0
~ ~ oo o
0 0
B
~r,,,I
-3
10
O
to
r~
~

10 -40-1 1 ...... 'i' ' ' ' ' ' ~' 10'" . . . . . "i~) 2 ' ' "' '"10 3

Shear R a t e ~ in 1//s
Figure 4. Influence of actual gap width AR- R - R, on the apparent flow curve
o. AR=l.085 mm; , AR=l.63 mm; o- AR=1.375 and 1.63 mm
(K = 1.078 in all three cases)
124

2.2. Theoretical considerations


In a concentric cylinder viscometer (Couette type) the outer cylinder (radius Ro)
rotates with angular velocity f2 relative to the inner one, (radius R~, height h).
The torque M, needed to hold the inner cylinder at rest, is measured. If
viscometric flow prevails the shear surfaces are concentric cylinders. Thus, the
shear rate ) will be a function of r, r being the distance from the axis of rotation,
do (1)
dr
Here o = o(r) is the angular velocity of a cylindrical surface of radius r. A
torque balance reveals that the shear stress x varies like r-2, so that

with "r., t h e shear stress at the inner cylinder. Expressed in terms of measurable
quantities it is given by
M (3)
x ~ = 27r.hRiZ

For the measuring systems employed by us the shear stress thus varies across the
gap width by 7.5, 16.2 and by 38.3 % (~: =1.037, 1.078 and 1.176).
Assuming the fluid to adhere at a solid surface the other measurable quantity,
namely f~, is given by

Ro do 1 "ci (4)

Rj dr 2 r" 2 T,
Ti/

It is important to note that this implies f~ - f~('t,, "t o ) = f~('r,,, K 2 ) .


Principally it is possible to invert this relation (and thus to get the shear
viscosity rl ) if experiments are performed with different systems (different K's)
using the same shear stress at the outer cylinder, i.e. fixed x o = "r., / K : . For in that
case we get from equation (4) the shear rate at the inner cylinder, ~,
Of~ (5)
c31ogK "tO
125

Thus, in a plot of f2 versus log~: the slope will be ~,. It is clear that such a
procedure will be extremely time consuming (to obtain from experimental data
the slope rather accurately will require many measurements) and we know of no
attempt that this procedure has ever been tried.
An alternate way is to employ a mean value theorem to equation (4) with the
result

2K2 ~ (6)

K2 -1 xi

where u is the appropriate mean shear stress, x o _<u _<x,. If the fluids rheology is
known then u can be calculated. Note that one can write u / x,= ~'~, I < ~ <_K .
This result implies for the shear viscosity the relation
q-rl(T)-~=x C M (7)
f2

with
9 -1 (8)
C _ _

4n~:2hR 2

a purely geometrical quantity.


Thus, measurements of M and f2 suffice to determine rlcorrectly.
Unfortunately ~(or ~ ) is not known and it is necessary to a priori choose a
particular value for it. Commercial instnmaents, and also the one used by us, use
for the shear rate the expression [13]
2~:2 (9)
Ira K 2 -- 1

which by equation (6) corresponds to the choice u = x,. While correct for a
Newtonian fluid as well as in the limit of infinitely small gap width (~: ~ 1 ) it
will for f'mite gap widths in general be incorrect. Not the fluid-specific shear
viscosity rl but an apparent viscosity rio, which depends on relative gap width as
well, is the consequence,
126

M (10)
f2

All experimental results reported thus far are apparent ones (since all results
reported in the last chapter are apparent ones we have for simplicity omitted the
subscript a ~ w h i c h emphasizes this fact ~ on q and ~. When we now compare
the theoretical results with experimental ones the latter results will get this
subscript) so that the observed dependence upon K (see figure 3) was to be
expected. Since these results point towards decrease in the height of the jump for
decreasing K (see figure 3) and since ~t and ~ coincide in the limit K ~ 1, the
question arises whether the jump seen in rio (see figure 3) is indeed consistent
with a jump in rl at some critical rate ~,~.
Such a jump of q implies a jump of rio at a critical apparent shear rate ~ a~, which
by equation (4) is given by

K 2 InK2 (11)
r - - ~gC
1( 2 --1

Since ~cShOuld be a fluid specific critical shear rate, ~ , according to this


relation, should increase with ~c. Experimentally, the opposite is found (figure 3).
The rheological relations presented do not admit a jump of rl~ at some critical
shear stress x . This, however, is observed (see figure 2). The way out of this
dilemma is to abandon (at least) one of the assumptions on which the formulas
rest, which are i) the flow is visco-metric (with concentric cylinders as shearing
surfaces) and ii) there is no slip. If the flow is not viscometric then none of our
theoretical relations are valid. On the other hand if the flow is viscometric but
slip occurs then equation (4) has to be modified,

-- ~'~]f + ( ~'] s i + ~'~'2s o )


(12)

with f~e given by equation (4) and f2~(f~,o) the angular slip velocity at the inner
(outer) cylinder. The relations 4-10 would involve the rheologically relevant
127

angular velocity f21 -f2~(x,,K) rather than the angular velocity f2of the outer
cylinder. To obtain it, the slip contribution has to be known.
Making the usual assumption that the angular slip velocity can be written as
shear stress dependent slip velocity u divided by the radius of the corresponding
cylinder, i.e.

f2,~ + f~,o = ~ ,[ , 2]
u,(x~) +-u,(x~~: / K )
(13)

shows that slip effects become more and more important the smaller system
(small R~) is. This was to be expected and the experimental results presented in
figure 4 for the two Mooney -Evart systems (ME) are in accord with this feature.
That the results for the double gap device (DA) do not follow this vend could
possibly be of different origin. Labeling for the DA system the radii of the
cylinders (in increasing order)/~ to R,, we have (with ~c=/~ //~ = R~ / R3)
~"~-- ~"~fl (~1, K) + (~'~S1 + ~"~$2) (14 a)
for the 1-2 gap, and simultaneously

~'-'~-- ~"~/3($3, K) + (~"2S3 + ~"2S4 ) (14b)

for the 3-4 gap. If no slipping takes place then f~z,- f2z:and consequently
x3 - x,. It is this very fact which ~ without slip ~ allows one to cast the result
for rl = rl(~ ) (for rl, = 1"1a (~t,,, K )) in the form of equation 7 (or 10) with the
system constant C given by

(K 2 -1) (15)
C~
47tK2h(R~ + R32)

With slip, however, (f2~,+ t~:) will, in general, not equal (~"~S3"~~"~$4) SO that the
relation between x3and x~is not known (Using the fact that f2~ cannot decrease
with increasing shear stress and that u should increase with x requires
128

(f2s3 + f2s,)<__(f2s~ + f2s,) and consequemly x~ >_x,. This implies that slip effects
in the 1-2 gap are larger than in the 3-4 gap. Since K =/~ //~ - R3 / R, the 1-2
gap is the smaller one so that this result was to be expected.)
This being the case there is no possibility to list a formula for the shear
viscosity. Using a double gap device in viscometric measurements in cases where
slip might occur is an extremely poor choice. Formally using equation (10) (with
C given by equation (15), as was used to obtain the corresponding results
displayed in figure 4) and f'mding odd results could possibly be traced back to
this feature.
Since f2eaccording to equation (4) stays constant at constant shear stress
x, (equation (4)) the sudden decrease of rl~ at x would require a sudden increase
of (f2" + f2o ) . In other words on reaching x any increase in f2 would leave f2I
unaffected but merely serve to increase the slip. Physically such a possibility
cannot be discarded.

3. SLIT FLOW

To possibly shed light on some of the puzzles presented we arm to slit flow
(dimensions of the channel: length l m, height: 3 mm, width 30 mm). Previous
investigations, employing a modification of the customary Laser Doppler
velocimetry (termed GRLDA-method [14]) such that local shear rates can
directly be determined, revealed that the jump of the shear viscosity, determined
via the relation

11 - - ( A p I L)y I (dul d y ) - (Ap I LI)Yl/~ (16)

where Ap / L is the applied pressure drop per unit length, u is the (axial) velocity
in the x direction and y is the traverse direction measured from the midplane, is
accompanied by a sudden jump of the local axial velocity fluctuations, u'. Prior
to the jump the root mean square (RMS)-values are in the 2 % range, while they
suddenly jump into the 10% range exactly at that point, where rlsuddenly
increases, too [14]. The fact that RMS-values in the 2% range are found for low
shear rates already is astonishing and most likely a reflection of the dynamic
129

nature of surfactant solutions. The sudden jump to values of about 10% does
come a big surprise. From a fluid mechanical point of view this hints more in the
direction of turbulent flow. Since the Reynolds number was 81 (based on 1"1+)
this would be associated with some type of structural (or early) turbulence.
Proceeding along that fluid mechanical line of thought one would put for the
axial velocity u,
u=~+u' (17)
(with ~ ' - 0 ), where ~ denotes the mean. The average of the equation of motion
(in the average flow direction, x) thus reads (Here we have assumed that the
mean of fluctuating quantities show no systematic variations in the axial
direction, i.e. ~ ~2 _ ~ - f , = 0).

d -p~'v-') (t8)
0=@+ (~v+-'

Here xxy - ~ + X~yis the x-y component of the extra stress tensor and v' denotes
the traverse velocity fluctuation.
Equation 18 implies for the average shear stress ~+ the relation

_ Ap -' + pu' v ' (19)


+xy - (----L--y- + xy )

It was shown elsewhere [14] that if isotropic tm'bulence prevails (v'= u', i.e.
pu'v'=-p~ '2) Newtonian behavior ('t~,=rls(du/dy), i.e. ~'~v= 0) would
quantitatively account for our results. In other words, the observed increase in q,
as determined via equation 16, would be fictitious, since that increase is solely
due to the (sudden) occurrence of the Reynolds shear stress 9u'v'=-p~ '2. The
viscosity remains constant, 1"1-- q~. (According to equation (19)
q satisfies rl~(d-a/ dy)= -(Ap / L)y if ~ < ~ and - -(Ap / L)-O~ '2 if ~ > Yc)
To study the hypothesis of isotropic turbulence, v' ,~-u', Laser Doppler
velocimetry (in the backward scattering mode) was employed, which allowed us
not only to determine the axial velocity u but also (by rotation of the optical
system by 90 ~ the traverse velocity v - v'[ 15]. Figure 5 shows a typical result. It
130

0.16
I"-"-I
C,6TMA-Sal + NaBr
-~0.14 1000 wppm
25 ~
~"JO. 12
1:3 ~ -~- - m -

~o.~o -

"~ 0 0. 8 -

/ cccco RMS o f t h e v-component \


>,0.06 ~ ,( -'--'--'--'--'-RMS of t h e u-com-ponent
r
o 0.04 -

t.,.~
9
> 0.02 -

0.00 ~T P i I I t t w w I ~ = i -I l I i w I ~ i I i I i-w I = 3"

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5


y - p o s i t i o n in flat c h a n n e l [ m m ]

Figure 5. The average velocity ~ as well as the RMS-values for


the fluctuation of the u-component and, respectively,
the v-component

0.030

t C,eTMA-Sal + NaBr
0.025 : I000 wppm
- 25 ~

/ =~
U~
rn 0.020

0.015
-
. ,,~,,,,,~ RMS~
o o o o o RMS v - c o m p o n e n t

0 . 0 1 0 -2_ ~

_ o o
0.005 0---.- 0
o o o

0.000
o., o.~ o.~ o.~ o.~ o~. 1.b 1.'1 1~. 1~. 1~. 1.~ 1.o
y - p o s i t i o n [mini

Figure 6. The RMS-values for, respectively, u' and v', close to the wall
131

follows from these results that Iv'l << lu'l (even close to the wall, figure 6)i.e. the
Reynolds shear stress, though non zero, is too small to account for the observed
increase. If the flow is indeed ,,turbulent", then it has to be a rather anisotropic
type of turbulence. Interesting in this connection is the fact that the surfactant
solutions investigated here do indeed show an essentially one-dimensional type of
turbulence under truly turbulent conditions (high Reynolds number) e.g. [16].
Since isotropic turbulence would, in our case, imply Newtonian behavior and
vice versa, we conclude that the fluids rheology is definitely non-Newtonian.
Thus, the observed increase in ,, q" as given by equation 16,should mainly be
due to a sudden increase in average shear stress fluctuations. Formally assuming
for the shear stress an arbitrary functional dependence f to the shear rate
~, ~ - f ( ~ ) , we, get, by Taylor expanding the right hand side around ~ and then
averaging

In our case up to the onset the viscosity is essentially constant ( f ( ~ ) = rl,~ ), so


that one anticipates T' = 0 (as is true for a Newtonian fluid). However, if one
realizes t h a t - at least d i r e c t l y - after onset the flow is not (strictly) viscometric
then the notion z - f ( ~ ) is incorrect [17] and

x =rl,~ +x' with -V ~ 0 (21)

is entirely possible. Without a priori knowledge about the fluids rheology, we


know nothing about ~'. For completeness we note that x'will not only depend
upon the shear rate, but on other variables as well [17].
Another possibility exists which might also be responsible for the observed
increase of ~-Iyl/~ at onset. This has to do with the average itself. For an average
to describe the (average) situation properly it is necessary that the mean peaks
around the instantaneous distribution (e.g Gaussian distribution). Performing
measurements at a constant distance (0.5 ram) from the wall and changing the
local shear rate ~ by varying the volume flow rate we find that the axial velocity
distribution is indeed properly peaked for low shear rates (below onset) but is
132

essemially plateau like for high shear rates (figure 7). In this latter case the
arithmetic mean used in LDA studies,

1 (22)
-W- , U , N .

0.25 .., I r ; i ;J ! t I ; I I"1~ I I I 1 l i I l{ i i I' I l ~

CTA-SAL + N a B r 1000 ppm


T- 25~
@.2@~ i

O Shear Rate 0.5 1Is


Shear Rate 5.2 1Is
O)
0
~0.15 Shear Rate 236 1Is

"0
1:::;

<8.1@

O. 05

@. @@ , 9 f i ~o= , I ~ ~ I = =~ =~=~-ri =r :- r ~--" , -" ~-g-:'4-I--/ |

@.1 " 1@
4
@@
Ve]ocity [era/s]

Figure 7. The probability density function for various shear rates

in which N, is the number of measurements, for which the velocity u, has been
measured ( N = ~ N, ) seems highly questionable. Considering the logarithmic
i

scale of the abscissa of figure 7 it becomes clear that equation (17) is not
representative of the average behavior for high shear rates.
These results point more in the direction of some kind of multiphase flow, for
which it is (almost) equally likely to find velocity values well below fi (according
to equation (17)) as well as above. This hints in the direction of a shear induced
state, (SIS), in which monomers, micelles and, for the most part, larger entities
133

(SIS) are in thermodynamic equilibrium. Thus, if we assume that there is, after
some critical shear rate, ~,, an additional SIS state, and that this SIS state is far
more dynamic (Since turbulent flow is customarily associated with vortices of
different size and since we have no evidence of such vortices in our case we
refrain from calling the flow turbulent.) then the state of wormlike micelles
existing up to ~c, then the viscometric results make sense. What we have to
abandon totally is that after onset there is viscometric flow. Quite the contrary,
for ~ > ~ c, the flow is definitely non-viscometric. It may be viscometric in the
mean (even this is open to question, see figure 7), but it is fluctuations around the
mean which bring in the strong (or odd) dependence upon the actual geometry in
relations valid only for truely viscometric flow. Since even prior to onset the
flow is not strictly viscometric (~'2/~2 ~ 2% for slit flow), we expect the critical
shear rate 1'c (actually ~ ) to be geometry dependent, too.

! .00

i CTA-SAL + NaBr 1000 p p m


T- 25~
X - 0~
~l- 8j~ S h e a r Rate 236 1/s

>a

oo 0 . 6 ~

~ I ' 0
o~
i:> I '

20 "

O 1O 2O 30 4~ 50 60
Time Is]

Figure 8. Intensity fluctuations for flow birefringence measurements (after


onset), extinction angle X - 0~

The fact that after onset the flow cannot be viscometric (at least not in its
classical terms) is strengthened further by two additional observations. According
134

to the SIS hypothesis the micelles are all lined up in the flow direction after the
critical shear rate ~ h a s been reached [e.g. 11]. Flow birefringence
measurements will thus produce the extinction angle %- 0 ~ . While we fred this
to be true on average there are rather huge instantaneous fluctuations around this
mean (see figure 8). Corresponding measurements of the instantaneous pressure
difference fluctuations also hint at an extremely dynamic behavior (one-
dimensional) of the flow after onset (figure 9). Attempts to study (after onset)
such behavior by considerations restricted to truly laminar viscometric flow seem
futile.

1 .013 ~,

~0.60

CTA-Sal + NaBr 1000 p p m


T - 25 ~
S h e a r R a t e 236 1 / s
~,1~1~
i
[ I i i i I i i r ' ] " l ~ r I i ~ r ~ i I ! I l k ' i l l t: I I I [ ~ IT i I'1 i r I I ! ~ '1 I 1 I I'l I ii '

.oo 3o. o 5o. e


Time [s]

Figure 9. Difference pressure fluctuations for the results presented in figure 8

For completeness we note that slip effects can def'mitely be discarded. In slit flow
we never observed any tendency of slip, irrespective of the volume flow rate
used. Figure 5 is a representative example for this fact.
135

4. SUMMARY AND CONCLUSIONS

Since the tendency for slip has to be discarded the results presented revealed
the non-viscometric flow behavior (at least in its classical sense) under seemingly
viscometric flow conditions. We do believe that these results reflect the dynamic
behavior of micellar surfactant solutions, especially in the so called SIS state.
This state is a rather dynamic one such that classical viscometric studies, though
useful if performed under identical conditions, may be of little use, if an
engineering attempt is made to use the data in any other system or for,
respectively, upscaling.

ACKNOWLEDGEMENT

Partial support of this study by the Deutsche Forschungsgemeinschaft (DFG) is


gratefully acknowledged.

REFERENCES

. S.Gravsholt, 1973, Pro. Intern. Congr. Surface Activity, 6, 807


2. A.J. Hyde, D.W.M. Johnstone, 1975, J Colloid Interf Sci 53, 349
3. D. Ohlendorf, W. lnterthal and H. Hoffmann, 1986, Rheol Acta 25, 468
4. T. Nash~ 1956, Nature 177, 948
5. H. Hofmann, M. L0bl, H. Rehage, I. Wunderlich, 1986, Tenside
Dertergents, 22, 290
. K. Schmitt, P.O. Bnmn and F. Durst, Progr. And Trends in Rheology II,
1988 Steinkopf, 249
. K. Schmitt, F. Durst, P.O. Brunn, Drag reduction in fluid flows, 1989 Ellis
Harwood Ltd., 205
. E. Ruckenstein, P.O. Brunn and H. Holweg, 1988, Langmuir 4, 350
9. J. Vorwerk, P.O. Brunn, 1994, J Non Newtonian Fluid Mech 51,79
10. A.M. Wunderlich, P.O. Brunn, 1989, Colloid Polym Sci, 267, 627
11. H. Hoffmann, G. Platz, H. Rehage, W. Schorr und W. Ulbricht, 1981, Ber
Bunsenges Phys Chem 85, 255
12. H. Rehage, I. Wunderlich und H. Hoffmann, 1986, Progr Colloid &
Polymer Sci 72,51
13. G. Schramm, Einf0hrung in die Rheologie und Rheometrie,
Gebr. HAAKE GmbH
136

14. A.M. Wunderlich, P.O. Brunn and F. Durst, 1989, Rheol Acta 28, 473
15. R. Steger, 1994, Dissertation, Universit~it Edangen-Ntirnberg
16. H.W. Bewerdorff and D Ohlendorf, 1985, Proc 5 Symp Turb Shear
Flows, (ed. J. Lumley), Ithaca, NY, 241
17. G. B0hme, 1981, Str0mungsmechank nicht Newtonscher Fluide, Teubner,
Stuttgart
137

TIME PERIODIC FLOWS


J. D u n w o o d y

Department of Applied Mathematics ~ Theoretical Physics,


The Queen's University, Belfast B T7 1NN, N. Ireland.

1. I N T R O D U C T I O N

Time periodic flows have a prominent position in rheological ex-


periment and fluid characterization, since they are used extensively to
determine the relaxation functions, or their spectra, which figure in the
constitutive description of viscoelastic fluids such as polymer solutions
and polymer melts. Invariably the amplitude of the oscillations in the
viscometers designed for this purpose are small, or infinitesimal, so that
linearised constitutive theories have been deemed adequate to describe
the flow(cf. Ferry[I]). In addition it is standard practice to neglect the
inertia of the motion generated by some external stimulus, particularly
in parallel plate and cone and plate instruments. Such simplifications, if
made a matter of course, are a source of error in the analysis of exper-
imental data as stressed by Schrag[2] in the case of parallel shear flow.
This is even more likely for parallel cylinder viscometers, such as those
used by Oldroyd et. al.[3] and Markowitz[4], because of the nature of the
fluids tested in these instruments.
Of course, viscoelastic fluids and non-Newtonian fluids in general al-
ways require a nonlinear constitutive law to adequately describe their var-
ied behaviour, particularly when subject to dynamic forces which produce
strains of considerable magnitude. There is some doubt as to whether or
not the standard regime of experiments used to obtain data on particular
fluids is broad enough to yield a material description which is complete
and suitable for all circumstances(cf. Vrentas et. al.[5]). For this reason,
but also because the flows have inherent scientific interest, the review
below is considered opportune.
138

In what follows attention is restricted to those time periodic flows


which are defined to be viscometric according to the definition given by
Truesdell & Noll[6], which is nonstandard and includes time dependent
flows. Thus, elongational flows are excluded. Time periodic perturba-
tions of steady viscometric flows are included, but only if both the steady
flow and the perturbation satisfy a common, specific constraint on the
velocity components which delineate a particular class of curvilineal flows.
While Schrag[2] has emphasised the role of inertia, too often it is
understated, and indeed it is frequently overlooked that some of the
standard viscometric flows are dependent on neglible inertia for their
existence. Also, the effects of nonlinearity must be a consideration when
the amplitude of the oscillations is not small; but not so large as to make
the existence of the flow questionable(cf. Hatzikiriakos & Dealy[7]). If
the physical circumstances are right, both nonlinearity and inertia will
have contributory and interacting effects on the possible motion.
Throughout this article it is assumed that a fluid adheres to solid
boundaries, that is to say that no-slip boundary conditions are applied.

2. V I S C O M E T R I C FLOW

The geometry of the most commonly used viscometers (cf. Ferry [1])
is such that the flow may be considered to be one of the curvilineal flows
described by Truesdell & Noll[6], in which the particle paths are con-
veniently expressed in terms of orthogonal curvilinear co-ordinates. For
the viscometer flows these are invariably rectangular Cartesian, cylindri-
cal polar or spherical polar co-ordinates. In general these flows may be
time dependent, and in particular time periodic in the sense that the flow
characteristics at a given point in the flow repeat after a given period of
time.
In the appropriate set of orthogonal curvilinear co-ordinates the ve-
locity of a fluid particle at x and time t has the contravariant components
with respect to base vectors ei(x):
v - o , v - v t) , - v 1, t ) , (1)

where x ~ - x.ei(x) and ei(x) are the duals of ei(x). Furthermore, it is


assumed that the dependence on the coordinate x I is such that the ratio
Or
) (2)
139

is independent of time, or equivalently that

Ov 2 Ov 3
Ox ~ = f (x * )q(x 1 , t) , OX 1 -- h ( x 1 ) q ( x 1 , t ) (3)

As seen below, this is necessary for the physical components of the shear
stress on the surfaces x I constant to be in time independent ratio, so
that while fluctuating in magnitude they maintain their directions at
each point on the surfaces x 1 constant.
For these motions the displacement vector u(x, r; t) of a particle at
past time r relative to its position x at present time t has the components

,,1_0, u2 _ v 2 (x 1 , a ) d a , u a-
S v2(x 1,a)da ,-oc < r <_ t . (4)

The relative deformation gradient is then

Ft(x,r) - 6~ + -~Txjda ek({) | &j(x), (5)

where
{ -- x + u(x, r; t) (6)
and the base vectors

(r)

are orthonormal. For these orthonormal vectors there exists a time de-
pendent orthogonal tensor Qt(x, r) such that

a~(x) - q,(x, ~).a({), (8)


and so

,/
q~(x, ~).F,(x, ~) - v 77)(~) ~ (9)

Vg-~JJ)({) ~j + k(~; t)
/O00/
Equivalently, the tensor defined by (9) has the matrix of components

g(kk) (x) k
~ o o (10)
/300
140

where

k(s; t) -- --~/ f : q(x 1, t -- r ) d r , - I g 1 1 1 ( g 2 2 ( f ) 2 + g33(h)2), /


(11)
O~ - -

,/ 11
,.)/--1 lg22f, - - 7--1 v/ g111g 33 h ,

and 0 < s - t - 7 < oc. The incompressibility condition requires that


the Jacobians satisfy the condition

g(x) - H g(,,)(x) - I I - (12)


i i

This condition is satisfied if


(i) the coordinates are rectangular Cartesian or circular cylindrical po-
lars
(ii) the coordinates are spherical polars and the velocity component
v 2 - 0.

Indeed in both these cases

g(ii)(x) - g(ii)(~) (13)

for all 7-, and the ratio c~ 9 fl is then independent of time, being dependent
on x only as is clear from (ii). At each point x at present time t a basis
i~:

i~ - e l ( x ) , i2 - c~62(x) + fle3(x) , i3 - -f162(x) + o / e 3 ( x ) (14)

may be defined such that the relative right Cauchy-Green strain tensor is

Ct(x, t - s) - FT(x, t - s).Ft(x, t - s)


= I + k(s; t ) ( N T + N) + k2(s; t ) N T . N ,

where N has the component matrix with respect to the basis ik"

O 0 0/
1 0 0 (16)
0 0 0
141

For the class of incompressible simple fluids with constitutive law ex-
pressing the stress tensor T as a non-linear tensor valued function of the
strain tensor history:

T + pI - S,% 0 ( C t ( t - s) - I) - S ~,=0 (k(s; t ) ) , (17)

it follows from the principle of material frame indifference:

Q . T . Q T - - p I + S~%o ( q . c t ( t - s ) . q T - I) (18)

for arbitrary orthogonal Q (cf. Truesdell & Noll [6]), that

T - (i3.T.i3) - T ~ o (k(s; t ) ) ( N + N T) + Sls%o (k(8; t)) N T . N


(19)
+ ,$2~=o (k(s; t ) ) N . N T

From this relation and (14) the physical components

T<ij> -- ei(x).T.ej(x) (20)


of the stress tensor may then be readily obtained. In particular

T<12> - o/~C~=o (]g(8; t)) , T<13> - fl~c~_-0 (k(8; t)) (21)

with similar relations obtainable for T~23> , T < l l > - T<33> and 7'<22> -
T<33>. Hence, in these time dependent curvilineal flows the physical
components of the rate of strain tensor

1 [Vv + (Vv) T] (22)


and stress tensor T have the property

D<12> " D<13> -- T<12> " T ( 1 3 > - a - f l (23)

2.1 T i m e P e r i o d i c Flow
By a suitable choice of Q in (18) it is readily established that the
shear stress functional in (19) and (21) is odd in the history k(.; t), i.e.

:r, o t)) - t)) (24)


142

Also, if the non-zero components of the velocity are time periodic, i.e.

271 i
v i ( x 1, t ~- - - 1-- V (X 1 t) for i -- (2, 3), (25)
a2
then
q ( x 1, t "F- -271"
) -- q ( x 1 , t ) (26)
a2
and
k ( s ; t + 2a- ) - -7 ~o ~ q ( x 1, t -~- -27r
- - r)dr
~d CO
(27)
-- --'7
/o q ( x 1 , t -- r ) d r - k ( s ; t)

Hence, time periodic curvilineal flows have the property that the shear
stress components defined in (21) are periodic:

271
T<li> (x 1, t -}- - - ) -- T<li> (x 1, t) for i -- (2, 3). (28)
a)

In passing it is noted that SlsC~__0 (k(8; t)) and 8 2 ~ 0 (k(s; t)) are even func-
tionals of their arguments, so that all the stress components are periodic.
If the velocity is anti-periodic,

v i ( x 1 t + -7r) - -v i(xl t) (29)


~d

and therefore periodic, then the history

k(8; t -~- - - ) -- - - 7
~d
/o s q(x 1, t + ~d
r)dr
(30)
-- "7
/o q ( x 1, t -- r ) d r - - k ( s ; t)

for all 0 _ s < ~ . It follows from (21) that if (29) applies then
71"
T<lk> (X 1, t -~ - - ) -- - - T < l k > (X 1, t) fo r k - (2,3), (31)
a2
i.e. the shear stresses are anti-periodic. However the normal stress dif-
ferences are not(cf. Coleman & Noll[8]).
143

It is evident that the converse of the statements that (25) implies


(28) and (29) implies (31) is true if and only if

ceT<12> -~-/~T<13> - ~~ 0 (k(8; t)) (32)


is invertible. While this is to be expected in the neighbourhood of the
zero history k(s; t) for all 0 ___ s < ~ , it should not be taken for granted
elsewhere (el. Hunter & Slemrod [9] and Truesdell & Noll, w
Thus the generation of time periodic flows of arbitrarily large amplitude
by application of time periodic forces must be treated with caution. The
observations of Giacomin, Hatzikiriakos et al. in a series of papers in-
eluding references [7,10,11 ] and also those of Durand et al. [12 ] and
their introduction of 'wall slip' are also pertinent to this point.
Because of the above observations, it is considered prudent to con-
sider firstly those flows as above which give rise to strain histories such
that Ik(s; t)l is moderate for all 0 _ s < co. To this end it is assumed
that k(s; t) C E where

)U--
{ k(s;t)"
/0 e-P~k(s;t)ds < o o , ~ ( p ) > 0
} , (33)

and
k(t; p) -- e-P~k(s; t)ds < e ~ , ~(p) > 0 (34)
j~0~176
is an element of a Fr(~chet space of functions analytic in a right half-plane
of the complex plane. Further, it is assumed that E is a subset of the
domain of the functionals in (19). All k(s; t) such that

[k(t;P)l<M<~176 ,~(p)>0 (35)

belong to/C, and in particular all periodic flows with transform


27r

k(t;- p) - 1 -Pfe- :~____e~fo -z- e-P~q(t - s)ds (36)

satisfy this requirement for all p such that ~(p) > 0 through proper choice
1
of v q ( t - s) -- 2 (t rD 2) ~. In addition
27r - 27r -
q(t + - - - s) - q ( t - s) 4, k(t + ;p) - k(t; p) (37)
cO CO
144

Among the materials covered are the Coleman-Noll[8] fading memory


fluids requiring that for some real/3 > 0
1
> fO c~ e -;3~ { k2(s;t) + -~
lk4(s;t)} ~ ds > fo c~ e-Z~[k(s;t)lds, (38)

the K-BKZ class for which

7-(k(~; t)) - ~- (k(~; t), ~)d~ (39)


j~0(X)
exists on (33), and various special models such as those named for
Maxwell, which form sub-species of the above (cf. Crochet, Davies &
Walters [13]).

2.2 Equations of M o t i o n
It is assumed that each of the stress functionals appearing in (17)
and (19) is Fr~chet differentiable over its domain in some normed vector
space of functions on [0, ~ ) , which intersects with (33). For example,
corresponding to T~0(. ) there exists a linear functional dT~=o(.) defined
by

%~o (k(~; t) + ,~xk(~; t)) - LTo (k(~; t))


d T ~ o (k(s; t)). Ak(s; t) - lim
e----~0 6
(40)
for all Ak(s; t). Formally, the relation (40) may be replaced by

dT (/c(t; p)). Ak(t; p) - lim 7 (k(t; p) + eAk(t; p)) - r (k(t; p)) (41)
e----~0 6
where
T (k(t; p)) -- ~'s~ (/~-1 [k(t; p)]) , }
(42)
dT (~(t; p)) -- d~/-s~ (/~-1 [k(t; p)])./~-1
are compositions of operators.
Also, if
Vi (xl , t; p) _ --[c~ e -psv i(x I t -- s)ds (43)
Jo
145

it follows that
O~ i
Vi = _~_pOi i -- (2 3) (44)
Ot ' '
and
0 ~2 0~3
~1~ Ox 1 = -apk(t; p) g_33_
v lOxl
= -flpk(t;p). (45)

In general the extra stress tensor

S-T+pI (46)

has the physical components

I g( ii) i (47)
S<ij> - g(JJ) Sj - v/g(ii)g(jj)S ij

and in terms of them the equations of momentum balance are

OvJ Ovj v/g(~) Ov/g(~) v~v~ + 1 Ov/g(JJ)ox


s vJ Vs ]
flv/g(jj) - - ~ -t- v s Ox-------
~ -- g(JJ) OxJ v/g(jj)

=~ v/g(j)) 0 S<iJ>~/g ) -t- 10v/g(jj) ~<jj>


i ~ Oxi v/g(ii)g(jj) g(jj) OxJ
1 { Ov/g(JJ) Ov/g(~) } 1 op
2 ~SKij> -- SKii> --
+ ~ir V/'g(ii)g(jj) OX i OxJ v/g(jj) OxJ
(48)
Whether or not the preceding purely kinematic considerations are empty
is determined by whether or not the equations (48) for specific choice
of coordinates admit appropriate solutions for the velocity components
(3) with rate of shear strain histories (11). It is evidently true that the
form of the constitutive equation (17) must have a bearing on this, but
the question of existence is most often avoided by the convenience of
equating the inertial terms making up the left hand side of (48) to zero
on the grounds that they are of no consequence. Solutions to the resulting
divergence equation, involving the stress tensor only, may then be sought
by the inverse, semi-inverse methods well known to elasticians. However,
the errors that can arise from this simplification are all too commonly
overlooked(cf. Schrag[2]).
146

2.2.1. Plane Flows.


For these flows the chosen coordinate system is rectangular Cartesian
with the cordinates of a fluid particle (x 1 , x 2 , x 3) identified with (x, y, z),
while
(~-1, ~-0,
V3 -- 0 ==~ k(S; t) -- -- ~0 ~ 0V2(X, t -- r) d r (49)
5x
so that (48) reduce to

0S<11> OP = o
~x Ox
0S<~2> @ Ov 2
(~0)
Ox N P-gi-
Op o
Oz
It follows from (50) that if a pressure

p - a(t)y + b(t)+ S<11> (51)

is applied the solution of equations (50) reduces to the solution of


~v 2
aS<12> _ a(t) -- p (52)
Ox Ot
subject to
ak(~; t) Ova(x, t - ~)
(53)
Os Ox
If it is assumed that the inertial term in the right hand side of (52) is
neglible and then set to zero, solutions to (52) and (53) may be sought
in the separable form

v~ - ~ {~y(x)}, ~- r (54)

which implies that

S<12> ----~/-s~ ( ~ If( x) --~(1 - e-'~) ]) f(x)- V'(x),. (55)

Then there are solutions if and only if


147

(i) a(t) = 0 = f ' ( x ) = 0 while dT~=o exists,


~ - a (t) # 0.
(ii) T~--0 is linear, f'(x) is constant and a(t + -5-)
The first of these conditions relate to oscillatory plane shear flow and
are appropriate to gap loading(cf. Ferry[I]). However, in order that (52)
yield a unique value of S<12> corresponding to each f ( x ) it is necessary
and sufficient, since the frequency w must be absolutely small for neglible
inertia, that

lim ~ ,, (1 - e r O. (56)
w--~0 b(M

If this condition does not hold then the possible effects of hysteresis must
be considered(cf. Hunter &: Slemrod[9]).
The second set of conditions relate to plane pulsatile flow, but the
condition that T ~ 0 is linear is very restrictive. Among the fluid types
satisfying it are the Navier- Stokes fluids and certain 'second order fluids'
(cf. Truesdell & Noll[6]). Of course if the strain amplitudes are infinitesi-
mal a linearized theory, which is compatible with both sets of conditions,
is applicable.
Existence of periodic solutions to (52), for a special model proposed
by Slemrod[14], is assured by a theorem of Rabinowitz [15] (cf. Dun-
woody[16]), if the nonlinearity in the equation is sufficiently weak.

2.2.2. Cylindrical Flows.


For these flows the curvilinear coordinates (x 1, x 2, x 3) are identified
with the cylindrical coordinates(r, 0, z), while in (11)

- ~(~) , ~- ~(~) , ~- Z(~) , (57)

so that the physical components of the extra stress tensor S are from (17)
s<~j> - ~<~j>(~). (58)
Hence, the equations (48) become for these flows
1 0 S<22> Op = -pro 2
rot (rS<ll>)- r Or
1 0 (~'S<12>)-~- S<12> 1 0p = - p r o
oo

(59)
r Or r r O0
10 Op oo

rot (rS<13>) Oz = - p z
148

where t~ - o0 etc. It follows from these that the pressure required to


support such flows is of the form

P--~<11>-t-f ~<11>--~<22>r - ~ - p ( r b ) dr + a(t)O + b(t)z. (60)

The velocity components (t), i) are obtained as the solutions, if such exist,
to the equations

OS<12> 2 pro.9 + a(t)


Or + r S<12> -- r '
(61)
0S<13> 1
Or + -S<13>r - p2 + b(t) ,

where

S<12> = c t ( r ) T ~ o (k(s; t)) , S<13) -- /~(r)~s~_-o (k(8; t)) , (62)

subject to

O 0 ( t - s) Ok(s; t) O i ( t - s) -
ok( ;t)
--r OF ~ oz.(r)~ ~ --
Os Or (63)
1 - c~(r) 2 +/3(r) 2

and a ( r ) , / 3 ( r ) , a(t) and b(t) are arbitrary. In most practical cases

a(t) -- 0 (64)

due to the physical requirement that the pressure be single valued in 0.


If the inertial terms pro and p~ are neglible and equated to zero, and
the choice
b(t) - 0 (65)
is made, then equations (61) are satisfied by

, (66)
r
149

where C1 and C2 are real constants.This corresponds to the choice

OZ(I')" / ~ ( f ) - C1" rC2, (67)

in (23), and so from (62) and (63) it follows that

t)) - -;r- + r2C~. (68)

Given that the modulus of the right hand side of this relation, for all r
in a finite interval, is within the range of the functional T,~=0(-) over the
domain/C, then if the functional is one-to-one and invertible over/C the
shear strain history k(s; t) is periodic by (27) and uniquely determined by
(68). The velocity components are then determined by (63)1,2. However,
only in the case that the functional is linear would the shear strain history
be simple harmonic.
Couette flow corresponds to the degenerate case

/ 3 - 0 e=> (72 - O, (69)

and sliding cylinder flow to

Og -- 0 ~ C1 -- 0. (70)

Even for these cases there is no known physical device which could enforce
stresses of the form (66) at cylindrical boundaries.
Another interesting type of flow falling within this general class is
pulsatile (axial) flow in a tube or annular pipe. However, for it

b(t) 7k 0 (71)

in (61), so that the stresses (66) no longer satisfy the equations of mo-
tion. Most often practical interest has been centred on small oscillations
superimposed on steady flow and inertial effects are not always neglected
(cf. Barnes, Townsend & Walters [17]). Mena, Serrania & Van Ziegler
[18] have performed experiments in which perturbation flows of all the
above types have been superimposed on steady axial flow.
150

2.2.3 Torsional Flows.


The type of flow envisaged is that which may be caused by inde-
pendently rotating about the common perpindicular axis through their
centres one or both of two parallel circular plates containing a fluid be-
tween them. Again the chosen coordinates are circular cylindrical except
that in this case the curvilinear coordinates (x 1, x 2, x 3) are identified with
(z, 0, r). Also

c~-1, /3-0, O-w(z,t),


V3--#--O=~ (72)
k(s; t) - - r fo ~ O~(z,Ozt - ~) du .

so that
S<12> -- "~s~ (~:(8; t)) -- S < 1 2 > ( r , Z, t) (73)
while from (14) and (19)

S<13> -- S<23> -- O. (74)

The equations of motion then become

0S<11> OF = 0
Oz Oz
0S<12> 10p = pr iiv (75)
Oz r O0
0S<33> S<33> -- S<22> 019 02
Or + r Or = - p r

subject to
0k(,; t) O~(z, t - ~)
~r (76)
Ot Oz
From the first and second of these

P -- S<11> -I- g(r, t)O, (77)

and for a single valued pressure

g(r,t) --0. (78)


151

Therefore,
S<12> - fir f (:u(z,t)dz + h(r, t) (79)

and it is clear that (73) and (76) are incompatible with (79) in general,
unless some simplifying assumptions are made.
If the inertial terms in the right side of (75) are neglected then (73),
00
(76) and (79) are compatible if in (76) ~z is independent of z. In partic-
ular this is true if the velocity ~) has the time periodic, separable, linear
form
O - N(az + b)e *"~ (80)
which corresponds to the physical condition of 'gap loading' again. The
analysis for the steady flow case and references to experimental work is
to be found in Truesdell & Noll[6].
The restriction that time periodic flows of this type may only occur
under gap loading conditions also applies to the following type of flows.

2.2.4 Cone and Plate Flows.


The chosen coordinate system is spherical polars, and for convenience
(x l, x 2, x 3) are identified with (0, r r ) i n this system. Again the velocity
has only the one component which is

and

v3 - - / ~ .-j0~
k ( s ; t ) - -sinO
0 (0, t -
~
u)
du.
(83)

Hence, as for the previous type of flows,

S<13> -- S<23 = 0, (84)

while
S<12> - g < 1 2 > ( 0 , t ) . (85)
The flows envisaged could be generated by the independent rotations
of a circular plate and half-cone about the axis of the cone which is
perpindicular to the plate surface, the fluid being contained between.
152

The equations of motion for these flows are from (48)

1 0S<11> cot 0 1 cop _ p r sin 20 ~2


r 00 +r (S<11>-S<22>) r CO0 -2 '
1 0S<12> cot 0 1 0/9 sin 0 r
r O0 ~- 2 r S<12> r sin 0 0r --pr

0S<33> 1 {S<II> -~- S < 2 2 > -- 2 S < 3 3 > } Op --pr sin 2 0 ~2.
Or r Or
(86)
It is clear from (85) and (86)2 that if

0p
= o, (87)
or
which is the condition necessary for the pressure to be single valued in r
there is incompatibility between (83) and (86)2 unless the inertial term
is zero, which is the case for steady flow. In addition,the relations (19)
and (83) imply that

0 (S<11>- S<33>)-
o-7 ~0 ( s < ~ > - s<~>) - 0 (88)

so that there is compatibility between (83) and (86)1,3 if and only if

0
(~----~{Sis%0 (]9(8" t)) -~- $2c~ 0 (]9(8" t))} - 2 p r 2 sin 2 0r O~ (89)
' ~= ' 00'
which cannot be satisfied for all r and 0 since

O0 r O. (90)

This incompatibility has not always been recognised(cf. MacDonald,


Marsh & Ashare [19]).
On equating all the inertial terms to zero in the right hand sides of
the equations (86) they and (83) are compatible if and only if

c(t)
~<12> = - 2
sln 0
(91)
0 {SiCs=O
00
c~ (k(8"
~ (k(8", t))-~- $2 s=O .,
t))} - O.
153

These conditions are satisfied for all fluids, but only approximately, when

2 c___0_~ 7~, 0r = - a (t) ,


(90 (92)

T~=o (fo a ( t - u)du) - c(t),

where e is small. In particular, the time periodic form

- (aO + b)e ~ t (93)

is possible under these stringent conditions, which again corresponds to


a form of of 'gap loading'. The stresses 7'<12>, T < l l > and T<33> at the
boundaries required to maintain the flow are then readily obtained from
(86) 1,3 once r is prescribed.
As noted by Ericksen[20], there are two distinct approximations in-
volved in the above analysis, one dynamical in nature and the other
geometrical.

3. S M A L L A M P L I T U D E FLOWS

In these flows a time periodic viscometric flow with small amplitude


of displacement is assumed superimposed either on a rest state of the
fluid, or a steady viscometric flow of it of the same type. The contribution
that the neglect of inertia might make to error of measurement is the basic
consideration.

3.1 Gap Loading and Inertia


Since inertia is not being ignored, the discussion must be restricted
to either plane or cylindrical flows, and to avoid the unnecessary com-
plications arising from the geometry of the flow the former of these two
types is chosen. A comprehensive treament of 'simple shear' has been
given by Schrag [2], but a revised version influenced by the desire to ex-
tend it to the nonlinear theory of viscoelastic materials has been given
by Dunwoody [21] and is preferred.
It is assumed that in a time time periodic plane shear flow

m~x Ik(~;t)l << 1 V t. (94)


0<s<c~
154

Therefore, if in particular (55) applies


~u 2
f(x) << 1 ~ I-~ I/~ << 1, (95)

so that while the relative strain is required to be small there is no such


restriction on the strain rate, except at low frequencies; but, this obser-
vation is general and a derivative of (11). Under these circumstances, an
e ~ 1 may be chosen such that

k(s; t) - Ek(s; t) (96)

and it then follows from (41) that

7~0 (~k(~; t)) - ~ 0 ( 0 ) . ~(~;t)+ o(~). (97)

In the case of oscillatory plane shear flow an appropriate value for E is the
A
ratio ~ , where A is the amplitude of the oscillations of one of a pair of
parallel plates which are a distance L apart, the other being fixed. There
is then no motion when A - 0 (cf. Dunwoody [21]).
Also, for a periodic flow of the form (54) with f ( x ) - V' (x) constant,
a complex valued dynamic viscosity may be defined by

~ * ( ~ M ) - d~]"s~ 9 [/,03-1 ( 1 - e-*"~*)] , (98)

which has the assumed limiting values

#- lira q*(w) (99)


~,~ "--+ 0

the Newtonian viscosity corresponding to steady shearing for which


k(s; t) - - e s , and
E- lim ~wq*(w), (100)
o.J ----~ o o

the instantaneous elasticity, or rigidity corresponding to an impulsive


uniform shear for which k(s; t) - - e H ( s ) . For example, if the linear
operator in (97) has the representation

d%~o(O), k(~; t)- fo ~ G(s)k(s; t)ds (101)


155

and G(s) the relaxation spectral decomposition

ak
-~/ak (102)
k=l

then
n n

k=l k=l
To facilitate estimation of the importance of the inertial forces rela-
tive to the viscoelastic forces in these motions it is convenient to introduce
a set of nondimensional variables as follows:

LSc - x, ewLv - v2 co--li- t ~COIT]*(CO)lCr -- S<12> (104)

In terms of these new variables a nondimensional form of the equation of


motion (52) with a(t) - 0 is obtained which is

Oa = R* O--v-v R* = PwL2
o~ oi' I~*(~)1 ' (105)
(7 -- --(O2]T]*(a2)l) -1 d T 2 o ( O ) 9 f0 ~ 0~(~,-0~i - r d+.

Corresponding to the above limits on the dynamic viscosity the nondi-


mensional parameter R* has the limits

lim R* = pwL2 , lim R* = pw2L2 , (106)


.o--+0 # .J--+~ E

which are the equivalents of a Reynolds Number and Mach Number re-
spectively.
A separable solution to (105) of the form

v(~,b- ~ [v(~)~l, } (107)


V(O)-O, V(1)-l,
is readily obtained by solving the ordinary differential equation in V(2)"

'*-(~) v"
I~*(~)1 ( ~ ) - m*v(~) - o, (lO8)
156
since from (98)

d ~ o ( O ) 9f0 De - * ~ ' d ~ - coU*(co) . (109)

The appropriate solution is

sin/3:~ /32 _ - ~ n * l n * ( ~ ) l (110)


V(:~) - sin/3 ' r/*(aJ) '

and it is uniquely determined because q*(co) is by assumption strictly


complex, so that cos/32 and sin/~2 have no zeros for any values of 2 7~ 0.
For neglible inertial effects, equivalently 'gap loading',

R* < 1 ,:~ IA~l << 1 (111)

and then
V(2)- 2 (112)
approximately. The shear stress arising from the motion (110) is

- ~
[~*(~)cos~]
i-~*-~-)lsin3 ' (113)

and, if (111) holds, it is approximately

q*(co) ] (114)
(l -- O'g - - ~1~ Iv*(~)l "

The error involved in using (114) in place of (113), regardless of the


constraint (111), is indicated by the ratio of their amplitudes and the
difference in their phase angles which are evaluated from

I~--2 ~-~ I ~d arg ~gg -- arg s--~n~ (115)

respectively.
If
arg [r/*(co)] -- - r (116)
157

then it follows from (110) that

3 - - OZl -- LOZ2-- R*2"


1{ cos 2 4 -4-~sin 2 4 " (117)

Important limiting values of the ratio ~_x


C~l
are obtained from (99) and
(100), viz.
Od2
(i) 1,
OZl
7: a2 (118)
(ii) w --+ e c ~ r ~ -~ , , -- - O,
OZl
which correspond to N e w t o n i a n f l u i d response and e l a s t i c s o l i d response
respectively as expected of these materials. Dunwoody[21] has demon-
strated through a plot of the ratio ~-~ against frequency ca using a specific
Ctl
K-BKZ model designed by Luo & Tanner[22] to fit data on LDPE melts,
how these asymptotic values are connected by a monotone decreasing
function of w. In addition, using the same model, he has demonstrated
the monotone increasing behaviour of ,7.(~o) as a function of ca, which
implies that R* for the specific fluid increases with w for fixed gap width
L and density p.
For the computation of relaxation moduli it is essential to have com-
puted values of r/* (w) over a wide range of frequencies 0 < w _< wM < oc
such that
<< 1 V w > WM (119)
max It/* (w) ]
0 <(-O~WM

(cf. K a m a t h & Mackley[23]). This condition is in clear conflict with


the condition (111) for gap loading, unless the gap width L is severely
reduced at the higher frequencies.
Following the example of Schrag[2], the arguments of the function in
the right sides of (115) may be written as

/~COS/~20Zl(1--~a2)COS{Ctl(1--~c~---~-2)
2 } a l al
(120)
sin/3 sin (OZl (1 - ~2-~-)}

Then, assuming 0 < r < 2, values of the ratios in the left hand sides of
(115) are plotted in Figure 1 and Figure 2 as functions of 2 for various
values of ~_x
O~1
in the range [0 ~ 1] and al > 0 9
158

0~1 -- 0 . 2 5 O~ 1 - - 0.5
1 . 0 4 ~
1.00 1.02
. . . . . . . . . ' X
~ . . .~ . . . . . . . 1i x
0 99 " " "~\"K~ " 0.98 0.2 0.4 0"~'N.~K"-5 0.75
0.9~ ~ 0.75 0.96
0.98 0.50 0.94 0.5
0.98 0.25 0.92 0.25

a 1 -- 0.75 a 1 -- 1.0

1.05 0.75
/
. . . . . . . X ' ' " " " " ~ X

0.95 0.2 0.4 0 0.75 0.2 0.4 O.


. . ~ I 0.5
0.9
0.9 05 0.8
0.85 0.25
0 25 o.7

Figure 1. Variation of [a___[ across the gap width for specific


Crg

values of a l and ~a2 _ 0.25, 0.5, 0.75 in each plot.


cr

al - - 0.25 OL1 ~ 0.5


0.03 0.75 0.75
0 5 0.I 0 5
0.02
0.01 0 25 0.05 0 25

-0.01 ~ . . . . . . . .

.6 0.8 I
X

- 0 . 0 5
~ . . . . . .

.6 0.8 I
X

OL1 -- 0 . 7 5 al-- 1.0


0.3 0.75 0.75
0.5
0.2 0.5 0.4
0 25
0.i 0.25 0.2

~
. . . . . . . . . . .
. . . . . . . X

_ 0 . 1 ~ . 6 0.8 1 .6 0.8 i
-0.2

Figure 2. Variation of arg (~-~_)across the gap width for the


\"91

same values of a l and a___22_ 0.25, 0.5, 0.75 again in each plot.
C~l
159

It is remarkable that their variations are so significant at relatively small

values of a l - R*89 + ~~_a and values of ~


~__xin the middle of the
range [0, 1].
The above observations cannot be overemphasised.

3.2 A s y m p t o t i c a p p r o x i m a t i o n s
In the case of small strain oscillating cylinder flows the constraint
(94) is satisfied and the relations (96) and (97) again hold. Thus, the
equations of motion (59), with/3 - 0 , reduce with errors O(e 2) to the
single equation
1 0 S<12> "
/" C~" ( / ' S < 1 2 > ) - t " -- - p r o (121)
r

and
P - po + . / p r O 2 d r (122)

is single valued in 0. The strain history is from (63)

k ( s ; t ) - - / r O 0 (Os*
t - S*)ds, _ ek(s; t). (123)

The flows envisaged are contained by two cylinders of radii a and b, a < b,
so that the gap width is
L - ( b - a). (124)
Therefore, there are two fundamental lengths which must figure in the
considerations of dimensional analysis, viz. the the inner radius a and the
gap width L. It is evident from (41), (97) and (123) that, as in the case
of plane flows, a separable solution to (121) may be sought which is time
periodic. Since equation (121) is explicit in r it is appropriate to choose
a as the more fundamental of these two lengths, and so corrresponding
to (104) the non- dimensional set of variables

a~ -- r, ea~av -- tO, a~-lt- t, elU*(a~)la -- S<12> (125)

is introduced for these flows. In terms of them the equation (121) is


replaced, on setting
v - ~ [V(~)e ~t] , (126)
160

by the ordinary differential equation

,~v"(, ~) + ,~v(,~)+ ( Z ~ - 1 ) - 0 (127)

where
pwa 2
~ _ _~R:I,*(~)I and R; = (~)1 (128)
,*(~) I,* "
The boundary conditions which are applied may be

V(1)-0, V(I+L)-I+L, (129)


which correspond to the standard Couette flow. Alternatively, they may
be
V(1 + L ) - (1 + L)~ [~e'~]
_ _ (130) ,

d (~_lv(?~)) -- fl2:~V(1) -- 0
d
where the non-dimensional parameter S defined by the relation
aS= K-w2I
27rPw2a3L (131)

depends on K, the restoring constant of a torsion wire attached to an


inner cylinder of radius a, length L and moment of inertia I. Thus the
conditions (130) are representative of a popular device in which the oscil-
lations of an inner cylinder, created by the forced oscillations of an outer
cylinder through the medium of a fluid contained between the cylinders,
are resisted by the torsion wire (cf. Oldroyd et. al. [3], Markovitz [4],
Walters [24]). The phase constant c in (130) allows for a difference in
the phase angle of the oscillations of the two cylinders, and in arriving at
(130) the identity
b- ~0 (132)
has been employed to eliminate 0 in favour of O.
The solution corresponding to the boundary conditions (129) is

(1+L)
V ( ~ ) -- -- 2-W-(L-) { J l ( ~ r ) Y l ( f l ) - Yl(flr)Jl(fl)) ,
(133)
Jl(~) v~(~)
tL/- [ t1+ Lt] [ t1+ Lt]
161

That corresponding to the the boundary conditions (130) is more com-


plex,but the amplitude of the velocity at the inner cylinder and the phase
angle of the outer cylinder relative to it are obtained from

V(1)- ~(1 + L ) e *c [Y2(/3)J1(/3)- J2(/3)Y1( z ) ] / D ,


D - Sl [/3(1+].)] Y1 [/3(1+ L)]
J~(9) v~(~) (134)
+ ~ J1 [/3(1+ L)] Y1 [/3(1+ L)] r
Jl(/~) YI(~)
and the requirement that V(1) is real. The conditions that W and D are
non-zero guarantees that there are no 'normal modes' in either case.
Because the arguments of the Bessel functions in (133) and (134)
are complex, asymptotic expansions have been employed in their evalu-
ation(cf. Oldroyd [25]). In practice this causes no serious error as it is
most likely that the fluids tested in cylindrical viscometers will be 'short
memory' fluids of low viscosity. Interest will therefore be concentrated on
reasonably high frequencies. As noted by Joseph[26], high frequencies in
the argument of rj* (co) correspond to small times in the arguments of the
relaxation modulus G(s) in (101), from which the asymptotic formula

~*(~)~E~ G(0)
~ , G'(0)
(~)~ ~ o(~co2) (1 35)

is readily derived, assuming G(s) is sufficiently smooth at s - 0.


For small enough values of co
(136)
and then the Bessel functions may be replaced by their power series
expansions to obtain a power series in/3 to replace (133) say (cf. Walters
[27]). On the other hand, for intermediate values of w Waiters [27], in
his critique of previous work by Markovitz [4], has shown how to obtain
^

a Taylor series in powers^of ~L for V(1) by expanding its expression in


(134) about its value at L - 0. The condition for strong convergence of
this series is
pooL2
19IZ, << 1 ~ R*L 2 = = n* << 1, (137)
162

which is the equivalent of the 'gap loading' condition of plane flow. The
condition (135) is of course even more stringent than (136).
Under the conditions

Ifll >> 1 r >> 1, R* <<R;**L<< 1 (138)


the primitive asymptotic approximation to (131)

1 sin fld: ~/( 1 + ].)a


1 + 2 -- F, (139)
V(~) ,-, ~ sin fir_, 1+2 '

may be obtained through asymptotic expansions of the expressions in


(133) similar to those used by Oldroyd[25]. Aschoff & Schiimmer[28] have
performed direct calculations to compute the shear stress from (133) and
have concluded that there was no significant error resulting from the use
of (139) in place of (133) when L , - 1 and

R* > 0.01 r R* > 2.25. (140)

It is clear from comparison of (110) and^(139), on equating flL in the


latter to/3 in the former and similarly ~L -1 to d:, that the approxima-
tion (139) is not based on a 'gap loading' assumption, i.e. the influence
of inertia is accounted for fully. In this respect it differs fundamentally
from the Taylor expansions proposed by Waiters [27], also Markovitz [4],
because (139) is derived via asymptotic expansions in inverse powers of
/3, while the Taylor expansions derived by Waiters [27,] involve positive
powers of/3. The approximation (139) is actually based on geometric con-
siderations, and succeeds so well, as claimed by Aschoff & Schiimmer[28],
because the Bessel functions in (133) have asymptotic expansions for large
flF, -rr < arg(fl) < re and 1 _< F ___ (1 + ]_,).

3.2 P e r t u r b e d S t e a d y Flows
The flows considered in this subsection are most often described
as nearly viscometric flows (cf. Pipkin & Owen[29]). However, here
the term viscometric flow is inclusive of the time dependent curvilineal
flows described in w and indeed of all the time independent, or steady
viscometric flows which in turn include the steady curvilineal flows. Thus,
the terminology adopted is that of the critical reference work by Truesdell
163

& Noll[6]. Nontheless, the analysis of perturbed steady viscometric flows


must be subject to consistency conditions on material functions which in
their most general form have been derived by Pipkin & Owen[29].
In all cases the velocity components in the appropriate curvilinear
coordinate system are assumed to have the form

?31 - - 0, Vi -- V(0
i ) (X 1 )-~-- V~I ) (X 1 , t)
"

for i - - (2, 3) (141)

where again equations (2), (3) and the analysis following them in w
applies. It is also assumed that the condition

[v(1)l
<<1 (142)
Iv(0)l
holds. The split (141) is compatible with (3) if and only if,

Ov 2 OV 3
Ox I - f ( x 1) [1 + q(x 1, t)] , Ox I = h(x 1) [1 + q(x l, t)] , (143)

where
[q[ << 1, (144)
so that in (11)

k(~; t) - -
( .~, + .y
/o q ( x ~, t - ,.)d~

-- ]~(0)(8; t) -~- ]~(1)(8; t).


) (145)

The condition (143) ensures that for all s finite

k(1)(8; t) ~ ]~(0)(8; t), (146)

and if q(x 1, t) is time periodic then all of w applies, i.e. k(1)(s; t) and
the extra stress tensor S are also time periodic.

]The flow considered by Tanner & Simm0ns[30] which is the sum of a


time periodic perturbation orthogonal to a steady plane shear is not a
curvilineal, or locally viscometric flow as defined by Truesdell & Noll[6].
164

For a time periodic perturbation of a steady viscometric flow the


parameter e in (40) may be chosen as

e -- max Iq(x I , t)l , (147)

so that in (19) and (21)

~Ys~ (k(0)(8; t)-t- k(1)(8; t)) -- ~s~ (k(0)(s; t))


-]- d~K20 (k(0)(8; t)). k(1)(8; t) -1- 0(6)
__ r-p(1) O(6),
- T(<~189 + ~ <12> +
(148)
with similar asymptotic forms obtainable in an obvious way for the other
components of S and in particular for the normal stress differences
Sl s%0 (') and 82 ~%o('). Therefore,
S - S (0) -~-S (1) -}-O(6), (149)

where
S - S,% o ( C t - I) - S ~ = o ( k ( s ; t)) , (150)

and if the pressure


p _ p(0) + p(1) _1_0(6), (151)

where p(0) is the pressure arising from the steady flow:

V p (~ - V. S (~ , (152)

then S (1) and p(1) are the perturbation stress tensor and pressure which
must be substituted in (48) to balance the perturbing motion with veloc-
ity v(1) ( xl , t).
It follows from the above that all of the classes of flow considered in
~2 may be treated in this way, subject to the proviso of their existence in
the presence of inertia, as before.
The consistency relations of Pipkin & Owen[29] relate the 13 inde-
pendent components of the tensor valued linear functional in the expan-
sion

s- + G 0 (st-oI0 ( - 8)- i, t153/


165

where g(s) is a general perturbation of a steady viscometric history, to


the viscometric functions

T -- "/'s=0(--")/8) ~1 - - S l c~ ~
- & e~
s=0
, (154)

and their derivatives with respect to 7. Dunwoody & Joseph[31] provided


their own derivation of these relations appropriate to a 'locally' two di-
mensional perturbation of the steady flow. Their relations involve just 9
of the components of the linear functional. Since the perturbations con-
sidered here are 'locally' one dimensional a sufficient set of consistency
conditions are

- . - .
(155)

derived from (17), (19)and

dS (o) / h 0G (o)
d7 07
(156)
= dS %o t)) t) "

07
The consistency relations quoted by Barnes et. a1.[17], attributed by
them to Pipkin[32], apply to a very special material class having a mul-
tilinear functional constitutive form. Bernstein[33] has obtained similar
consistency relations for the general K-BKZ fluid class by setting

~(1)(8; t ) - Ae Lwt ( 1 - e - ~ ) , A ~< 1, (157)

in (145), substituting the corresponding g(s) in (152) and then taking the
limit a~ ~ 0, but again some of his relations are peculiar to K-BKZ flu-
ids. Coleman & Markovitz[34] using the same method obtained relations
similar to Bernstein's for a class of second order fluids approximating the
fading memory fluids of Coleman & Noll[8].
The flows which have been dealt with in the literature using this
type of approximate analysis fall within the classes described in w so
that in the case of torsional and cone and plate flows (cf. Jones & Wai-
ters[35]) inertial effects must be neglected. Bernstein[33] in his analysis
of sinusoidal oscillations superimposed on steady plane shear flow of K-
BKZ fluids has taken no account of inertia. Barnes et. al.[17] have taken
166

account of inertia in their analysis of pulsatile pipe flow and also 'second
order effects', but have employed a special second order theory in doing
so. Phan-Thien[36] and others have followed suit. Their aim of course
has been to match their results to specific observations.
In all cases a single ordinary differential equation governs the spatial
dependence of the time periodic perturbation, so that the analysis differs
little from that required for the equation resulting from letting 7 -~ 0.
Indeed, Osaki et. a1.[37] interpreted their experimental data in terms of
the analysis of Markovitz[4] referred to in w However, besides the effect
that 7 7~ 0 has on the perturbation functionals (154), the interaction of
the perturbation with the basic flow has considerable practical interest(cf.
Jones & Waiters[35]). Both of the two previous nondimensional analyses
leading to conditions for the use of the 'gap loading' approximation may
be modified to be applicable to the corresponding perturbed steady flows.

4. L A R G E A M P L I T U D E F L O W S

The above description is applied to all those viscometric, time peri-


odic flows which are excluded from the considerations of w because their
amplitude of oscillation in strain satisfy neither (94) nor (146). For the
sake of simplicity the discussion is restricted to flows oscillating about a
rest state, and only plane shear and cylindrical shear flows are consid-
ered because the effects of inertia, and in particular its interaction with
non-linearity in the constitutive relation for the fluid, are the main in-
terest. Therefore, the work of Jones & Waiters[35] on cone and plate
flow while noted is not detailed. Besides, their work is limited to special
second order theories, which is also true of the work of Barnes et. a1.[17],
Townsend[38], Davies et. al. [39] and Phan-Thien[36] on pulsatile flow in
cylindrical pipes, and all are limited by the restriction that they involve
time periodic perturbations which are small in an absolute sense.
Here the aim is generality which is illustrated by application to a
K-BKZ model proposed by Luo & Tanner[22] to describe LDPE melts.

4.1 M o d e r a t e A m p l i t u d e Expansions
Considerable experimental data obtained from large amplitude os-
cillatory shear(LAOS) flows has quite recently been recorded by such as
Hatzikiriakos & Dealy[7], Giacomin, Jeyaseelan, Samarkas & Dealy[10].
Many of the perplexities which these investigators have attempted to ad-
167

dress in interpreting their data and observations, such as whether or not


there is 'wall slip', are avoided by resricting attention to oscillations in
the shear with amplitudes which are moderate, but in a sense dictated
by material parameters. Mathematically one is then also on safer ground
vis-a-vis the question of the existence of time periodic flows using the clas-
sical assumption of 'no slip' of the fluid at solid boundaries. Of course if
one neglects inertia such questions are trivially answered in the case of
plane shear, but unfortunately the predictions of the theory are at odds
with the observations for arbitrary LAOS.

~. 1.1. P l a n e S h e a r Flows.
For simplicity of presentation the shear stress is written in the form
(42)1 but as a function of the rate of strain using the identity

p k ( t ; p) - -O~f~(t; p) , ~(t; p) -- e-P~v(t - s)ds , (158)


f0 X)
where a simplified notation for partial derivatives is being introduced, to
obtain
S < 1 2 > -- O" -- 7" ( - - p - 1 O x V ) -- g ((:;gxV;p) . (159)

Then the equations of motion (52) and (53) governing plane shear flow
with a ( t ) - 0 reduce to the single equation in terms of ~(x, t)"

p-laxG(x, t) - O2tv(x, t) -F pOtv(x, t) , (160)

where the identity (44) has been used.


For strains which are moderate in some sense, best defined within
the context of the constitutive description of an individual material, it is
assumed that g ( 0 ~ ; p) has a polynomial approximation to any desired
degree of accuracy, i.e.

g (axy;p) - ~ g2r+l (axV,..., axv;p) <~ (161)


r:0
for some m, where the terms under the sum are formed by multilinear
operators defined on /(:2,-+1. It follows from (24) and (159) that these
multilinear operators must be of odd order. Such approximations may
168

be obtained by Taylor expansions involving Fr6chet derivatives or dif-


ferentials of the functional T~0(. ) if it is defined on a vector space of
functions intersecting /(:(cf. Liusternik & Sobolev[40]). But, there are
alternative to such esoteric methods as is now shown for a specific model.
It is important to obtain estimates of the error involved in such approx-
imation, rather than appeal to asymptotics based on infinitesimals(cf.
Dunwoody[21]. The use of asymptotic methods in viscoelastic fluid dy-
namics has had an unfortunate history, due to indiscriminate application.
The model chosen to illustrate the preceding argument is that of
Luo & Tanner[22], which is a specific form of the general K-BKZ model
designed to fit data obtained from experiment on LDPE melts(vide. Wag-
ner[41], Papanastasiou, Striven & Macosko[42]). For it the shear stress
in a plane shear flow is

a -- -- J~0 M ( s ) + k2(s;t) t ) d s ' M(s) - ~n --ak (162)

n an integer. If the flow is time periodic with frequency co, so that

k(s; t) - k s; t + - - 61C, (163)


co
and
Ik(s; t)l < c~2 V s 6 [0, ec) a n d t 6 ( - c o , ec) , (164)
then as shown by Dunwoody[21] the formula (162) for the shear stress is
approximated by

O" m
~ f (1 - e -2~/w'~k ) - 1 ak j[o27r/w -s k2r+l(s't)
,
r=0 k=l ~ e /ak (--a) ~ ds + R m
(165)
where the remainder
k2m+3(s;t)
Rm <E O<s<27r/~s
max ctm+l (166)

which is independent of t. The physically significant constant


n
E - ~ ak (166)
k=l
169

is the i n s t a n t a n e o u s elasticity as defined by (100). The equivalent of (165)


in the form (162) and the notation of (161)is

G -- -- M(8) ]~2r+1(8;t)

?n
7: :f::
~=o (_a)~ ds+Rm
(167)
= ~ _
g2r+l (Oxv,... OxV;p)-F nm.
r--O

D y n a m i c viscosities r/~*(a~) of order r may be defined in terms of the


multilinear operators g2r+l(') defined on the Fr&het space of functions
analytic in the right half of the complex plane. They are

?];(CO) -- g2r+l ((PAr-bDO)--I, -. ., (P-~-bCd)--l;P) (168)

which for the sample fluid have the complex values


n ~2r-F1
, akA k
r#~(co)- (-c~)-~(2r + 1)! ~ ]-[2r+l (169)

In particular,
r/~)(co) -- r/* (aJ) (170)
is the dynamic viscosity defined by (98), and it has the limits (99) and
(100).
A nondimensional form of the equation (160) is obtained in terms of
the nondimensional variables (104) except that the quantity e is replaced
A
by the ratio ~ , where A is the amplitude of the displacement of the
'upper' of two plates, a distance L apart, which contain the fluid between
them. In addition to (104)

p - - c@, fj2 _ A ~ , k(t) - - (A)


~ /3-10~,~, (171)

and

1,7:(~,)I-Z-- ~2,.+1 (o~v,..., o~; p) - g2,~+1 ( o ~ 2, , 0 ~ 2"p)


(172)
170

which generalises (104)4. It is noted that the equivalence

0~ _ (@) (p ~+ ' ~,) ** o~ ~ - (p ~+~)' (173)

the definition of r/; (w) in (168) and the multilinear property of the oper-
ators g2~+l (',. 9 "; p) imply that

!) ((i5 + ~ ) - 1 , . . . , (/5 + t ) - i ;i5) -- arg (q*(w)) . (174)

The nondimensional form of (160) sought is then

0~ [Dr02~+l (0~O,..., 0e#; t5) + / ) m ] -- R* (0t2O +/30tO) , (175)


r--0

where
R*= pwL2 ( wA ) 2r r/r*(w))
I~;(~)1' D,.- -~ rl~)(w . (176)

The parameter D~ has asymptotic limits as w --+ oc and w --+ 0. In


particular, for the sample fluid

2) ~
lim D ~ - aL 2 , (177)
CO--* OO

which corresponds with the limit (106)2 on R*. The low frequency limit
corresponds with (106)1 and is appropriate for 'retarded ' flow(cf. Dun-
woody[21]).
The set of D~, 1 _< r <_ m, where rn is arbitrary, have the norm

O<D- sup (D,.) ~ -/) w,~ .


l<r<m

This positive scalar, which is assumed to be bounded above as a function


of w, as it is for the sample fluid, acts as a measure of the degree of
nonlinearity in (175). If and only if D << 1 is the nonlinearity of no
consequence. Also, it is clear from (176) that the frequency not only
has an influence on the relevance of the inertial terms represented by the
171

right side of (175), but it also contributes to the nonlinear effects caused
A
primarily by relatively large ~ .
A nontransient solution to (175) and the boundary conditions

v(0, t) - 0, 9(1, i) - ~ \/5 + (179)

may be sought in the form of an asymptotic series"

v(:c, t) -- v0(z, t)-4- Dvl(z, t)-t- 9 9 9 9 (180)

As is to be expected the leading term in this series is given by (107) and


(110). The second and succeeding terms satisfy the boundary conditions

~(0, t ' ) - ~ ( 1 , t ) - 0 r__l, (181)

so that they are obtained successively and independently from

D1 oq~0 3 (oqs,~o ' oqs,~o, oqs,~o; 1~)

O, - R* + -
D2 0~5 (0~ ~o,.. , 0~ ~o; t3)
D

D1 0~.03 (0~,91 0~,g'o 0~9o;/5)


D ~
+ etc.
(182)
and so on, by the method of variation of parameters, or what is much
the same t h i n g - though better for computation- the method of Green's
functions. The unique solution of the homogeneous equations satisfying
the boundary condtions (180) is zero for all 1 _< r, so that it is then
apparent from the multilinear form of the operators in the right hand
sides that the particular solutions have the form

f~r(X; t) -- ~ ~[p'4- (2k q- 1)/,]-le(2k+l)~iXl(2k+l)(:~)


k=O
} . (183)

The variable coefficients Xl(2k. t_1) in these finite harmonic series are found
by solving, by either of the above methods, the ordinary differential equa-
tions obtained by substitution of (183) in the appropriate equation in
172

(182) and equating coefficients of e (2k+l)Li. The left hand sides in the
resulting equations involve the dynamic viscosities with argument w re-
placed by (2k + 1)w. The details of the calculations are described by
Dunwoody[21], who computed the actual shear stress occurring at the
upper plate using the Luo & Tanner[22] model and two terms in the se-
ries (180) only. His numerical results obtained for A / L - 2 , w - 100
rad/sec, and various Reynolds numbers are plotted in Figure 3, where
the upper profile is obtained from the linear theory.
If the gap loading assumption is made, equivalently R* is set to zero,
then the series (179) reduces to the single term

eti ) ~. (184)
- p+

This transform of the nondimensional gap loading velocity produces a


shear stress, constant across the gap,

ag- Iq~(w)] ( ~ ) 5 g +R~, (185)

where

~g -- ~ ~-~g2rA-1 -~- ^ ,... ~-t- ^ "/3 . (186)


,.=o t5+ ~ p- ~ '/5+ ~ p - ~'

Dunwoody[21] has computed the value of the shear stress crg again using
the Luo & Tanner[22] model for the same frequency and amplitude ratio
as above, but retaining four harmonics and terms O ([A/v/--dL]S). For
comparison a plot from his results is shown in Figure 4,where again the
upper profile is obtained from the linear theory, and it is apparent that
the higher order harmonics have a significant effect. The shear softening
due to nonlinearity alone is also significant.
It is also observable by comparison of the shear stress profiles in
Figure 3 and Figure 4 that the effect of increased inertia is to increase the
shear softening at the upper plate, which bears out the general conclusion
to be drawn from Figure 1. This observation can be compared with the
effect that a pulsatile pressure has on the mass flux of a fluid through a
pipe(cf. Barnes et. al.[17]).
173

R* - 0.0 R* - 0.7
60000
40000

-<0;of/ o 0.010.02k~3 0.010.0~03

~/:~176176176
o
R * - - 1.0 R*-- 1.2

40000 40000

-40000
0.0
\
3
--
0j/OoZeS
~j~oooo
-- 0
0.010 3

~L
Figure 3. Profiles of measured in Pascals as calculated by
A
A
Dunwoody[21] f o r - - 2, w - l O O r a d / s e c , both fixed, and
L
R* -- 0.0, 0.7, 1.0, 1.2. The upper profile was obtained from
the linear theory, and the ordinate scale varies with each plot.

75000
50000
250
,'5/
, , . . . . | . . . . | , / / . . . . . , . . . . | .

_o.o~ _o,o~, _o.// 0.01 0.02 ~ 3


ooo

000

-75000

Figure 4. Profile of ~r~L measured in Pascals as calculated by


A
A
Dunwoody [21] for - - - 2 , w - l O O r a d / s e c and R* -- 0.0, the
L
upper profile again obtained from the linear theory.
174

4.1.2. Cylindrical Shear Flows.


As in the case of plane shear flow either of the solutions (133) or (134)
may be made the basic term of a series solution of the equations (121) for
Couette flow, taking account of both inertia and nonlinearity in the con-
stitutive equation as in w With slight variations in the argument the
method to be employed is that described by Dunwoody[43], where essen-
tially just two terms in the expansion (161) were retained, which as seen
in w may not be quite sufficient for accurate quantitative analysis.
Also, as discussed in w the fluids likely to be tested in oscillating cylin-
der instruments would lead to circumstances necessitating full account of
inertia in the analysis; but, the use of asymptotic approximations such
as (139) would reduce the computations considerably.
As previously related, Barnes et. al. [17] have analysed pulsatile
flow in circular pipes, where the flow is supported by a pressure

P - Po - t5 [1 + ee ~'t] . (187)

For particular models they have shown via an analysis in which terms
O(e 3) are n e g l e c t e d - the theory is therefore a second order one - that
the mean flux of the fluid is enhanced by the presence of the oscillations in
the pressure, hence by inertia also, and that the effect is O(e2). Davies et.
a1.[39] highlighted the fact that there was disagreement between theory
and experiment in that the theory of Barnes et. a1.[17], as well as their
own based on a different model, predicted that flow enhancement should
decrease with the frequency of the oscillations while experiment showed
the opposite. Phan-Thien[36] used a generalised Maxwell fluid model
for the same problem, and he found that the mean flux enhancement
increased with frequency.

4.2 E x i s t e n c e of L A O S
In order that a time periodic flow exist it must be possible to repre-
sent the velocity with the property (25) in the form of a Fourier series
oo

, , kfk- ]fk]. (188)


k~--oo

To be specific, the assumption that a solution to the equations (52), (53)


and (21) with a - 1 a n d / 3 - 0 is gross if p 7~ 0, because it is evidently
175

not true in general that the shear stress in the left hand side will separate
into the form
S<12> -- ~ gk (f(x))e ~kwt , (189)
k'- -cx:,
where
f(z)-(...fl2(x ) , f l 1 (x) , f) (z) , fl (z) ,.. .) , (190)
in general.It is therefore surprising that a perturbation analysis for LAOS
on the assumption that R* << 1 has never been undertaken. It is the au-
thor's conjecture that plane, rectilineal, oscillatory shear flow of arbitrar-
ily large amplitude satisfying the standard no slip boundary conditions
does not exist, because the velocity must be two dimensional if plane
shear flow is to exist at all, i.e. it must be n o n r e c t i l i n e a l . If his con-
jecture is correct then it would explain the complexities encountered by
Hatzikiriakos & Dealy[7] and Adrian & Giacomin[44] arising from the
imposition of a unidirectional oscillatory displacement of large amplitude
on one plate of a parallel plate viscometer. Theories involving wallslip
(cf. Graham[45]) have been put forward to account for the complicated
'quasiperiodic' experimental data.
Nonrectilineal oscillatory plane shear of arbitrarily large amplitude,
based on work of Antman & Guo[46] on large amplitude shearing oscil-
lations of incompressible, nonlinearly elastic solids, is the subject of a
current study by the author.

REFERENCES

1. J.D. Ferry, Viscoelastic Properties of Polymers, Wiley, New-York,


1980.
2. J.L. Schrag, Deviation of velocity gradient profiles from the gap load-
ing and surface loading limits in dynamic simple shear experiments,
Trans. Soc. Rheol., 2 (1977) 399-413.
3. J.G. Oldroyd, D.J. Strawbridge and B.A. Toms, A coaxial-cylinder
elastoviscometer, Proc. Phys. Soc. B, 64 (1951) 44-57.
4. H. Markovitz, A property of Bessel functions and its application to
the theory of two rheometers, J. Appl. Phys., 23 (1952) 1070-1077.
5. J.S. Vrentas, D.C. Venerus and C.M. Vrentas, Finite amplitude os-
cillations of viscoelastic fluids, J. Non-Newtonian Fluid Mech., 40
(1991) 1-24.
176

o C. Truesdell and W. Noll, Non-linear Field Theories of Mechanics,


Encyclopedia of Physics (Ed. S. Flfigge), Band III/3, Springer-
Verlag, Berlin, 1965.
.
S.G. Hatzikiriakos and J.M. Dealy, Wall slip of molten high density
polyethylene I. Sliding plate rheometer studies, J. Rheol., 35 (1991)
497-523.
.
B.D. Coleman and W. Noll, Recent results in the continuum theory
of viscoelastic fluids, Ann. N.Y. Acad. Sci, 89 (1961) 672-714.
.
J.K. Hunter and M. Slemrod, Viscoelastic fluid flow exhibiting hys-
teritic phase changes, Phys. Fluids, 26 (1983) 2345-2351.
10. A.J. Giacomin, R.S. Jeyaseelan, T. Samarkas and J.M. Dealy, Va-
lidity of separable BKZ model for large amplitude oscillatory shear,
J. Rheol., 27 (1993) 387-410.
11. A.J. Giacomin and R.S. Jeyaseelan, A constitutive theory for poly-
olefins in large amplitude oscillatory shear, Polym. Eng. Sci., 35
(1995) 768-777.
12. V. Durand, B. Vergnes, J.F. Agassant, E. Benoit and R.J. Koop-
mans, Experimental study and modelling of oscillating flow of high
density polyethylenes, J. Rheol., 40 (1996) 383-394.
13. M.J. Crochet, A.R. Davies and K. Walters, Numerical Simulation of
Non-Newtonian Flow, Elsevier, Amsterdam, 1984.
14. M. Slemrod, Damped conservation laws in continuum mechanics,
Non-linear Analysis and Mechanics (Ed. R.J. Knops), Research
Notes in mathematics, Pitman, 1978.
15. P.H. Rabinowitz, Periodic solutions of non-linear hyperbolic partial
differential equations II, Comm. Pure App. Math., XXll (1969)
15-39.
16. J. Dunwoody, Time periodic rectilineal simple fluid flows, J. Non-
Newtonian Fluid Mech, 53 (1994) 83-98.
17. H.A. Barnes, P. Townsend and K. Walters, On pulsatile flow of non-
Newtonian liquids, Rheol. Acta, 10 (1971) 517-527.
18. B. Mena, F. Serrania and A. yon Ziegler, Further developments in the
use of oscillatory dies in polymer extrusion, Progress and Trends in
Rheology IV (Ed. C. Gallegos), Steinkopff-Verlag Darmstadt, 1994.
19. I.F. MacDonald, B.D. Marshe and E.A. Share, Rheological behaviour
for large amplitude oscillatory motion, Chem. Eng. Sci., 24 (1969)
1615-1625.
177

20. J.L. Ericksen, The behaviour of certain visco-elastic materials in lam-


inar shearing motions, Visco-elasticity :Phenomenological Aspects,
Academic Press, 1960.
21. J. Dunwoody. The effects of inertia and finite amplitude on oscil-
latory plane shear flow of K-BKZ fluids such as LDPE melts, J.
Non-Newtonian Fluid Mech., 65 (1996) 195-220.
22. X.-L. Luo and R.I. Tanner, Finite element simulation of long and
short circular die extrusion experiments using integral models, Int.
J. Numer. Methods Eng., 25 (1988) 9-22.
23. V.M. Kamath and M.R. Mackley, Redetermination of polymer re-
laxation moduli and memory functions using integral transforms, J.
Non-Newtonian Fluid Mech, 6 (1989) 119-144.
24. K. Walters, The motion of an elastico-viscous liquid contained be-
tween coaxial cylinders. II, Quart. J. Mech. Appl. Math., 13 (1960)
444-461.
25. J.G. Oldroyd, The motion of an elastico-viscous liquid contained
between coaxial cylinders. I, Quart. J. Mech. Appl. Math., 4 (1951)
271-282.
26. D.D. Joseph, Fluid Dynamics of Viscoelastic Liquids, Springer-
Verlag, New York, (1990).
27. K. Walters, The motion of an elastico-viscous liquid contained be-
tween coaxial cylinders. II, Quart. J. Mech. Appl. Math., 14 (1961)
431-436.
28. D. Aschoff and P. Schiimmer, Evaluation of unsteady Couette flow
measurement under the influence of inertia, J. Rheol., 37 (1993)
1237-1251.
29. A.C. Pipkin and D. Owen, Nearly viscometric flows, Phys. Fluids,
10 (1967)836-843.
30. R.I. Tanner and J.M. Simmons, Chem. Eng. Sci., 22 (1967) 1079-
1082.
31. J. Dunwoody and D.D. Joseph, Systematic linearization for stability
of shear flows of viscoelastic fluids, Arch. Rational Mech. Anal., 86
(1984) 65-84.
32. A.C. Pipkin, Modern Developments in the Mechanics of Continua,
Academic Press, London, 1966.
33. B. Bernstein, Small shearing oscillations superposed on large steady
shear of the BKZ fluid, Int. J. Non-linear Mech., 4 (1969) 183-200.
178

34. B.D. Coleman and H. Markovitz, Normal stress effects in second


order fluids, J. App. Phys., 35(1964) 1-9.
35. T.E.R. Jones and K. Walters, The behaviour of materials under com-
bined steady and oscillatory shear, J. Phys. A, 4 (1971) 85-99.
36. N. Phan-Thien, Pulsatile flow of polymeric fluids, J. Newtonian Fluid
Mech., 4 (1978) 167-176.
37. K. Osaki, M. Tamura, M. Kurata and T. Kotaka, Complex modulus
of concentrated polymer solutions in steady shear, J. Phys. Chem.,
69 (1965) 4183-4191.
38. P. Townsend, Numerical solution of some unsteady flows of elastico-
viscous liquids, Rheol. Acta, 12 (1973) 13-18.
39. J.M. Davies, S. Bhumiratana and R.B. Bird, Elastic and inertial
effects in pulsatile flow of polymeric liquids in circular tubes, J. Non-
Newtonian Fluid Mech.,3 (1977) 237-259.
40. L.A. Liusternik and V.J. Sobolev, Elements of Functional Analysis,
Ungar, New York, 1961.
41. M.H. Wagner, Analysis of stress-growth data for simple extension of
a low- density polyethylene melt, Rheol. Acta, 15 (1976) 133-135.
42. A.C. Papanastasiou, L.E. Scriven and C.W. Macosko, Integral con-
stitutive equation for mixed flows : viscoelastic characterization, J.
Rheol., 27 (1983) 387- 410.
43. J. Dunwoody, A comparison of periodic flow of moderate amplitude
in two viscometers, Developments in Non-Newtonian Flows (Eds.
D.A. Siginer and S.E.Bechtel), ASME, FED-Vol. 206, AMD-Vol.
191, 1994.
44. D.W. Adrian and A.J. Giacomin, The quasi-periodic nature of a
polyurethane melt in oscillatory shear, J. Rheol., 36 (1992) 1227-
1243.
45. M.D. Graham, Wall slip and the nonlinear dynamics of large ampli-
tude oscillatory shear flows, J. Rheol., 39 (1995) 697-712.
46. S.S. Antman and Z.-H. Guo, Large shearing oscillations of incom-
pressible nonlinearly elastic bodies, J. Elasticity, 14 (1984) 249-262.
179

S E C O N D A R Y F L O W S IN T U B E S OF A R B I T R A R Y S H A P E

Mario F. Letelier a , Dennis A. Siginer b

a Departamento de Ingenieria Mecdmica, Universidad de Santiago de Chile,


Casilla 10233, Santiago, Chile

o Department of Mechanical Engineering, New Jersey Institute of Technology,


University Heights, Newark, NJ 07102-1982, USA

1. INTRODUCTION

Conduit flow is a common occurrence in many industrial and biological


systems. It is also, in some cases, a convenient approximate model for studying
fluid motion through porous media, filters, tissues, and other slow fluid motion
in complex solid matrices. Current technological advances in many fields require
a good understanding of the dynamics of fluids other than Newtonian in conduit
flow. This knowledge is necessary in order to estimate energy loss, transport
properties, and many other variables of industrial interest.
Viscoelastic fluids constitute an important class among non-Newtonian fluids,
the study of which is rendered more difficult by several properties and
phenomena exhibited by these fluids such as stress relaxation, strain recovery,
die swell, normal stress differences, drag reduction, and flow enhancement [1].
The flow of non-linear viscoelastic fluids in non-circular pipes may lead to the
occurrence of secondary flows, a phenomenon not well covered in the technical
literature. Secondary flows have a significant influence on important industrial
phenomena, such as transport, and energy loss.
A large number of competing viscoelastic constitutive models exist to predict
flow phenomena. Integral models seem to predict experimental data better [2]. In
this chapter, the simple fluid of multiple integral-type models with fading
memory is considered. Secondary flows are determined in the case of laminar
180

longitudinal flows in approximately triangular and square conduits, when the


flow is driven by small-amplitude oscillatory pressure gradients.
The chapter is organized as follows. The mathematical background is
developed and the summary of a novel analytical method devised by the first
author and co-authors for determining the velocity field of laminar Newtonian
unsteady flow in non-circular pipes is presented in Section 2. This is followed by
an analysis in Section 3 of the pulsating flow in circular pipes of a viscoelastic
fading memory fluid of the multiple integral type. Results of these sections are
combined in the next section, where a mathematical expression for the axial
velocity is developed for flow in non-circular pipes driven by a pressure gradient
oscillating around a non-zero mean. The chapter closes with Section 5 where
analytical steps that lead to the determination of the transversal velocity field are
developed. Plots that depict the main features of axial and secondary flow fields
are also presented.

2. NEWTONIAN UNSTEADY FLOW IN NON-CIRCULAR PIPES

Fully developed Newtonian flows in pipes are rectilinear when they are driven
by a longitudinal pressure gradient, properties are kept constant, no wall-suction
is considered, and the pipe is a straight cylinder of arbitrary cross-section. No
secondary flows are possible under these conditions.
The linearity of the equations of motion makes it possible to use a wide
variety of analytical methods that yield either exact closed-form, or approximate
solutions. In this section, a summary of these results is presented based on a
unified approach that is applicable to an infinite spectrum of cases where the
main differentiating factor is the pipe cross-section geometry.

2.1. Analysis
The flow of a Newtonian, incompressible, and isoviscous fluid through
straight tubes of constant-cross-section is governed by a linear form of the
Navier-Stokes equations when the flow is parallel and laminar. In this case,
velocity is of the form,

u = (0, 0, w), (1)

where w is the axial velocity. If z is the axial coordinate, then, the continuity
equation reduces to,
181

c~
- o, (2)
c~

and w only is a function of the coordinates attached to the cross-section, and of


time. The Navier-Stokes equations become,

0t9 3t9 c~ 319


0 - 0 - , p - +/.tV 2 w, (3)
~x' ~ 3t ,~

in which Cartesian coordinates have been used. In this, P is the piezometric


pressure, that is, P = p + p g ( and P, (, g, /9, fl, t represent the fluid pressure,
vertically upward directed length coordinate, acceleration due to gravity, fluid
density, dynamic viscosity, and time, respectively. Since in equations (2) and (3),
w = (x, y, t), and P = P (z, t), physical homogeneity demands that,

319
= pf(t), (4)
3z

where f(t) is an arbitrary function of time, usually called the "forcing function"
of the flow.
The velocity shown in equation (3) is subject to the no-slip condition at an
arbitrarily curved boundary 3D~ and, in general, to an initial condition for t=O,
that is,

w (x, y , t) --- 0, w (x, y, 0) = w (x, y). (5)


DE

Closed-form exact solutions of equation (3) for non-circular pipes are


available for a few cross-section shapes and a variety of forcing functions. Some
well-known contributions are those of Mt~eller [3] and Yih [4], for concentric
annular pipes, Fan and Chao [5] for rectangular pipes, and Hepworth and Rice
[6] for circular and annular sector, and rectangular shapes.
Most of the results obtained are for simple types of forcing functions such as
start-up, oscillatory, linearly accelerated, and similarly driven flows. Analytical
methods used include separation of varfables, eigenfunction expansion, and
Duhamel's principle.
182

The method of eigenfunction expansion provides a convenient tool for finding


new solutions for cases where cross-sectional geometry makes possible the
definition of associated eigenfunctions. In general, the equation of motion can be
written as,

c~ _ vV 2 w =f~ +f2(t), (6)


3t

where v = ~ / 9 is the kinematic viscosity, and, where the forcing function has
been split into a steady-state, constant part f,, and, a time-dependent part f2 ( t ) .
The linearity of equation (6) allows one to define,

w = w~(x, y) + w2(x, y, t), (7)

and to rewrite equation (6) as,

vV ~ w. = f . . C~V2h vV ~ w:, = O. (8)


dt

3Wa p
v V 2 w 2 p - f 2 ( t ), Wz = Wz h + w zp" (9)
dt

Let E (x, y ) be an infinite set of admissible eigenfunctions (m, n = O, 1 , 2 . . . ) ,


in which Cartesian coordinates are used with eigenvalues a, and /3,,. Then, it
follows that,

mn
-_
)2Om
[( ] +
~12J

Relationships shown in equations (8) and (9) lead to

oo oo

w, = Z Z~m~ , ~.. = A m2. ( L / v ) 2~ Z AmnEmn = l , (10)


o
m-O .--0 am + 12)
183

W2h = E Z CmnEmn e2mnt "/~mn= V a m nt" (11)


m=O n--O ' [\l 1 ) kl 2 ) '

ao aO
W2p= ~_. ~_,Omn(t) Emn " (12)
m=O n=O

Here, A~. are constant coefficients determined by means of Fourier-series


analysis. C,,, are constant arbitrary coefficients determined by applying an initial
condition and the orthogonality of the eigenfunctions, l~ and l 2 are appropriate
length scales, and the time-dependent coefficients Dmn a r e to be computed by
solving equation (9)~. When equation (12) is substituted, equation (9)~becomes,

dDm, -JrVF(IQC'm~2 -~-(i~n ~2] Omn = Amnf2. (13)


dt I\ l 1 ) \ 12 )

Only the particular solution in equation (13) applies, since the homogenous
solution has been already taken into account in equation (11). An example of this
kind of exact analysis is applied to the problem of a flow that decelerates under a
zero pressure gradient starting from a given initial steady state velocity W(x, y) in
a rectangular pipe of cross-section 2a(x-direction) and 2b(y-direction). It is
assumed that the z coordinate lies along the pipe axis, so that the initial and
boundary conditions are,

w(x,y,O + ] : W(x,y), w(+ a,y,t) : w(x, + b, t) : O, (14)

further,

f =0, w(x,y,t --+~o)-0.

A convenient set of eigenfunctions for this problem is defined by,

(2m+l)x (2m+l)y
Emn = cos cos , m,n = 0,1,2... (15)
2a 2b
184

The initial steady-state velocity can be obtained by solving equation (8)~,


where the constant f l will be defined as fl = k for clarity's sake. This solution
is shown in equation (10). Thus,

W= 64k
oo
Z ~
oo
( - - l ) m+n 2 Bran"
(2re+l) (2n+l) (2re+l) (2n+l)
a~ + b2

Since the pressure gradient driving the flow for t > 0 + is zero, the time-
dependent velocity obeys equation (8), whose solution is shown in equation (11).
Thus,

64k ~ oe (_1)m+n - - 2r ant


w= 4 ZZ Em, e
rc r m=~ "=~(2m + l)(2n + l ) I ( 2 m +~ + ( 2n+l)2]bi '

2, -
n~v I(2m+1)2 +
(2n+l)~
___
1. (16)
mn 4 a~ b~

Another useful approach is Duhamel's principle. Using Duhamel's principle,


velocity is found through an integral superposition method. If Ws(x,y,t) is the
velocity which results from a suddenly imposed unit step-function pressure
gradient, then, for a general forcing function f, velocity is obtained using
equation (6),

w(x,y,t) = I t w s (w,y,t - s) d f (s)ds + f ( O ) w s (x,y,t) (17)


o -~s "

The structure of equation (17) shows that the boundary condition is met, since
that condition is first imposed on ws. Also, the initial condition follows from W,,
that is,

w (x, y, O) = f (O) w, (x, y, O).


185

The associated start-up velocity of the flow has to be worked out first in order
to apply equation (17). This is not necessary when the eigenfunction method is
used, since the transient response to any forcing is incorporated in equation (13).

2.2. Boundary perturbation approach


Exact analytical methods are very restrictive as to the scope of admissible pipe
shapes. A more flexible approach was developed by Letelier and Leutheusser [7],
and Letelier, et al [8]. An updated version of the approach is presented here. In
the following, cylindrical polar coordinates (r, O, z) will be used. A steady state
expression for the axial velocity is given by,

f (R 2 r 2 (18)
w~ = 4 v - ) +e,w~,

where w,~ is a homogeneous solution of equation (8)2, R is a constant length, and


E, is a small arbitrary parameter. For ~ -+ 0 the velocity takes the classical
Poiseuille's form. The velocity should meet the boundary condition given in
equation (5)1 on a closed curve, and be unique within the domain. This can be
achieved if,

w,~ (o, o = o), Ic, I ~ Ic,~], (19)

Condition (19)1 makes the velocity unique at the center of the pipe, and
condition (19)2 restricts the absolute value of c~ to a value equal, or less than
some critical value e~c, in order that the contour be a closed curve. A convenient
set of functions to this end are,

= ~ pn sinnO (20)
Wlh
L r n cosnO

where n is any positive integer. For each value of n, the small parameter ~c1 has
some specific physical dimension stemming from the physical homogeneity of
equation (18). For the sake of convenience, a dimensionless form of equation
(18) can be used by selecting R as length scale for r, f l R 2 / 4 v for Wl, and
fl / 4 v R n-2 for OC'l.In dimensionless terms, equation (18) becomes,

w~ = 1- r 2 + ~1rn sin n O, (21)


186

and the pipe contour is given by

1 - r 2 + ~ r n s i n n 0 = O. (22)

A wide variety of contour shapes develop from equation (22) according to the
given values of ~, and n [8]. As l e~ I increases starting from c1 =0, the circle
changes continuously up to a shape that exhibits a number of sharp corners, or
cusps, equal to the integer constant n, and is obtained for ]c, ]= ]~]. When
]c~ I is greater than its corresponding critical value, then equation (22) no longer
describes a closed contour. For appropriate values of Cl equation (22) describes
ellipses (n = 2), approximate equilateral triangles (n = 3), approximate squares
(n = 4), and so forth. The critical values of ~ can be found by applying the
boundary condition, and the condition that at a cusp the velocity gradient should
be zero. This leads to,

n-2

I~lcl= - ( ~ n-2-2
)n2 . (23)

In general, expression (21) can be made up of several different components


such as given by equation (20), that would lead to many other shapes. For steady
flow, equation (21) is an exact solution of the Navier-Stokes equations.
The perturbation approach described here draws on both the steady state
solutions in non-circular pipes and on the available closed-form unsteady
solutions of flows in circular pipes. Let w~ be the velocity distribution of an
unsteady flow in a circular pipe driven by an arbitrary forcing function. Then
velocity for an unsteady flow driven by such forcing function in non-circular
pipes, can be conveniently expressed as,

w= ( 1 - r z + e~r" sin n 0 ) I we
-r ~ + c lF~ + c 2F z +... ), (24)

where Fj (r, O, t) ( j = 1, 2 .... ) are unknown functions to be determined from


equation (6). It has been shown that, in general, w~/(1- r 2) is finite for r = 1
[8]. In equation (24) the steady state velocity is now a "shape factor" that
becomes zero at the pipe contour.
187

In the following, the analysis is carried out for the case of an oscillatory flow
superimposed on a steady flow. It is, thus, assumed,

f = k~ + k 2 sin cot, (25)

where k, and k2 are constants. If the reciprocal of the angular velocity co is taken
as time scale, then equation (6) becomes, in dimensionless variables,

f) 0w~ _ vZ Wc = 4 + A sin t , if2 - (~ , 2 - 4k2 . (26)


Ot v k~

Velocity field is expressed as

w~ = 1 - r z + A ( r ) e it + A(r) e -it , (27)

where A(r) is a complex unknown function and A is its conjugate. Substitution of


equation (27) into equation (26) yields,

V 2 A - iA - iA, (28)
2

with

A -- 2, I1_o.( t"/~r)_- 11 A = a , + ia 2. (29)


2f'2 L I o(4iff2) '

The functions al(r ) and a2(r ) can be obtained from equation (29)1 by splitting
the modified Bessel function I o into its real and imaginary parts. In this form, the
expression for the velocity is,

w~ = 1 - r 2 + 2 (a~ cos t - a 2 sin t). (30)

Different versions of equation (30) have been developed in the past by several
authors, among them Sexl [9], and Uchida [10]. Since ~, is a small parameter
188

( [ G ] < 1 for n _> 2), a perturbation approach can be used to find the unknown
functions Fj. To this end w is expanded as,

W=W (0) "-~-oC'1 W(1)-'~ - oO? W (2) "+ ..., (31)

and it follows from a comparison of equations (31) and (24),

w (~ c = l - r z + 2 ( a ~ c o s t - a 2 sin 0 ,

w ( ' > = ( 1 - r 2) F 1 + w~r"


1-r 2 sinnO, (32)

w ( 2 > - ( 1 - r 2) F2 +r"F~ sin nO,

where,

~3v~ d2 1d 1 d2
V 2 w (i) = 0 " i = 1,2, .. V2 - + -~ + (33)
v2 ' " ' 3r 2 r Or r 2 ~5~2 "

An appropriate solution for W (1) in equation (33) can be found by defining,

W(~)= (.4 (') eit+ A(')e-;' )sin nO+ Dr" sin nO, (34)

where A ~1) is a complex function of r, and D is an arbitrary constant. Upon


substitution of equation (34) into equation (33) it follows that,

d2A(')2dr + rl dA(1)
(dr i ~ n_~) A <~) = 0. (35)

The admissible solution of equation (35) is the modified Bessel function I,


Thus,

A ~l> = C (I) I, (i~r). (36)


189

The constant C~) is complex, to be determined from physical continuity


requirements. Defining,

C (l) = C(l) + iC) ') , I, (~/if2 r) = In1 (r) + In2 (r),

we obtain,

A (1) - - a(~) + ia~ ~) , a(~) = C(~) I~, ( r ) - C ~ ~) 1.2 ( r ) ,

a~ ') = C~ l) Inl (r) + C(1) 1,2 (r). (37)

The first order velocity is now given by,

w (') = 2 (a( 'J cos cot - a z(~) sin cot) sin n 0 + D r n sin n 0. (38)

Setting D = 1 and equating equations (38) and (32) an equation for F/is found,

0-r2)Fl =2 [(a~1) -rnpl)COS t-(a~ 1) - r n p 2 ) s i n t] sin nO, (39)

al a2
P ~ - l_r 2 , P2- l_r 2.

The functions p~ and P2 are continuous functions of r. Since for r = 1 the


LHS of equation (39) is zero, this condition leads to,

a~') (1) - p~(1), a 2~) ( 1 ) = P 2 (1) (40)

which determines the constants in equation (37). In this way F, becomes a


continuous function of r throughout the domain. Equation (32)3 defines the
second order velocity, where,

w ~2) = (1-r 2) F 2 + rn(F~ cos t-F~2 sin t) sin 2nO,

F~(r) = 2(a~) - r " p , ) F~2(r) = 2(a~ ~)- rnp2) (41)


l_r z ' l_r 2 9
190

Equation (41) may be rendered linear in trigonometric functions by using the


relation,

sin 2 nO = 1- cos2n0

Physical continuity at r = 1 demands that w ~2) should be of the form,

w '2' = 2(b~ c o s t - b 2 sin t ) + 2(a( 2' c o s t - a ~ 2' sin t ) c o s 2 n 0 . (42)

The unknown functions b, (r), b2 (r) can be obtained from a 0 - free solution of
equation (33),

B(r) e i' + - B ( r ) e -it , B(r) = C o I o (x]~r), C O = Co, + i Co2, I o = Iol +ilo2, (43)

and similarly,

A ~2' : C ~2' 12, (x/-f-t-~r) = a~2) + ia~2' , C ~2) : C(2) + iC~2' , I2n : 12n I .qt_ //2n2 "

The four c o n s t a n t s Col , C02 , C~ 2) and C~2) can be determined by equating


equations (41) and (42), and setting r = 1. In this way, the expression for F 2 is
derived as,

2 r "F~I cos t - b2 sin t +


F2 - 1 - r 2 bl 4 4

2 [(a(2)+ r n F ~ ) c o s t - ( a ~ 2 ) + rnF~z)sint 1 cos 2n 0. (44)


1-r 2 4 4

Higher order terms can be found in a similar manner. Computations for n = 6


and several values of .O were made by Letelier et al [8], for the case of a purely
oscillating flow. It is noted that the pipe contour approaches a regular hexagon,
in this case for e, = 0.1481. Wall shear stress was also computed at finite time
intervals.
191

This method can be applied to flows other than oscillatory, as long as it is not
necessary to impose an initial condition. Such a condition cannot be satisfied by
the present analysis.

3. VISCOELASTIC FLOW IN CIRCULAR PIPES

Steady, fully developed, viscoelastic pipe flows are rectilinear when the pipe is
a circular cylinder [11, 12]. This is true even when the flow is unsteady,
regardless the constitutive behavior. Some special results of practical interest
arise from non-linear effects. One of these results is flow enhancement in
periodic flows. Several authors have explored this important aspect both
theoretically and experimentally [1, 2, 14 - 18]. Integral models seem to simulate
experimental data better than other models.
An example of integral models is Green and Rivlin's [19] simple fluid model
of the multiple integral type. For incompressible fluids, the total stress T, can be
expressed in terms of the strain history G, on a particle X,

T + p l = Fs:o[G(X,s)] , s = t-r, (45)

where p, t and r are the pressure, time, and past time, respectively. Following
Green and Rivlin, the functional F is expressed as a Fr6chet series,

F [ G ( X ; ~)] : e.S <') + ~2S(2) --1-oo3S (3, -+- 0(0o 4 ) (46)

in terms of the small parameter e. S ~i) is defined as the Fr6chet stress at order i.
The first two terms in equation (46) are,

= I o4(S)a, (s)ds, (47)

oo oo
S(2) "-- I 0 ~(S)G2 (s)ds -]- I o I o Pl (s1, $2 )al (s1)a2 ($2)dsldS2, (48)

where the history of the strain has already been expanded in a series in c,

G(s) = o~G1(s) + c2G2 (s) + 83G 3(s) + 0(6~ 4 ). (49)

Similarly, the velocity and piezometric pressure are written as


192

c~3 OO

u(X,t," c) = ~ c n u ( " ) ( X , 0 ; P ( X , t ; g ) - zo~np(n)(x,t). (50)


n =1 n=l

The equations of motion and boundary conditions are,

Ou
p --VP+VoS ; V.u=O inD,, (51)
Dt

D, - {(r, O,z).'O < r < R,O < 0 < 2rc, - m < z < +~},

u (R ,0 ,z) - O, u ( O , O , z ) < ~ .

We define u = (u,v, w). The axial symmetry of the flow requires u = v = 0


and, w (-~) = -w. Siginer [2] presented a solution for pressure gradiems
oscillating about a non-zero mean. First and third order velocities were found in
closed form. Flow enhancement appears at the third order of the analysis. This
does not imply that the enhancement effects are small. On the contrary, as
predicted by the analysis, they are quite large, as the series expansion outlined
above may be considered asymptotic in nature. The second order velocity is zero,
as the axial velocity is odd in c.
A general analytical method for circular pipes can be developed by extending
the eigenfunction approach presented in the previous section to the first order
problem. The equation of motion is given by,

~42 (1)
P - I o G ( s ) V2 w(')(r, t-s) ds + p f ( t ) . (52)
c~

Let,

w (~) = Z C, (t) Jo , r , (53)


n=l

where Jo is the Bessel function of first kind and order zero, and
Jo(bn) = 0; n = 1,2, .... Substitution of equation (53) into equation (52) gives,
193

dC, + R 2 ~ o G(s ) Cn (t -- S) as ~ D n p f (t ), (54)


P dt

with

oo

ZDnJ ~ (b n r
n:, ~):1.

The time-dependent coefficients C. can be obtained from equation (54) with


the help of an appropriate initial condition. For oscillatory forcing around a non-
zero mean,

P f = Pfo + A sincot. (55)

C n is given the form,

C n : Cnl e ic~ q- Cnl e -i~ q- C ,

that when using equation (54) yields,

R2
Cnl = (17"- p(oR 2 -ib2Dn#7'),
2[(b.2r/') 2 + (pooR 2 - r/") 2]

C= fo R2 D.
P bZ" , /.t = ~o G(s)ds, (56)

77 - ~ o G(s)e- ~<~ - ~7' - i~7", (57)

From which we obtain,

R 2
w ~'~ = Z c~ljoR~ Dn
. ~ 2 (b~. u')~ + (poor ~ _ u")~ [(u"- poor ~-) cos cot
n=l bn
194

+ D. b.2 r/' sin co t]} Jo (b. R ) . (58)

The shear relaxation modulus G(s) determines the viscosity/z in equation (56)
and the complex viscosity 7-/*in equation (57). The constant parameters fo and
are assumed to be 0(1). Other cases can be analyzed for different forms of the
forcing function, provided it remains bounded.
Another analytical approach can be developed by exploiting the linearity of
equation (52), through an extension of Duhamel's principle, in the form,

t df (59)
w (') : I oWsv(r,t- 4 ) - : = ( 4 ) d 4 + f(O)w~v(r,t),
ag

where Ws~ is the start-up velocity for the viscoelastic fluid at O(E), and for a unit
pressure gradient given by,

Wsv(r't)-2pR2 ~-'~J ~3 (1 - e -x"' ), An = - ~0~176 (60)


p .--, b.J, (b.) p

4. LONGITUDINAL VISCOELASTIC FLOW IN NON-CIRCULAR PIPES

To order c, the flow velocity field in non-circular pipes is governed by,

02
Ow
p--=r
oo
0
G(s)V2w(r, O. t - s)ds + pf(t), V 2 - 3z 1 0
+ - /"
1 ~2 ~ (61)
c,-r2 O-9F + _ /,.2

where cylindrical coordinates (r, O, z) have been selected and the superscript in w
has been dropped. The condition at the boundary cO, is,

wl =0.

Following the analysis presented in Section 2, velocity is assumed to be of the


form,
195

w = (R2-r z + F~,r n sin n 0) [w o (r, t) + c,H,(r, O, t) + ~ H 2 ( r O , t)+O(c3)].(62)

Where wo(R 2 - r 2) is the closed form axial velocity of the flow in a circular
tube. The analysis that makes it possible to find the unknown functions
H i ( r , O , t ) , i = 1,2,.., follows a close parallel to its counterpart for Newtonian

flow as presented in Section 2. The main steps that lead to the determination of
the velocity field for pulsating flow in pipes of arbitrary cross-section are
described in the following using dimensional variables. Flow is driven by the
pressure gradient defined in equation (55) with constants j~, and 2 of 0(1). The
base flow that is the flow in a circular pipe under the forcing given by equation
(55), has already been solved in Section 3, and is described by equation (58) in
exact form. Assuming c~ small, equation (62) can be rewritten as a series in ~ ,

w = w ~~ + clw ~'~ + e,w ~2~ + 0 (e~), w (~ = wo(R 2 - r2), (63)

W(1)= WO rn sin n 0 + H~ (RZ-r2), w ~ = r n Hj_~ sin n 0 + H j (R2-r2). (64)

Equations associated with equation (64) are,

~(J) (j)
P cTt -~o~G(s)V2w (r, O , t - s ) ds ; j _> 1. (65)

In order to take advantage of the similarity between this problem and its
Newtonian counterpart presented in Section 2, the velocity at zero order in ~,
will be written as,

w(O) _ Pfo (R2_r 2) + 2 [a 1(r) cos cot - a 2(r) sin co t], (66)
4/a

where a, (r) and a2(r ) a r e infinite polynomials explicitly available from equation
(58). Note that several symbols used in previous sections are used in this section
with equivalent meaning, but they do not necessarily have the same algebraic
structure. For example, the functions a, and a 2 have the same meaning in
equations (30) and (66), but their algebraic structure is not the same in both
equations. Assuming w ~ has the following structure,
196

w"' = (A ~'' (r) e i~ + --A(" (r) e -ia~ ) sin nO + D r" sin nO, (67)

we obtain from equation (65),

d2A (~) + 1 dA (~) rt 2


dr 2 r dr - (12 + - 7 )A(1)= 0, 12
. . . . . .

ipco (68)
r]*'

with r/* given by equation (57). The admissible solution of equation (68) reads as

A (') = C (') I, (Ar) = a(') + ia~ '), C (') = C~ ') + i C~ '). (69)

A comparison of equation (64)~ and equations (67) through (69), yields,

2
H, = h, (r, 0 sin n O, h , - [(a( ~)- r ' p ~ ) coscot-(a~ ')- r ' p 2 ) sin cot],
R2_r 2

al a2 (70)
P ~ - R2_r 2 , P2- R2_r 2.

Similar to the steps which led to equation (40) in Section 2, the required
continuity of the function h~ for R = r leads to a set of algebraic equations for
the constants C(~) and c"<~) The expressions for wc~) j > 1 can be worked out
"'2 ~

following the steps shown in equations (41) through (44).


Computations of equation (62) have been made by Letelier and Siginer [20] for
some values of the fluid, flow, and geometry parameters. The fluid constants
were selected following Siginer [1], thus,

G(s) =/-to - U'oo Sk-I s/O 0= al , (71)


k( r/'oo-p)

Here r/'oo is a second Newtonian viscosity at high frequencies,


O, a,, 0 <k _<1, F and d are the relaxation time, the first Rivlin-Ericksen
constant, the viscosity index, and the Gamma and Dirac functions, respectively.
The specific values chosen for the material constants are,
197

r/'~ = O, a~ = - 5 0 g / c m , /-to = 200 p o i s e , p = 0.89 g/cm 3.

Other parameters concerning the flow and the geometry are given as,

co = 10 r a d/s , P f o _ I,
2

and the geometrical constants are,

R = 3cm, ~ = 0.3845 for n = 3, ~ = 0.22 for n = 4.

The selected values of n and the associated values for E1 determine through
equation (22) pipe cross-section contours that are approximately triangular and
square. The longitudinal flow in these pipes exhibits a very complex pattern that
is very much dependent on time and on the effect of sharp corners.
In Figures 1 through 6 the isovels in the cross-section of the pipes are shown
for one value of the viscosity index k. The specific times at which the isovels
were drawn are such that the velocity profiles undergo several inflexion points,
as shown in the companion plots. At such times, local velocity extrema appear at
several points, a phenomenon that tends to increase as k decreases. All plots are
dimensionless. The length scale for r is R, and the velocity scale for u is
PR 2/4f10. Velocity profiles refer to the larger half symmetry axis, and to the
shorter half symmetry axis for all figures.
198

F i g u r e 1. I s o v e l s f o r n = 3 ; c= 0.3845 ; k = 0.007 ; t = 0.14257r.

F i g u r e 2. I s o v e l s f o r n = 4 ; c= 0.22 ; k = 0.007 ; t = 0.257r.


199

1 "

0.8

0.6

0.4

0.2

0 I I I, I " ~
0 0.002 0.004 0.006 0.008 0.01

Figure 3. V e l o c i t y profile at shorter symmetry axis for n = 3 ; ~ = 0 . 3 8 4 5 ;


k = 0.007 ; t = 0.1425rc.

2 ! . . . . . .....

~I
o ...... i
0 0.002 0.004 0.006 0.008 0.01

Figure 4. Velocity profile at larger sylmnetry axis for n = 3 ; e = 0.3845 ;


k = 0.007 ; t = 0.1425rc.
200

1-
r

0.8

0.6

0.4

0.2

0 I I

-0.035 -0.03 -0.025 -0.02 -0.015 -o.o~ -o.oo5 o

Figure 5. V e l o c i t y profile at s h o r t e r s y m m e t r y axis for n = 4 ; e = 0.22 ; k = 0 . 0 0 7 ;


t - 0.25~.

1.4
r
1.2

0.8 -

0.6-

0.4

0.2

o I I

-0.05 -0.04 -0.03 -0.02 -0.01 0


U

F i g u r e 6. V e l o c i t y profile at l a r g e r s y m m e t r y for n = 4 ; c = 0.22 ; k = 0.007 9


t = 0.25r~.
201

5. SECONDARY FLOWS

Reported studies of secondary flows of viscoelastic fluid are scarce. Such


studies have been restricted to a limited set of geometry, such as elliptical and
rectangular [12]. Xue, et al [12], apply an implicit finite volume method to a
modified Phan-Thien Tanner fluid in rectangular tubes when the flow is steady.
Numerical computations describe the transversal flow velocity and pathlines for
several sets of relevant parameters. In Xue, et al [12], also provided a review of
the relevant literature.
In this section the basic aspects of an analytical method for computing
secondary flow in pipes of arbitrary cross-section are presented when the fluid is
described by the multiple-integral simple fluid model, and the flow is unsteady.
The method will be applied to a pulsating flow driven by a forcing of the form
defined in equation (55).
At O(d) the equations of motion read as,

8u (2)
=- VP (2) + V'S (2) ; V'u (2) - 0 in D,, (72)
c~

u I =0 ;u(O,O,z)<~. (73)
0D,

The velocity and extra-stress are given as,

U (2) -- U ( r, O, t) e r + v ( r, O, t) eo, (74)

S(2) - I oG(s)A(2) (t - s)ds + I oG(s) L, (t - s)ds +

(75)

The transversal velocity components are u and v, and e, and e o represent the
radial and azimuthal unit vectors. At the second order the axial velocity
component is zero. G(s) and 7"(s1,s2) represent the shear relaxation and the
quadratic shear relaxation moduli, respectively. The first Rivlin-Ericksen tensor
A1O) is of the form

A( j~ (t-s)= 2D [u (s) (X, t-s)] , j =1,2, (76)


202

and, following Siginer [2],

L I (t - s ) : ~ * ' V A~ l) + A ~ ' ) V ~ * + (A~I)v~*) T , (77)

4 - I ~u ''~(x, ~')d~', t)~:, 4" - *ez. (78)

The longitudinal velocity, u a) =wez is shown in equations (62) through (70).


The longitudinal velocity substitution into equations (76) through (78) using
equation (72), leads to the transversal momentum equations which define the
velocity field (v, w),

U~t --- _ P~r ~oG(s)


2
2u- - T v , O - ds

-r (rW'r r + --~ (W'r ~,o + w , o ~, r ),o- 72 w'O r ds

~
+J0JoY(S,,S2) { _l [rw, r(S,)W, (s2)]
r r

1 }
"~- V [w'r ($2) w'O (SI)])0- 13 W~o (SI) W'o ($2) NSI ds2, (79)

PV't ~--- - - 1 P,o + ~~ G(s )[V2v + 2 u,o __~T]d


v s
r o -~T

1 2 1 + ]
+~o~ G ( s ) (W'r ~:'0 + W'O ~, r )'r +----~'(W,O ~'0 )'0 +-'7(W' ~'0 W'O ~ ) ds
r r~ r- r ~r

1
+ J'O J'O~ Y (S1, $2) -1 [w, r (s,) w, o (s 2)],r +
r
~3- [W,o (S,)W,o (S~)],O

1 W'r (s2)W,o (s 1) } dSl Ms2"


+ -U (80)
203

Equations (79) and (80) can be combined and the pressure field may be
eliminated using a stream function 5Udefined as,

u (2) = V x 7-'.

The resulting equation has driving terms both time independent and time
dependent of type e i~ and e 2i~. This means that the secondary field has a time
averaged mean part ~ (r, O) and oscillatory components with frequencies co and
2oo. The time averaged field ~ is thus defined by

--i Ld-" ~-'r --" I o G ( s ) { [ V " L1]r,O - [ r V . tl]O,r}mdS ,

+ ~o IoT"(s,,s2){[V'(A, (s,) A,(s2)]r,O - [ r V .(A, (s~) A,(s2)]O,r }rods, (81)

where { }. refers to the mean, that is time independent terms in the expression
between { }, and the operator L has the following structure,

34 1 34 ~3 1 ~2 4 ~2 2 33 1 C
L = r ~-g + r3 ~5~04 ~t_ 2 c7r3 r cTr2 + r 3 302 r 2 drO02 + -rT ~dr" (82)

An approximate solution for ~ is obtained by defining,

~-/m "-- H2 (r, O)~b(r, O) + ~o, (83)

H(r, O)= R 2 - r 2 + ~elrn sinnO. (84)

H is the shape factor, q~(r,O) is an unknown function, and To is an arbitrary


constant. The structure of equation (83) is the simplest form that meets the
required conditions,

lC9~ cgk'
u [ - m [ =0, v [ - m [ = 0 , Wm =constant. (85)
cqD r 00 c~D~ cqD~ ~ cqD~ c~D~
204

Next, ~bis specified as,

~b(r, 0)- ~ r ~bj (r, 0). (86)


j=l

In this form, ~ becomes,

% (r, 0) = c, (R 2- r 2)2 ~b,(r, 0) + e~ [(R z- r 2)2 ~b2(r, 0)

+ 2(R 2- r 2) r n sin n 0 ~b~(r, 0)1 + 0 (o~3), (87)

where ~m has been expressed as a series in powers of the small parameter e,. To
0(~), the equation to be solved reads as,

0(~, ) 9L [(RZ-r 2) ~b~(r, 0)] = g,(r) sin n 0 + gz(r) cos n 0, (88)

where the known functions gl (r) and g2(r) are involved expressions in terms of r
and of the fluid constants. The solution of equation (88) can be found by
replicating the method of Section 4. It is convenient to develop the functions g/as
finite polynomials in r. This can be easily achieved since g, are continuous and
smooth functions and the applicable range is of the order of R. Given the
structure of equation (82), for any term of the form rmsin nO on the RHS of
equation (87), this equation provides a non-homogeneous solution of the form
Cm r .m+3 sin nO, the constant Cm being dependent on m and n. It also provides, in

general, homogeneous solutions of type,

r/r
c~r~cos (fl In -R'\cos n C2F sin (fl In rR)
sin n 0 )
cos n 0
(89)

where c,, c 2 are arbitrary constants, and a ~ ifl are complex powers of r, with
a_~O. The solution for ~ at 0 (c~), thus, reads

5u~1) = [Pl + clr~ cos(fl In R)] sin nO+ [~2 + c2 r~ cos ~ ln R)] cos n O. (90)
205

Only the cosine term in equation (89) has been used because subsequent steps
demand that the homogeneous solution should be non-zero at r = R. We note
that the functions Pi(r) are known polynomials of r. The function ~1 is
determined next by equating ~m') as defined in equations (90) and (87), where the
constants c~ and c2are found after putting r = R.
The expression for ~m2) follows from the equation at 0(e~),

L 7/~m2) =g3(r)+g4(r) s i n n 0 + gs(r)cosnO+ g6(r) sin 2n0 + g7(r) c o s 2 n 0 , (91)

in which all terms in the RHS have been rendered linear in sinjnO and cosjnO, j
= 1,2. The functions gt (i =3 .... 7) depend on r and on the material constants
embedded in the shear relaxation and the quadratic shear relaxation moduli G(s)
and y(s~,s2) which appear in equation (81). Equation (91) can be solved using a
similar technique as that applied in the first order problem. The functions g, are
approximated as finite polynomials of r. Computation of the homogeneous and
non-homogeneous solutions of equation (91) is straightforward. All arbitrary
constants are determined from the required spatial continuity for r = R, after
-equating ~m~) computed from equation (91), and its equivalent expression
obtained from equation (87).
Material constants involved in the driving terms in equation (81) were adopted
from Siginer in [2]. Plots of the streamlines for the same values of the fluid, flow
and geometry parameters used in Section 4 are shown in figures 7 and 8. A very
complex pattern of cells is common between the two cross-sectional shapes
chosen. The main cells appear a number of times equal to twice the number of
cross-sectional sides. Minor cells are present close to the corners, and to the
center of the pipe, depending on the value of k.
These results can be extended to other shapes by changing the values of n
and/or c~. Secondary flows appear for any value of ez ~ 0. They do not exist
only for c ~ - 0 , in which c a s e ~gCm ~ ~/0 =constant in the whole domain,
consequently u = v = 0.
206

Figure 7. Transversal streamlines for n = 3 ; e = 0.3485 ; k = 0.007.

Figure 8. Transversal streamlines for n = 4 ; e = 0.22; k = 0 . 0 0 7 .


207

ACKNOWLEDGEMENTS
This work was partially supported by grant 197-0810 of the Chilean National
Fund for Scientific and Technological Development. The authors acknowledge
important institutional support provided by the departments of Mechanical
Engineering at the University of Santiago of Chile, and the New Jersey Institute
of Technology.

REFERENCES
1. Siginer, D. A., J. Fluids and Structures, 6 (1992) 719.
2. Siginer, D. A., Int. J. Engng.Sci., 29 (1991) 1557.
3. Mueller, W., Z. Angew. Math. Mech., 16 (1936) 227.
4. Yih, C. S., Fluid Mechanics, McGraw-Hill, New York, NY 1969.
5. Fan, C., and Chao, B. T., Z. Angew. Math. Phys., 16 (1965) 351.
6. Hepworth, H. K., and Rice, W., J. Appl. Mech., 37 (1970) 861.
7. Letelier, M. F., and Leutheusser, H. J., J. Engng. Mech., 111 (1985) 768.
8. Letelier, M. F., Leutheusser, H. J., and M~rquez, G., J. Engng. Mech., 121
(1995) 1075.
9. Sexl, T., Z. Phys., 61 (1930) 349.
10. Uchida, S., Z. Agnew Math. Phys., 7 (1956) 413.
11. Oldroyd, J. G., Proc. R. Soc. London Ser. A, 283 (1965) 115.
12. Xue, S. C., Phan-Thien, N., and Tanner, R. I., J. Non-Newt. Fluid Mech.,
59 (1995) 191.
13. Rahaman, K. D. and Ramkissoon, H., J. Non-Newt. Fluid Mech., 57 (1995)
27.
14. Barnes, H. A., Townsend, P., and Waiters, K., Rheologica Acta, 10 (1971)
517.
15. Sundstrom, D. W., and Kaufman, A., Ind. Eng. Chem. Proc. Des. Dev., 16
(1977) 320.
16. Manero, O., and Waiters, K., Rheologica Acta, 19 (1980) 277.
17. Davies, J. M., Bhumiratana, S., and Bird, R. B., J. Non-Newt. Fluid Mech.,
3 (1977) 237.
18. Phan-Thien, N., J. Rheol., 25 (1981) 293.
19. Green, A. E., and Rivlin, R. S., Arch. Rat. Mech. Anal., 1 (1957) 1.
20. Letelier, M. F., and Siginer, D. A., Rheology and Fluid Mechanics of
Nonlinear Materials, S. G. Advani and D. A. Siginer (eds.), 121-128,
American Society of Mechanical Engineers, New York, NY 1997.
209

EFFECTS OF N O N - N E W T O N I A N FLUIDS ON
CAVITATION

D.H. Fruman

Groupe Ph~nom~nes d'Interface


Ecole Nationale Sup~rieure de TECHNIQUES A VANCEES
91761 Palaiseau Cedex- FRANCE

1. I N T R O D U C T I O N

The change of phase, from liquid to vapour, occurring when the temperature
is kept constant and the local pressure is decreased to become equal or less than
the vapour pressure of the liquid, is called cavitation. For this definition to
hold, it is necessary for the liquid to contain dispersed micro bubbles of non
condensable gas, called nuclei, offering interfaces where opportunities exist
for the change of phase. The liquid can be either stagnant or in movement. In
the first case, the pressure change is due to the propagation of pressure waves
generated by ultrasound generators, boundary vibrations, sudden valve
closures, underwater explosions, etc. In the second case, the pressure reduction
is associated, through the Bernoulli equation, with the increase of the liquid
velocity as it occurs in the flow around hydrofoils, and by extension in pumps,
turbines and propellers, in the contraction of pipes, in venturi tubes, etc. In
separated flows, such as jets and wakes, cavitation occurs in the low pressure
regions of the vortices developing in the free shear layers. Cavitation over
solid surfaces can occur in the form of either individual travelling bubbles or
sheets. The latter continuously shed bubbles which, when penetrating in a
region where the pressure has increased above the vapour pressure, will
collapse and rebound with an amplitude decaying with time as a result of,
essentially, viscous dissipation. The consequences of cavitation are the
reduction of the hydrodynamic performance of liquid machinery - pumps,
turbines and propellers - and hydraulic circuits, the emission of noise, the
generation of vibration and the erosion of materials.
Bubble cavitation occurrence is essentially determined by the vapour pressure
of the liquid. Surface tension opposes bubble growth and determines the level
of "tension" to which the fluid has to be subjected before the onset of cavita-
210

tion. Also, the viscosity of the liquid, assumed to behave as a Newtonian fluid,
will affect the bubble growth and contribute to the damping of the oscillations
following collapse. Even in the case of such liquids, characterized by the
simplest relation between strains and stresses, explaining and interpreting the
inception of cavitation and the consequences of its development on engineering
applications is not a simple task and numerous textbooks [1-4] have been
dedicated to consider these issues.
When the liquid displays a non-Newtonian behavior, as it often happens in the
chemical engineering processes and when polymer additives are dissolved in
Newtonian liquids, the problem becomes much more complex. It has to be
recognized that the larger impetus on the research concerning cavitation in
non-Newtonian fluids comes from the interest developed during the 1960's by
the use of dilute solutions of high molecular weight polymers, known as being
very effective turbulent drag reducers, in naval hydrodynamics and hydraulics
(see the review paper by Acosta and Parkin [5] for a summary of work on
cavitation in polymer solutions up to 1975). It was soon recognized that these
solutions were capable of modifying strongly the flow field on a length scale
characterizing the boundary layer for very small concentrations, even below
one ppm. As a consequence, the experimental investigations on the effect of
flowing drag reducing polymer solutions on cavitation were faced with the
following question, i.e. was the behavior of individual bubbles altered because
of:
i) the modification of the surface tension, the vapour pressure and the
rheological properties of the liquid phase, or,
ii) the changes of the local pressures, as a result of the modifications of the
flow structure and of the boundary layer.
In presenting the most remarkable results we will try to single out which of
these approaches is predominant.
The spherical growth and collapse of a gas bubble has interested rheologists
because it creates an extensional flow which can be used to estimate the
elongational viscosity of polymeric liquids. Although numerous theoretical and
experimental investigations have been conducted on this matter, we will only
talk about them occasionally because they belong, in our opinion, to the area of
characterization methods rather than to that of hydrodynamic cavitation.
In this presentation we consider first the theoretical approach of the dynamic
equilibrium of a single spherical bubble in an unbounded liquid. This is
followed by some theoretical results concerning the effect of Newtonian and
non-Newtonian properties on bubble growth and collapse. Because the
conclusion of this section is that these properties have a very limited impact on
the life of spherical bubbles, we move to the presentation of experimental
results obtained in the somehow academic conditions of a single bubble
interacting with solid walls or affected by shear flows. For these non-spherical
deformations it is shown that the viscoelastic properties have a significant
211

effect. We proceed next to review results obtained in situations such as


cavitation of submerged jets, on hemispherical bodies and hydrofoils, and in
vortices. We show that, it is usually difficult to decide whether the flow or the
bubble dynamics modifications are responsible of the observed changes on
cavitation behavior. However, it appears that the former are, generally,
predominant. We proceed by presenting recent results concerning the effect of
polymer solutions on the cavitation occurrence in very confined spaces.
Finally, we consider the important problem of cavitation erosion.

2. SINGLE SPHERICAL BUBBLE BEHAVIOR

2.1 The generalized Rayleigh-Plesset equation


Let us consider a single bubble growing spherically in an unbounded liquid.
In a spherical coordinate system (r, 0, ~), with the origin at the center of the
bubble, the only velocity component is radial and, in order to satisfy the
continuity equation, is given by,

u-
R(t))
r
/~(t) (1)

where R ( t ) is the bubble radius at time t, and [ ~ ( t ) = d R / d t . The radial


component of the momentum equation is given by,

OU OUI o~p f)Trr 2 Trr- TOO - Tr162


P -&-+UOrr - - - ~ - r + Or + r (2)

where p is the liquid density, p is the pressure anywhere in the liquid, and
Trr, TOO and 7:~r are the normal components of the deviatoric stress tensor
related by,

Trr + Tr162+ TOO = Trr + 2 r162 - 0 (3)

Because of (1) and (3), expression (2) becomes, after integration between
infinity and R,

p ( RR + -
2
- p(R)- Poo (t) + Trr (oo)-- Tr r ( R ) + 3
JR r
dr (4)

where p ( R ) and poo (t) are, respectively, the pressure at the interface on the
liquid side and the pressure far from the bubble. The stress Trr (cx~) can be
212

assumed to vanish far from the bubble. The pressure at the interface on the
liquid side is given by,

2~
p(R) - Pv + Pg R + Trr (R) (5)

where pg is the non condensible gas pressure within the bubble, p v the vapour
pressure and 7' the surface tension. By determining p g from the initial
conditions of the nucleus of radius R0 and gas pressure PgO, expression (5)
becomes,

P(R ) - Pv +
( P O - P v + -~O
27'/(R~/ 3k 2~+Trr(R )
--R R
(6)
where k is the polytropic index and P0 the pressure at infinity at t = 0.
Substituting (6) into (4) we obtain the generalized Rayleigh-Plesset equation

3
) - - - - ( P o o ( t ) - - P v ) + ( P o - p v ) ( - R ~ ) 3k
2
(7)

R0 R
.01
R
+3
R r
--dr

allowing one to compute the time evolution of the radius of the bubble as a
function of p~ (t) under the assumption that the liquid is incompressible and
that thermal and mass transfer effects are negligible. The terms on the left
hand side represent the contribution of inertia, while, on the right, the first
term is the cause of the bubble growth, the second term the contribution of the
partial gas pressure, the third term the effect of surface tension and the last
term that of the rheological properties of the liquid.
We would like to point out here that, whatever the rheological behavior of
the liquid, for cavitation to exist, the pressure far from the bubble has to be
reduced to a value at least equal to a critical pressure for which an asymptotic
growth phase of the bubble, at constant velocity, will exist. In the case of the
static equilibrium of a bubble undergoing an isothermal expansion and
ignoring the contribution of the last term in equation (7), the critical pressure
is given by,

4 ?'
Pc - P ~ (8)
3 Rc
213

with,

Re-Ro
E3R0/ 2
-~ --~- P O - P v + -~o (9)

The order of magnitude of the asymptotic growth velocity when the bubble is
subjected to p~ < Pv is given by,

[~g =I3Pv--PC~p (10)

The characteristic time for a vapour bubble, having an initial radius Rch, to
collapse under the effect of a pressure step is,
[

~R = O. 915 Rch ~[ P (11)


VP~ -Pv
called the Rayleigh time. The velocity at the final stages of the collapse is given
by,

/~c= -3 p --R-- (12)

where p~ > Pv. It is clear that the velocity at the interface during collapse can
reach values much higher than those of the velocity for asymptotic growth.
If the pressure is reduced but remains above the critical value obtained using
expressions (8) and (9) and a given radius of the nucleus, it will not achieve
the conditions for asymptotic growth and will oscillate before reaching its
equilibrium. Therefore, depending on the nuclei size distribution and the local
pressure, some bubbles will enter the asymptotic growth phase and cavitate and
others will only oscillate with limited amplitude.

2.1.1 Newtonian fluid


In the case of a Newtonian liquid, equation (7) reduces to,

3 ~'rr d r - - 4 1 ~ - - (13)
R r R
214

where ~t is the viscosity of the liquid. Numerical integrations of equation (13)


have been conducted for numerous cases of time dependent reference
pressures. As an example, Figure l a shows the radius of a bubble having a
10 ~tm initial radius as a function of time for a steep decrease of pressure,
from 101 000 Pa to 100 Pa, constant pressure during 0.3 ms, and a steep
increase thereafter. The initial slow growth of the bubble corresponds to the
non cavitation region before the critical radius (= 4.2 10 -5 m) is reached.
Comparison of the inviscid case (/~ = 0) to the water case (Ft = 10 -3 mPa. s)
shows that the effect of the viscosity is relatively small and results in a slight
decrease of the maximum radius. This decrease becomes larger if the viscosity
increases to five times that of water. If the results are plotted in non
dimensional form, using as scales the maximum radius and the Rayleigh time
computed with the maximum radius, the results for the collapse can be hardly
distinguished whatever the viscosity, Figure lb. The difference in the dimen-
sionless time for the growth phase is associated to the modification of the
Rayleigh time of each one of the bubbles.
The straight discontinuous line has a slope corresponding to the asymptotic
growth velocity, computed using expression (10) and pv = 2400 Pa. This
velocity is larger than the one achieved by the bubbles at the end of the growth
phase. This effect is essentially associated with the initial diameter of the
nuclei; for smaller diameters, the asymptotic velocity is reached more
rapidly, leading to bubbles whose final diameters will not be too dependent on
the initial ones. For example, a tenfold decrease of the initial radius, from 10
to 1 ~tm, will, under the same pressure versus time conditions, lead to a
maximum radius of the bubble only 60% smaller.

2.1.2 Non-Newtonian purely viscous fluid


By ignoring the probable contribution of elasticity, a power-law fluid has
been adopted by Shima and Tsujino [6] as a model of carboxymethylcellulose
(CMC) aqueous polymer solutions. The stress term in Equation (7) is now
written in the form,

n-1
f S Zr__Zrdr = _4 ( 2 )n-1 m
r n R

where m and n are the power law parameters and 77(R,/~, m, n) the apparent
viscosity. They numerically investigate the collapse and rebound of bubbles of
different diameters to a sudden increase of the pressure far from the bubble
and do not consider the bubble growth phase. Their results do not allow one to
show that power law behavior is responsible for the acceleration of the
damping of the oscillations, as compared to those in water.
215

0.35 [- R, m m p, Pa
0.30 ~Pa.s /
0.25 t- 0 //~'/ \//
I 10 -3 ///" X \~ 101000
0.20 -3 /

o.15 /~
0.10

0.05 ~ f / lO0
ms
0 t t t
0 0.1 0.2 0.3 0.4

1.0
_R/Rma 1 b)
0.8
Pa.s

0.6 0
10-3
5x 10 -3
0.4

0.2
t-t(R max )

0 I rR
- 12 - 10 -8 -6 -4 -2 0 2

Figure 1 : a) Pressure and radius of the bubble for an inviscid fluid (~t = 0),
water (~t = 10-3 Pa.s) and a more viscous fluid (r = 5 10-3 Pa.s) as a
function of time, b) radius normalized with the maximum radius and time
normalized with the Rayleigh time (expression (11)) for the maximum radius.
The straight discontinuous line has a slope corresponding to the asymptotic
growth velocity, expression (10).
216

Table 1 9 Fluid properties and main characteristics of the radius versus time
evolution in Figure 2.
rn n R ( t = 0.3 ms) /~(t = 0.3 ms) tmax R A
mPa.s n mm m/s ms mm %o
0.895 1 0.3228 1.053 0.3332 0.1861
10 0.8 0.3203 1.047 0.3330 0.1848 4.24
50 0.65 0.3190 1.041 0.3328 0.1841 6.14
450 0.4 0.3192 1.037 0.3328 0.1847 4.48

0.35 R, m m

0.30

0.25

O.2O

0.15

0.10

0.05 r n = 0.4
I
0 I //jl I I I
/
0 0.1 0.2 0.3 0 0.1 0.2 0.3 0 0.1 0.2 0.3 t, m s

Figure 2 : Radius of the bubble as a function of time resulting from the


pressure evolution shown in Figure 1 and a nucleus radius of 10 ~tm f o r : a)
N e w t o n i a n fluid /1 = 0 . 8 9 5 m P a . s ; b) p o w e r law f l u i d with
m = 10 m P a . s 0.8 and n = 0.8 ; c) power law fluid with m = 450 m P a . s 0.4
and n = 0.4. No effect due to the rheothinning property of the fluid can be
detected.

We have conducted the integration of the generalized Rayleigh-Plesset


equation with the integral term of equation (14) and the pressure variation
shown in Figure 1 for a Newtonian fluid of ~t = 0.895 mPa.s, and three
power law fluids whose characteristics are given in Table 1. These characte-
ristics are such that the evolution of the radius of the bubble with time is
practically indistinguishable from the reference Newtonian situation, as shown
in Figure 2 for two of the three power law fluids. Indeed, Table 1 shows that,
as compared to the Newtonian case : the radius achieved at the end of the low
pressure plateau (third column) differs, at the most, by 1.2 p e r c e n t ; the
217

interface velocity at the same time (fourth column) differs by a maximum of


1.5 percent; the total duration of the bubble life (fifth column), which ends
when the interface velocity reaches the sound velocity in the liquid, differs by
a maximum of 0.1 percent; and the mean bubble radius (sixth column) and
the differences of the mean bubble radius relative to the maximum radius in
the Newtonian fluid, A, differ by less than one percent. In spite of the
considerable differences of the instantaneous apparent viscosity, 77, the
behavior of the bubble remains the same as that of an equivalent Newtonian
fluid and vice versa.

2.1.3 Non-Newtonian viscoelastic fluid


The time history of spherical bubbles in non-Newtonian viscoelastic fluids
has received considerable attention since the earlier works of Fogler and
Goddard [7, 8] for a fluid model including stress accumulation with fading
memory. Later, Ellis and Ting [9] have employed a second order fluid and
Yang and Lawson [10], Chahine [11] and Ting [12] an Oldroyd fluid. The
Fogler and Goddard results presented large elastic effects, while all other
authors conclude that viscoelasticity has a very limited retardation effect
(Ting, [12]) on bubble growth and collapse, provided the material constants
are compatible with dilute polymer solutions properties.
As stated in the paper by Kezios and Schowalter [13], when equation (7) "is
solved for conditions appropriate for cavitation .... the dynamics are dominated
by inertia, and the contribution of rheology to the integral on the right-hand
side is of little consequence. This is fortunate, in that one need not be
concerned with fluid viscosity, let alone elasticity. But it is also inconclusive,
because it leaves open the question of interactions between bubbles and the fact
that in a shear field one does not expect the bubbles to be spherical".
These findings have been further substantiated theoretically by a recent and
very well documented paper by Riskin [14]. By incorporated the polymer-
induced stress calculated using a "yo-yo model" which accounts for the
unravelling of the polymer molecules into equation (7) without the non-
condensible gas contribution, the author computes the growth and collapse
phase of a vapour bubble. He concludes that the growth of the bubble is not
affected by the polymer, but the final stage of collapse is. He shows that, there
is a total arrest of the collapse, with a bubble wall velocity reduced to nearly
zero when the radius becomes about or less than 10% of the radius at the
initiation of the collapse. However, under such situations, the velocity of the
wall prior the sudden onset of the viscoelastic effects will be, because of
expression (12), if (P~-pv) = 10 5 Pa and p = 10 3 k g / m 3, = 815 m/s. This
is large enough to invalidate the assumption of incompressibility for both the
liquid and the gas phases.
Therefore, none of the theoretical approaches of spherical bubble deforma-
tion in purely viscous and viscoelastic fluids indicate the possibility of strong
218

changes on the bubble behavior. Thus, we next consider results concerning


experimental single bubble behavior for a variety of conditions to establish
under what circumstances the non-Newtonian behavior becomes significant.

3. S I N G L E B U B B L E 9 E X P E R I M E N T A L

3.1 Stagnant fluid


The results of numerous experiments conducted to investigate the behavior of
spark generated bubbles in an unbounded f l u i d - water, polyethylene oxide
(PEO) and Guar Gum solutions - were presented in a paper by Ellis and Ting
[9]. The authors showed that the polymer additives, even at concentrations as
high as 1000 ppm of PEO, for which the aqueous solution display very
marked viscoelastic effects, did not affect the behavior of the bubbles in any
significant way, that they remain spherical during growth and collapse and that
the duration of the collapse phase is equal to the Rayleigh time (equation (11))
computed using the maximum radius of the bubble.

1.2
- Rcma x RA
_ A AAA 10
1.0
o ~ o 9 I
0.8
- o~ A ~ AA I
0.6 - ~ 0 9
0 Q
0.4
0 9 A
0.2 17= Rc max / L
0~
o 77 Fluid
o |zx 9 1.39 250ppmPEO
o o 1.25 Water
I I I I I
9 0.56 250ppmPEO
0.5 1.0 1.5 2.0 t /"t" R
A 0.50 Water

Figure 3 : Behavior of spark generated bubbles in water and a polymer


solution in the vicinity of a solid wall. The distance between the center of the
initial bubble and the wall is L, RA is the distance between the position of the
interface opposite to the wall and the center at the origin, Rc is the lateral
extension of the bubble, t is the time and z is the Rayleigh time computed for
Rcmax. From Chahine and Fruman [15].
219

These results were confirmed by Chahine and Fruman [15] using the same
bubble production technique and distilled water and a 250 ppm solution of
PEO (Polyox WSR 301) as testing fluids. They also show that the time and
amplitude of the first and second rebounds were unaffected by the polymer
additive. Very recently, Brujan et al. [16] produced laser-induced cavitation
bubbles in solutions of CMC and polyacrylamide (PAM) with a concentration
twenty times larger than the one used by Fruman and Chahine. Even for this
very large concentration and for the very viscoelastic PAM solution, the
bubble behavior is "little affected by the rheology of the fluid and the most
important parameter for bubble oscillation is the infinite-shear viscosity".
Since it was substantiated that a re-entering jet developed if the bubble collap-
sed close enough to a solid wall, and hence the viscoelastic properties of the
polymer solution might affect the development of the perturbation initiating
the re-entering jet, Chahine and Fruman [15] also investigated this
configuration. Typical results are shown in Figure 3, and demonstrate that
when the distance to the wall decreases as compared to the maximum lateral
extension of the bubble, the polymer solution introduces a retardation effect
over the initiation of the re-entering jet. Notice in this figure that the position
of the interface is, at the end of the collapse, below the center of the bubble at
origin. This investigation was, we believe, the first to demonstrate that
polymer solutions seem to have a stabilizing effect on bubble dynamics,
especially with regard to deviation from sphericity. Brujan et al. [16] did also
experiments in the vicinity of a solid wall and confirmed these early results.
Chahine and Morine [17] and Morine [18] conducted tests with spark
generated bubbles confined between two parallel walls. This geometry offers a
rather peculiar situation of collapse because the bubble, initially spherical,
develops an hour-glass shape during the collapse phase and splits into two
bubbles which collapse against each wall. In this situation the departure from
sphericity occurs along the mid-plane between the two parallel walls. Figure 4
shows some of the results obtained with concentrations of 125 and 250 ppm of
PEO. Clearly, the lengthening effect on the bubble life duration as a result of
the presence of the walls is significantly reduced in the case of the polymer
solutions. The most striking is that, in these fluids, the departure from
sphericity of the bubbles is considerably delayed. Interestingly enough, results
obtained with a 50% glycerin/water mixture, having a viscosity about ten times
larger than water, demonstrate little change compared with those obtained in
pure water, in agreement with the results obtained by integration of the
Rayleigh-Plesset equation.
Chahine [19] published results concerning the behavior of bubbles in the
vicinity of a liquid-air interface and showed that, in this peculiar situation, the
deformation of the bubble in the polymer solution is enhanced and made more
non-spherical. This peculiarity can be associated to the fact that the free
surface also deforms considerably during the growing phase, forming a jet
220

which ascends vertically up to distances of many times the maximum bubble


diameter. Chahine [19] pointed out that, in the polymer solution, the jet is
axisymmetric, smoother and thinner and its surface looks stretched at the
beginning. "Later, the jet becomes distorted and loses its symmetry while in
water the axial symmetry is conserved".

1.2 m

CD I CD m a x
m a mcx nyi_n nl m
1.0 pOx. -o-o 9 []
m A I k v

_ AOAA O zx A []
0.8
AolLX o A D xe i
0.6 - 0 ~ 0 A
X @
o [] | m
0.4
OA x @
0 A [] Oi
0.2~
0 x
EX @ []
0 I I I I I I
0 0.5 1.0 1.5 2.0 2.5 3.0 t/TR
~N"~"N'~"~N'~NNN'~'N~ ]7 fluid fl fluid

I b 9
9
9

0.59
0.64
0.70
0.54
water
water
water
water +glycerin
o
A
[]
0.56
0.61
0.80
250 p p m
250 p p m
250 p p m

Figure 4 : Behavior of spark generated bubbles in water and a polymer


solution in between two parallel walls. The center of the initial bubble is in the
middle of the gap, the lateral extent, CD, of the bubble is plotted divided by its
maximum value, t is the time and "t" is the Rayleigh time computed for CDmax.
=CDmax/b. From Chahine and Morine [17] and Morine [18].

3.2 Shear flows


A detailed investigation of the effect of a controlled shear flow on the
deformation of laser-generated bubbles was conducted by Kezios [20] and
Kezios and Schowalter [13]. Because of the shear, the bubble is elongated, as
indicated in the insert of Figure 5, and as time proceeds it reaches a maximum
deformation amplitude and then recedes during the collapse phase. In Figure 5,
results obtained with polyacrylamide (PAM) and PEO solutions are shown.
221

100
-~ ,~m 0
0
80
0 A
0
Fluid 60 0 A
A
Water O O DOn [] A
500 ppm PAM 40
6o [] 0
I000 ppm PAM o
500 ppm PEO A
20 []
1000 ppm PEO*
O r7
0 I I I
0 0.5 1.0 1.5 t/'C R
100
- ~ ,~m 0
R 3=ab 2 0
0
A =a-R 80 0
ml
m m o
5 60
oOm
o m
AA
A
IIO
o [] A A l
0 0 mll A
40
@ A A

20- B
o
A
I I I I
0 0.5 1.0 1.5 t / Z"R

Figure 5 : Mean bubble deformation amplitude (see insert) as a function of


time divided by the Rayleigh time corresponding to the maximum equivalent
spherical radius for water and polymer solutions and a shear rate, y, of
170 s-1. The asterisk refers to a low molecular weight PEO. From Kezios and
Schowalter [ 13].

They demonstrate that the departure from sphericity is significantly reduced


in polymer solutions, in particular in PAM solutions. Interestingly enough, the
authors found that increasing the concentration beyond a critical value reverses
the results and they speculated that this can be caused by the relative increase
of the solution viscosity as compared to its elasticity. Needless to say that these
authors have consistently checked that, whatever the polymer and the polymer
222

concentrations, the bubbles remain spherical during all the process of growth
and collapse when the shear rate was absent.
Ligneul [21] performed experiments with spark-generated bubbles immersed
in the shear layer developed by a rotating cylinder, with velocities between 1
and 5 m/s, inside a reservoir of much larger dimensions. By comparing the
behavior in water and solutions of PEO with 50 and 250 ppm concentration,
he concludes that the influence of the polymer solution is to maintain
sphericity during bubble collapse (but not during bubble growth).

4. C A V I T A T I O N IN A M O V I N G F L U I D

4.1 Jet cavitation


The cavitation number is defined by,
a - Pr - Pv (15)
1

with P r the pressure in the reservoir where the jet is discharging and V j the
bulk velocity of the jet. Depending on the way the experiments are performed,
either by decreasing the pressure from a non r situation until
cavitation is reached or by increasing the pressure from a r situation
until the non cavitating conditions are achieved, the inception, a i, or desinent,
a d , cavitation numbers are obtained respectively.
Hoyt [22-25] conducted experiments with a nozzle situated 60 r below the
free surface of an open reservoir. Preliminary test results [22] showed a sharp
decrease of the incipient cavitation number when the concentration of the
homogeneous polymer (PEO) solutions was increased up to 10 p p m ;
afterwards the reduction was not as rapid. These results were later confirmed
and extended by modifying the upstream turbulence level [25]. All the data is
plotted in Figure 6.
For these concentrations, the fluid behaves as Newtonian with an increase of
the viscosity of about 10 and 25%, as compared to the solvent, for respectively
10 and 100 ppm. The surface tension of the polymer solutions was reduced
very sharply, by as much as 10%, when the concentration increased from zero
to about 40 ppm and reached a plateau afterwards. The tensile strength, the
nuclei concentration and the air content, at saturation or very slightly below,
are practically unaffected in the polymer solutions for the concentrations
mentioned earlier. Because the surface tension is reduced we may expect,
based on equations (8) and (9), an increase of the critical pressure and an
advanced cavitation. Since all other properties are unchanged and the
cavitation inception is delayed, Hoyt concludes that the causes are of an
hydrodynamic origin.
223

~f
0.4
Inception cavitation index

) Turb. Turb.
generator level, %
0.3
No <2
nX Yes 2
0.2 Yes 2.2
Yes 2.6
0.1 Hoyt [21]
Polymer concentration, ppm
I I I I I I I I
0 10 20 30 40 50 60 70 80

Figure 6 : Incipient cavitation number as a function of polymer concentration


for a submerged jet issued from a 6.35 mm diameter nozzle at four conditions
of the upstream turbulence level. Reynolds number were in the range of
2 105. From Hoyt [25].

Baker et al. [26] obtained the desinent cavitation numbers, o"d, instead of o"i
as in the case of Hoyt, for a constant concentration, 100 ppm, of the same
polymer, a nozzle diameter of 50.8 ram, and a Reynolds of 6 x 105, larger
than in the case of Hoyt's tests, for varying concentration of air in the
circulating fluids. When the concentration was below saturation (<4 ppm) the
polymer did not introduce any appreciable change, but, for supersaturated
conditions (>4 ppm) cavitation desinence was delayed. The considerable
difference in behavior between Baker's and Hoyt's results can be due to the
much larger size of the nozzle, by near a factor of ten, the reduction of the jet
velocity, by a factor of about three, and the choice of the desinent instead of
the incipient cavitation number as a reference. Indeed, even in the case of pure
water, increase of the total gas content causes a strong increase of c7d, in
particular for supersaturated conditions.
Hoyt and Taylor [27] conducted visualizations of cavitating events using a
nozzle of only 2.92 mm diameter for saturated and undersaturated conditions.
The photographs clearly demonstrated that, for cavitation numbers comparable
or lower than those for which the water jets show extensive cavitation in the
shear layer, the drag-reducing polymer solutions, with a concentration of only
25 ppm, drastically reduced the presence of individual cavities. Moreover, the
bubbles are larger and their surface is smoother than for the conditions of
limited cavitation in water jets.
224

1.4
- Desinent cavitation index
1.2
AA

1.0 -

e# Fluid
Jet
origine
0
0.8 - o o o ~ 0 0
Water Nozzle
0.6
4 Water Orifice
100 ppm PEO Nozzle
0.4 100 ppm PEO Orifice

0.2
Total gas content, ppm
0 I I I I I I I
0 2 4 6 8 10 12 14

Figure 7 9 Desinent cavitation number as a function of total gas content for


submerged jets of water and an aqueous solution of 100 ppm of poly(ethylene
oxide) issued from a 50.8 mm diameter nozzle and orifice at Reynolds
number of 6 105. From Baker et al. [26].

Arndt [28] advanced an explanation of the Baker et al. [26] and Hoyt [22-25]
results based on the difference in the hydrodynamic behavior of jets depending
on the values of the Reynolds numbers. His analysis is based on the fact that,
for equal Reynolds number, the order of magnitude of the strain rate in the
contraction of the small nozzles, will be larger than for the large nozzle. Thus,
this effect, coupled to the transitional state of the jets in the Hoyt and Taylor
[27] experiments, is expected to lead to a significant viscoelastic effect on the
pressure field and to justify the inhibition of cavitation. However, no
explanation is given on why cavitation is delayed in the Baker et al. [26]
experiments in the supersaturated regime.
Already in 1974, Ting [29] correlated jet cavitation inhibition data by Hoyt
[23] with the drag-reducing properties of the polymer solutions. He showed
that the percent reduction of incipient cavitation and of drag reduction can be
expressed in the same way; as a function of the polymer concentration and an
intrinsic concentration, which is a measure of how rapidly cavitation surpres-
sion (drag reduction) increases with increasing concentration. The
characteristic numerical values obtained in both situations are in very good
agreement.
In conclusion, jet experiments do not show conclusively that cavitation is
inhibited under all flow conditions. It appears rather clearly, however, that
modification of the flow structure is the main effect with little, if any, contri-
bution from the fluid properties on inception. When cavitation is well develo-
225

ped, the cavities are larger than those existing in the solvent and their surface
is smoother. These modifications can be directly related to experiments
showing that polymer solutions delay the departure from sphericity of single
bubbles and, as a consequence, their eventual break-up due to surface defor-
mation and instability.

4.2 C a v i t a t i o n near blunt bodies


The effects of polymer solutions on cavitation inception on blunt bodies
(ogives) was first investigated by Hoyt [30]. For axisymmetric bodies, many
other papers followed; Ellis et al. [31] and Baker et al. [32], in homogeneous
solutions, and van der Meulen [33, 34] and Gates and Acosta [35], for injected
polymer solutions. Reitzer et al. [36] conducted experiments using a circular
cylinder. As an example, Figure 8 reproduces the inception cavitation numbers
obtained by Gates and Acosta [35] for a hemispherical nosed body as a function
of the Reynolds number for various values of the non dimensional injection
rate of polymer fluid, defined as,

G- cQ (16)
rc D U Sts

where c is the concentration, Q the volume flow rate of the ejected solution, D
the diameter of the hemisphere, U the free stream velocity and 8ts the
boundary layer displacement thickness calculated at specified positions along
the body surface.
There is a marked reduction over the whole range of Reynolds numbers, but
the relative effect seems to decrease when the Reynolds number increases.
Moreover, when the concentration of the ejected fluid is increased, the
polymer effects increase also but seem to reach a saturation beyond which
further increase causes a reduction of the polymer efficiency. The saturation
effect is exemplified in Figure 9 where the cavitation inception number has
been plotted for a constant Reynolds number of 7.5 x 105 as a function of the
polymer solution injection rate (up) and the dimensionless polymer solution
injection rate (down). It is interesting to note that what seems to be really
important is the amount of polymer present in the reference thickness of the
boundary layer.
Gates and Acosta [35] have plotted the maximum percent reduction in
inception on the hemisphere nose as a function of the Reynolds number for
their results and those previously published (Figure 10). It is very interesting
to see that the effects are qualitatively comparable in spite of the differences in
the way the polymer molecules are brought to the body wall (carried by the
main flow or ejected at the nose). The differences, between the results of van
226

der Meulen [34] and Gates and Acosta [35], may be due to many factors and, in
particular, the level of turbulence and nuclei concentration of the cavitation
tunnels employed in their investigations.

0.7 - cyi

~ Hemisphere body

0.6 9 : 106x G

iBm i ~ o 0
[] 0.35
0.5
ii 1.75
. / Q 2.3

0.4 . o/.--. A
A
7
9
t 14.6
K/ 10-Sx Re
23.6
0.3 I I I I I I I
3 4 5 6 7 8 9
0.7
(3"i
NSRDC body

0.6 106 G

o 0
O 0.35
0.5
(9

0.4 O
|
o
1 0 - S x Re
0.3 I I I I I I I I I
3 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4

Figure 8 9 Inception cavitation numbers for a hemispherical and a NSRDC


body as a function of the Reynolds number for various values of the non-
dimensional injection rate of polymer fluid. From Gates and Acosta [35].
227

0.65 F O'i
D O
50 ppm
0.60 500 ppm
O

0.55 -0

- |
0.50

0.45
O Q, 10-6 m3/s
I I I I I I I I I I I I I I
0.40
0 5 10 15

0.65 ~- O"i
O 50 ppm
0.60 OO O 500 ppm
0.55

0.50

0.45
| 106 G
0.40 I I I I I I I I
0 1 2 3 4 5 6 7 8

Figure 9 9 Cavitation inception number for a constant Reynolds number of


7.5 as a function of the polymer solution injection rate (up) and the
dimensionless polymer solution injection rate (down). From Gates and Acosta
[35].

In an attempt to interpret these results, van der Meulen [37] and Gates and
Acosta [35] visualized the boundary layer along the body surface in the region
of separation. They both come to the same conclusion: the polymer additives
remove "the laminar separation by stimulation of transition (Gates and Acosta
[35])", causing "transition to turbulence at much lower Reynolds numbers than
the pure solvent (van der Meulen [37])". This conclusion is further
substantiated by the results of tests performed by these authors with a Schiebe
body, which, theoretically, should not exhibit laminar boundary layer
separation. They showed that, indeed, the polymer solution ejection has no
effect whatsoever on either the type of cavitation or the inception index.
Moreover, as signalled by Gates and Acosta [35], tripping the boundary layer
transition on the hemispherical headform has been demonstrated to be more
228

effective than the polymer solutions in delaying cavitation occurrence. It seems


therefore that, for these tests and for the Reynolds numbers employed, the
polymers act by modifying the behavior of the boundary layer and have little,
if any, effect on the individual bubbles at inception. We will see below that
even if the cavitation is very much developed, these boundary layer effects are
still of importance.

60 - A o ' i ,o~
O"i
50 - A
conc.,
ppm Reference
40 500 van der Meulen [33]
o

30
O0
0 | o
@
@
@
[] A 2o} Ellis et al. [36]
A 0 o A 50
0 0 20 H o l l et al.
20-
0 0 0 @ @ 500 Gates & Acosta [34]
0 0 @ [] [] 2o } Baker et al. [31]
10- [] 80
10-SRe
0 1 I I I I I I I I I
0 1 2 3 4 5 6 7 8 9 10

Figure 10 9 Composite plot of maximum percent reduction in inception on the


hemisphere nose as a function of the Reynolds number. From Gates and Acosta
[35].

Brennen [38] investigated the effects of homogeneous dilute polymer


solutions on the characteristics and appearance of the interface of well
developed cavities produced behind a cylinder and spheres operating at
Reynolds numbers, relative to the diameter of the obstacles, comprised
between 1 0 4 and 1.2 105. The author showed that the polymers caused a
wavy instability of the wetted surface flow around the headforms at the
initiation of the cavity. This instability can be related to the effect of the
polymers on the transition and separation mentioned above.
Bazin et al. [39] presented one of the earlier results on the effect of the
injection of rather concentrated (5000 and 10000 ppm) solutions of PEO
WSR 301 on the surface of a cylinder downstream of the stagnation point.
They found, for unspecified Reynolds numbers, that the injection of the
polymers inhibits the development of cavitation and that the noise level at
inception conditions is significantly reduced by the polymer additive,
especially in the high frequency band (> 25 kHz).
229

14 l O - 3 x Pac ' P a

12

10 c = 0 ppm
c = 3 ppm
m

- (

zone zone one


V

- }~ 10-SxRe
o~ t
0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4
I I I I I I I I I I (7
15 10 9 8 7 6 5 4 3 2

Figure 11 : Acoustic pressure as a function of the cavitation number and of the


Reynolds number for water and a 3 ppm PEO homogeneous polymer solution
flowing around a circular cylinder. From Reitzer et al. [36].

Reitzer et al. [36] have also investigated the flow around a cylinder in an
open loop cavitation tunnel. A 1000 ppm PEO solution was ejected upstream
of the test section with a flow rate such that the concentration of the homoge-
neous solution in the test section reached an homogeneous concentration of
3 ppm after mixing. Their main results are summarized in Figure 11, where
the acoustic pressure is plotted as a function of the cavitation and Reynolds
numbers for water and for the homogeneous polymer solution. For water,
there is a region, zone I, where the flow is non cavitating, up to a Reynolds
number, defined with the free stream velocity and the diameter of the
cylinder, of 85 000, for which cavitation is initiated and isolated bubbles can
be seen. Beyond this Reynolds number, cavitation develops and the acoustic
pressure increases up to a maximum (zone II), after which it decreases over a
limited range of Reynolds (zone III). After a relative minimum, the noise
increases very sharply, reaches another maximum much larger than the
previous one (zone IV) and then decreases as sharply as it had increased (zone
V). By comparison, the polymer solution completely inhibits zone II, which is
replaced by the continuation of zone I, and zone III, where the noise was
decreasing in the case of water, is replaced by an increasing region that fits
perfectly zone IV. The polymer has no effect whatsoever on zones IV and V
where cavitation occurs respectively in the form of alternating bursts and an
elongated vapour cavity. From these and other data, the authors conclude, as in
other previous works, that the cavitation retardation associated with the
230

presence of the polymer molecules is more likely due to the modification of


the flow field around the cylinder than an effect of the macromolecules on the
dynamics of individual bubbles.

0.65 - Inception parameter


00 9
0.60 9 9 9
t o~
9
.. PEO-FRA
oemo o o o 0 ppm
O 9 A 9
0.55 -O

mm []
9 Am~
~pr Ipr V
v~mvvV []
A
100 ppm
250 ppm
0.50
m m V 500 ppm
0.45 [] A Ak A A y y y v
A A
V" 9 IPr
0.40
1 0 - S x Re
0.35 I I I I I I I !
0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4

0.65 - Inception parameter


O0 9
0.60 - 9 9
9
i~ 9 PAM-273
-o e ~ 9 emo oo 9 nnn 0 0 ppm
0.55 mum
[] 100 ppm
[][] mmmam m 9 At A 300 ppm
0.50 mum
~?V V T 500 ppm
0.45 _ A
k V

0.40
10-5 Re
0.35 I I I I I I I I
0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4

Figure 12 9 Inception cavitation numbers as a function of the Reynolds


numbers, computed using the local azimuthal velocity and the diameter of the
protrusion, for pure water and three concentrations of PEO and PAM. From
Ting [40].

The experiments with cylinders can be compared to those conducted by Ting


[40] in a specially designed unit allowing him to perform experiments using
very small amounts of test fluids. It consisted of a circular plate, on the surface
of which were two protrusions having a diameter and height of 2.9 mm,
rotating inside a casing. The protrusions were observed through a transparent
wall in order to determine, visually, the initiation and the extent of the
231

cavitation. The inception cavitation numbers as a function of the Reynolds


numbers, computed using the local azimuthal velocity and the diameter of the
protrusion, for pure water and three concentrations of PEO and PAM, are
plotted in Figure 12. For the largest concentration, PAM seems to be more
effective over a larger range of Reynolds numbers as a result, probably, of its
larger capacity to sustain high shear rates with moderate degradation. Ting
also signalled that the appearance of the polymer cavities was more transparent
than in water and showed a regular and smooth wavy pattern, as beautifully
demonstrated by Brennen's [38] visualizations.

4.3 Vortex cavitation


It is well known that viscoelastic fluids cause very strong effects in rotating
flows. In particular, while in a Newtonian (or even inviscid) fluid the rotation
of a rod makes that the surface of the liquid is depressed near the rod, in a
polymeric fluid the liquid will climb the rod. This is the famous Weissenberg
effect. The non-Newtonian liquid develops a normal stress contribution which
acts in such a way that the depression induced by the rotation of the liquid is
overcome, and even exceeded. In fact, the so called "vortex inhibition
phenomena" has been used as a method to estimate quantitatively and
qualitatively the viscoelastic properties of liquids (Gordon and Balakrishnan
[41]). It seems therefore natural to assume that the contribution of normal
stresses may have a significant effect on the cavitation which may occur in
confined (vortex chambers) or unconfined (tip vortices) situations.
For an axisymmetric line vortex, the radial component of the momentum
equation, assuming that the radial component of the velocity is zero, is given
by,

= OTrr Trr -- TOO


(17)
Or R Or r

where V o is the tangential velocity, and Tii the extra stresses (i = r, 0). The
pressure on the axis of the vortex is obtained by integrating equation (17),
from infinity to r = 0. It is easy to see that the contribution of the rotation,
whatever the velocity distribution, is to decrease the pressure on the vortex
axis. In a Newtonian fluid the extra stress terms cancel. In non-Newtonian
viscoelastic fluids, the experience shows, as stated above, that these terms
contribute to diminish the centrifugal effect.
Inge and Bark [42] conducted experiments with a specially designed elliptical
wing having a maximum chord of 0.16 m and a half span of 0.238 m. They
reported results obtained by ejecting concentrated polymer solution into the
water stream through a multiperforated injector situated one meter upstream
of the foil and for homogeneous polymer solutions with concentrations
232

between 0.01 and 12 ppm of PEO Union Carbide WSR-301. In Figure 13, the
values of a modified cavitation number, taking due account of the different
values of the incidence angle and of the free stream velocity, are plotted as a
function of the polymer concentration in the homogeneous solutions for three
conditions of cavitation development. The results show very clearly that the
incipient cavitation occurrence is significantly delayed for concentrations
larger than 1 ppm. The same is apparent for the other conditions of cavitation
although the changes are proportionally smaller. Interestingly enough, the
permanent cavity and the attached cavity situations cannot be distinguished for
the largest concentration. The experiments with the polymer ejection gave
qualitatively similar results.

o"n
O O Incipient cavitation
[] Appearance o f a p e r m a n e n t cavity
o
0 ~ ZX Cavity reaches the w i n g

[] o ....
El

c, ppm
0 I I I ] ,,,

0.01 0.1 1 10

Figure 13 9 Values of a modified cavitation number, taking into account the


effect of the different values of the incidence angle and of the free stream
velocity, as a function of the polymer concentration for three conditions of
cavitation development. From Inge and Bark [42].

Fruman [43] investigated the behavior of the tip vortex of a NACA 16020
cross section elliptical foil in water and in water with polymer solutions
ejected from a 0.5 mm diameter tube at a distance of 20 mm upstream of the
tip. Prior to the polymer tests, experiments conducted by ejecting a 50%
water + glycerin mixture, with a viscosity of 10-2 Pa.s, did not show any
significant contribution of the augmentation of the viscosity of an otherwise
Newtonian fluid. Ejecting a 500 ppm PEO solution when a quite large vapour
tube occupies the vortex core considerably modifies the appearance of the
cavitation. At an ejection velocity of about half of the free stream the
233

continuous very long cavity is reduced to a very short cavity of about half of a
maximum chord length and only scattered isolated bubbles are carried
downstream. By doubling the ejection velocity these entrained bubbles are
eliminated and only the shortened cavity remains. In terms of the incipient
cavitation number, Figure 14 shows inception results obtained with an ejection
of a 250 ppm PEO WSR 301 solution at an ejection velocity slightly larger
than the free stream.

2.5
cYi
no ejection
2.0
polymer
.....

1.5

1.0

0.5 I I I
5 7.5 10 12.5

Figure 14 9 Inception cavitation numbers as a function of the wing incidence


angle during the ejection of a 250 ppm PEO WSR 301 solution from a
0.5 mm diameter tube situated 20 mm upstream of the wing tip. From
Frurnan [43].

Thus, it appears that the polymer solutions have a strong effect on the tip
vortex cavitation occurrence and development. However, it is not clear what is
the mechanism responsible for such changes because, in these experiments,
neither the tangential velocities of the vortex nor the lift of the wing have been
measured. The latter is a very important factor since the tip vortex intensity is
directly related to the mid-span circulation of the foil and, as shown by Wu
[44] and Kowalski [45], lifting surfaces immersed in flowing homogeneous
polymer solutions display a reduction of lift.
In order to elucidate these aspects, experiments were conducted by ejecting
the polymer solution from the tip of the wing through a small diameter orifice
(Fruman et al. [46]), measuring both the tangential velocity distributions and
the incipient cavitation numbers. The results show a significant reduction of
the incipient cavitation number and, more interesting, a decrease of the
234

maximum tangential velocity and of the slope of the velocities in the core
region when the polymer solution was ejected, Figure 15. Measurements of the
lift did not show any change during polymer ejection (Aflalo [47], Fruman and
Aflalo [48]) in agreement with the fact that the tangential velocities in the
potential region were unchanged.

V, m/s
2 qi, m 3/s

0
1 0 . 7 x 10 -6
2.2 x 10 -6

-1 - U o o = 5 m/s

~=10 ~
~2
I I I I I I I I I I
-12 -10 -8 -6 -4 -2 2 4 6 8 10

1.5
V, m/s
1.0

0.5

--0.5
U =5m/s

-1.0 - 0~=5 ~

I I I I I I I I I I
-1.5
-12 -10 -8 --6 -4 -2 0 2 4 6 8 10

Figure 15 : Vertical (tangential) velocities as a function of distance to the


vortex axis for different ejection rates of a 1000 ppm PEO WSR 301 solution.
Measurements conducted at 0.20 m (five maximum chords) downstream the
tip. From Aflalo [47] and Fruman and Aflalo [48].
235

1.0 (P ~ - P o)P / (P ~ - P o)w a; o


u , m/s
[]
[] 5 5
0.8 [] 5.5 5
[] [] 0 5 10
0.6 0 5.5 10

[]
0.4 0 0

0.2
107x q-/
0 I I I I
0 0.5 1.0 1.5 2.0

Figure 16 9 Ratio of the pressure difference between infinity and the vortex
axis for the ejection of the polymer solution and the pure water as a function
of the reduced injection rate. 1000 ppm PEO solutions. From Aflalo [47].

Using these velocity distributions it was possible to integrate between a


position far from the vortex, where the pressure is p=, to a position on the
vortex axis, where the pressure is P0, the Vff/r term of equation (17) in the
case of pure water flow, index w, and polymer ejection, index p. The ratio of
the resulting pressure difference contributions are plotted in Figure 16 as a
function of the non dimensional ejection flow rate,

m qi c
qi =
U~ b Cma x

The results show that the polymer ejection causes a reduction of the pressure
difference as compared to the pure water situation. This reduction is largely
sufficient to justify the delay on the cavitation occurrence even if the
contribution of the normal stresses in equation (17) is ignored.
An investigation on the effect of homogeneous polymer solution on tip vortex
cavitation was conducted by Fruman and Aflalo [48]. The homogeneous
solution, of about 10 ppm, was obtained by injecting, upstream of the water
tunnel recirculating pump, a concentrated polymer (1500 ppm).
236

1.00 ( , Cavitationnumber at time t


umber at t = 0
0.95

0.90

0.85 -

Time after beginning mixing, s


0.80 i i i 1 I
0 200 400 600 800 1000

At time t=O

Attime t= lOOs

At fin~ t = 9OO s

Figure 17 : Top : evolution of the cavitation number as a function of the time


elapsed after the beginning of the mixing of a 1500 ppm PEO solution into the
recirculating water of the cavitation tunnel. Bottom : photographs showing the
elliptical foil (left) and the tip vortex cavity initiated at the tip and extending
downstream for pure water (t =0) and with increasing homogeneous polymer
concentrations (t =100 and 900 s). From Fruman and Aflalo [48].
237

The photographs of Figure 17 show that, in spite of the fact that the
modification of the flow rate and the reference pressure caused a significant
reduction of the cavitation number during the process of mixing, for an
initially well developed cavity the build-up of the polymer concentration in the
cavitation tunnel recirculating water was responsible for a significant decrease
of the cavity length. It is interesting to note also that the polymer seems to be
more effective at distances from the tip of several maximum chords and do not
causes the cavity to detach from the foil.

3.0
8
0
@0 water
2.5 -- iO polymer (9
0
2.0- w~ [] 0 0
Om o o
[] 0 0~
1.5-
u o~
o
1.0- [] O~

~ c Z

0.5 I I I I I t I
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Figure 18 9 Desinent cavitation numbers as a function of the lift coefficient for


the NACA 16020 elliptical foil in an homogeneous polymer solution. Open and
closed symbols refer to two free stream velocities, 8 and 8.4 m/s. From
Fruman and Aflalo [48].

The measurements of the lift of the foil showed that it was markedly reduced
in the homogeneous polymer solution. The tangential velocities were also
considerably reduced in both the core and the potential region. The latter is in
direct agreement with the reduction of the lift and hence of the mid-span
circulation. Figure 18 shows the desinent cavitation numbers as a function of
the measured lift coefficients in water and homogeneous polymer solution. The
effect of the polymer solution is to degrade the performance of the foil since,
for the same lift coefficient, the critical cavitation number is larger. The effect
of the polymer solution on the lift is largely more important than that on the
inhibition of cavitation inception. Since the boundary layer of the NACA
16020 section used in these tests is very sensitive to laminar-to-turbulent layer
transition (Pichon et al. [49]) it may well be that the extremely large decrease
238

of the lift coefficient is due, in part, to the already mentioned destabilizing


effect of the homogeneous polymer solutions.
The effect of doubling the size of the foils and increasing the free stream
velocity, increasing thus the Reynolds number, on the effectiveness of the
polymer ejection at the tip of the foil was investigated recently by Fruman et
al. [50] with results comparable to those already reported, provided the
ejection flow rate was scaled with the Reynolds number. A very detailed
survey of the tangential and axial velocity distributions at different positions
along the vortex path was also conducted in the very near region (less than a
maximum chord) downstream of the tip. The original results, showing a
decrease of the maximum tangential velocity and of the slope of the solid body
rotation core, were confirmed. Again, the increase of the pressure on the
vortex axis due to the velocity modification gives a decrease of the desinent
cavitation number, Figure 19.

w Cp llum
2.0
I / "~-- crd withoutejection

J 0,~ I0 no ejection
1.5 ~ _ .. ~ m e r

L cre with polymer ejection


1.0

III

X /C max
0.5 I I I I I I I I I I
0 0.2 0.4 0.6 0.8 1.0

Figure 19 : Pressure coefficient on the vortex axis, computed using expression


(17) ignoring the stress terms, as a function of the distance to the tip. The
values of the desinent cavitation numbers for the conditions of pure water and
polymer ejection flows are also indicated. The minimum of the pressure
coefficients are very close to the desinent cavitation numbers. From Fruman et
al. [50].

The ejection of polymer solutions at the tip of the foils to delay cavitation has
been applied with success by Chahine et al. [51 ] to a five blade screw propeller
of about 29 cm diameter. Because of the relatively low aspect ratio of the
blades, the position of the tip vortex detachment was not as easy to establish as
239

in the case of the elliptical foils. Therefore, a preliminary investigation was


conducted in order to determine the positions of ejection ports allowing the
ejected fluid to be carried out into the tip vortex. Once these positions were
chosen, experiments were conducted to determine the effect of the polymer
concentration and solution flow rate. As an example, Figure 20 shows the
cavitation number for inception as a function of the polymer concentration for
the pure water conditions, for water/glycerin injection and for polymer
injection. The most important result is that, depending on the position of the
ejection ports, the effect of polymer injection can be unnoticeable or achieve a
decrease of the incipient cavitation number of up to 35%.

17 - C a v i t a t i o n n u m b e r
water + glycerin
16 injection
blade #5, no injection
15
blade #1 }injection
blade #3
14

13

12

~ ~ o l y m e r concentration, ppm
10 I I I I I I I
1000 2000 3000 4000 5000 6000 7000

Figure 20 9 Inception cavitation number as a function of the polymer


concentration for the pure water conditions, for water/glycerin injection and
for polymer injection. From Chahine et al. [51 ].

Inge [52] investigated the contribution of the stresses due to an Oldroyd B-


fluid to the pressure within a weak vortex. The numerical computations
showed that the pressure on the vortex axis was increased with only a small
influence on the velocity field. This finding is not confirmed by experiments,
which show that, whatever the conditions - polymer ejection or homogeneous
solutions -, the azimuthal velocity profiles are modified enough to give an
increase of the pressure on the vortex axis.
Hoyt [53] investigated the effect of homogeneous polymer solutions of PEO
on the onset of cavitation in a vortex device where the liquid was injected
tangentially from a single port and evacuated through an axial pipe. The
results are summarized in Table 2. Incipient cavitation numbers are defined as
240

the difference between the discharge pressure and the vapour pressure divided
by the pressure difference across the device, and the air content is non
dimensionalized by the value at saturation. There is a large inhibition effect
for the PEO solutions, even at very low concentrations. However, a non drag
reducing polymer, Carbopol, does not display an inhibition effect.

Table 2 : Incipient cavitation number and air content for tests in a vortex
chamber. From Hoyt [53].
Test liquid Incipient Cavitation Air content/Air
Number content at saturation
Dearated water 0.253 0.284
Tap water 1.480 0.701
1.330
PEO, 8.2 ppm 0.547 0.755
PEO, 12 ppm 0.513 0.709
PEO, 16 ppm 0.100 -
PEO, 17.5 ppm 0.107 0.709
Carbopol, 20 ppm 1.300 0.785

Bismuth [54] conducted tests in a long vortex chamber where the fluid was
introduced tangentially at one end through eight rectangular slits and evacuated
axially at the other end. The reference pressure was measured in a large
reservoir situated downstream of the vortex chamber and used to determine
the incipient cavitation number using, instead of the pressure drop across the
chamber as in Hoyt [53], the kinematic head at the exit section. For a 10 ppm
concentration of PEO, the cavitation onset was advanced, as compared to the
pure water results, Figure 21, in complete disagreement with the earlier
results by Hoyt [53]. It has to be pointed out that, because of the different
definition of the cavitation number, direct comparison of these results is
difficult. However, if it is accepted that the drag reducing properties of the
polymer solution should increase, at equal pressure drop, the flow rate, the
results by Hoyt [53] should correspond to much larger inhibition effects. On
the other hand, velocity measurements made by Bismuth [54] did show that the
tangential component of the velocity increased, as compared to pure water,
when moving from the wall towards the center of the vortex. This increase
was large enough to justify, by integration of the simplified radial momentum
equation (equation (17) without the normal stress terms), the enhanced
cavitation characteristic of this flow. Moreover, since the flow rate determines
the ejection tangential velocity, and hence the circulation, Figure 21 can be
compared to Figure 18, where the lift coefficient is also proportional to the tip
241

vortex circulation. In both cases there is augmentation of the incipient


cavitation numbers as compared to pure water.

20

[] []
19

18

17 -
EO]

16 -

f I I IQ' 10-3m3/~
15
1.0 1.1 1.2 1.3 1.4
I I I I Ejection tangential
11 12 13 14 velocity, m/s

Figure 21 9 Inception cavitation number in the vortex tube as a function of the


flow rate for water and a 10 ppm homogeneous polymer (PEO) solution.
From Bismuth [54].

Bismuth also performed experiments by ejecting, through an orifice of


1 mm diameter situated at the end opposite to the evacuation, semi dilute
solutions of PEO. In Figure 22, the ratio between the incipient cavitation
number with polymer ejection and the one for pure water are plotted as a
function of the ratio of flow rates, ejection/total, for different values of the
reference pressure in the downstream reservoir. As shown, the polymer
ejection significantly inhibits cavitation onset. This figure is very much
analogous to Figure 16 and leads us to think that polymer ejection in the center
of a confined vortex or of a tip vortex has the same qualitative effect. In well
developed cavitation conditions, the vapour tube on the chamber axis is made
to disappear over increased distances when the rates of ejection of the polymer
solution increase.
242

~Rp CrRw
1.00 PR
0.86
0.95
0.98
_ []
1.17
0.90 1.23
0
o 1.35
0.85

0.80 []
qi /Q x103
0.75 I I I I I I I I
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Figure 22 9 Ratio between the incipient cavitation number with polymer


ejection and the one for pure water as a function of the ratio of flow rates,
ejection/total, for different values of the reference pressure in the downstream
reservoir. From Bismuth [54].

4.4 Cavitation in confined spaces


The flow occurring between a moving and a fixed wall separated by micron
size gaps is of the lubrication type and is characterized by very large pressure
gradients in the flow direction and, within the usual assumptions, nearly
constant pressures in the direction normal to the flow. Moreover, because the
minimum pressure can be made equal or less than the vapor pressure,
cavitation can locally occur. It can be theorized that the elongational flow
prevailing in the vicinity of the minimum gap will promote the occurrence of
a viscoelastic contribution, if polymer solutions are employed, and modify the
conditions for cavitation onset and development. Ouibrahim et al. [55] have
conducted experiments in a micron size space, with a minimum gap, e,
comprised between 5 and 20 ~tm, confined by a rotating drum of radius, R,
equal to 7 cm, and a fixed flat wall with a Newtonian fluid (water) and a non-
Newtonian viscoelastic fluid (600 ppm aqueous solution of PEO WSR 301). In
this peculiar situation, the product of (e/R)Re, where Re is the Reynolds
number computed with the drum tangential velocity, mR, and e, are much
smaller than unity as lubrication theory requires. The cavitation number, is
defined as,

Prey -- Pv
or- I (18)
-~- P ( ( o R ) 2
243

where Prey is a reference pressure.


In the case of water at inception conditions, intermittent cavitation spots,
scattered along a very thin band parallel to the cylinder axis, appear
downstream of the minimum gap. If the cavitation number is decreased, the
cavitation spots increase in number, become less scattered, adopt the shape of
an arrowhead with the tip facing the upstream of the rotating cylinder and
show bubbles escaping from the trailing edge. When the cavitation number is
further reduced, a string of nearly regular cavities is formed. They have
characteristic dimensions which are of the order of a few millimetres,
separated by thin lateral liquid films.

O' i

(:rip= 2514 Re -1.88

-1.84
lO iw = 7287 Re

f O polymer
o water
Re
5 I llll I I I I
5 6 789 2 3 456789 2 2 3
10 10

Figure 23 9 Incipient cavitation numbers as a function of the Reynolds number


for water and polymer solution and e - 10 ~tm. From Ouibrahim et al. [55].

In the case of the polymer solution and of incipient conditions, intermittent


tiny cavitation spots in the shape of rod-like cells scattered along a very thin
band parallel to the cylinder axis, appear downstream of the position of the
minimum gap. When the reference pressure is decreased below the one for
cavitation inception, a nearly continuous vapour cavity, characterized by a
practically straight leading edge and a more or less undulating trailing edge,
develops parallel to the axis of the rotating cylinder. This continuous cavity
results from the lateral coalescence of the arrowhead cavities, present only for
a very small range of the reference pressure. The difference in morphology is
not a consequence of the modification of the shear viscosity, since tests
conducted with a water/glycerine solution, having a viscosity of 11 mPa.s,
much larger than the one for the polymer solution, 2.05 mPa.s, shows
244

arrowhead cavities analogous to the ones observed in the case of water. When
the cavitation number decreases, the trailing edge waviness increases while
oscillations occur in the flow and axial directions. For extreme conditions the
movement of the cavity trailing edge becomes chaotic.
In Figure 23, the inception cavitation numbers as a function of the Reynolds
number (computed using the solvent viscosity and the minimum gap) have
been plotted in logarithmic coordinates for e = 10 ~tm. As shown, the
initiation of the cavitation is delayed in the case of the polymer solution but the
slope remains the same as that in the case of water.
It can be hypothesized that the inhibition effect is associated with the elastic
contribution of the normal stresses, which has the net effect of reducing the
incipient cavitation number, that is to say, to increase the pressure over the
Newtonian contribution. The added pressure, APe, can be estimated thus from
the difference of incipient cavitation numbers at an equal Reynolds number,
by,

1
APe - 2 t% - tYiw l. (19)

fie, Pa.s APe, Pa


3
9 - ~ ,- -, --0.8236 -
8
7
6
5
2
4

Ou/Ox, s -1
10 I i I I i i I tt : 105
2 3 4 5 6789 2
10 3 10 4

Figure 24 9 Estimated elastic contribution to the pressure and elongational vis-


cosity as a function of the estimated elongational strain rate. From Ouibrahim
et al. [55].
245

In Figure 24, the values of APe have been plotted, together with an estimate
of the elongational viscosity defined as,

APe
~e = ~
c)u "/ Ox (20)

where c)u / Ox (=70 co) is the maximum elongational strain rate, as a function
of the latter. The elastic pressure contribution increases with the strain rate
increasing while the elongational viscosity decreases. The latter is up to four
orders of magnitude larger than the shear viscosity (a Trouton ratio as high as
104). These Trouton ratios and elongational viscosities are comparable to the
ones obtained for a variety of polymer solutions displaying an elastic behavior.
It seems that the effect of the polymer solution in the present situation is due to
the development of normal stresses in the elongational flow in the space
confined between the moving and the fixed wall.

5. C A V I T A T I O N EROSION

The inhibition or enhancement of material erosion caused by cavitation in


non-Newtonian liquids has been a subject of much conflict and debate. This is
not surprising since, in the case of Newtonian fluids, the mechanism of
cavitation erosion is not well understood either in stagnant or flowing liquids.
Indeed, erosion seems to result from the combined action of the pressure
wave, generated during the near spherical phase of collapse of the bubbles, and
the reentering jet generated during the terminal, highly non spherical, phase of
collapse. Moreover, in flowing liquids, the intensity of erosion (rate of
material removal) increases very rapidly when the free stream velocity
increases. Finally, the interaction between the mechanical and crystalline
properties of the eroded material and the hydrodynamics of the bubble can be
quite strong.
Tests in stagnant fluids have been conducted by inducing ultrasonic cavitation
on metallic samples and determining the weight loss as a function of time.
Ashworth and Procter [56], with copper test specimens placed at 1.3 mm
below the tip of an ultrasonic probe (horn) obtained, as compared to the
results in pure water, a decrease of the incubation time and an increase of the
erosion rate with a 1000 ppm concentration of the PAM agent in distilled
water, Figure 25. For a concentration of 100 ppm, no significant change in
behavior occurs. It has to be mention that, because polymers in solution are
partially degraded by extensive exposure to ultrasonic cavitation, these
investigators took the precaution to have a fresh solution continuously fed into
the test reservoir in order to achieve an equilibrium solution. In a glycerol-
distilled water mixture, having a viscosity comparable to that of the
"equilibrated" 1000 ppm PAM solution, the incubation time increased by a
246

factor of three as compared to pure water and the erosion rate decreases
slightly.

W e i g h t loss, mg J []
50

Fluid
40
water
J )J
"o glycerol-distilled
water mixture

20 100 p p m P A M
1000 p p m P A M

10
E r o s i o n t i m e , rain
I I
0 10 20 30 40 50 60

Figure 25 9 Weight loss as a function of time for copper specimens subject to


ultrasonic cavitation. From Ashworth and Procter [56].

Weight loss, mg f = 19.5 k H z


50
A = 38 ~tm
..O
t I = 295 K
40
J

Fluid
30
0 water
20 - .~/ [] 100 p p m P E O
A 500 p p m P E O
0 1000 p p m P E O
10

I i I I I I
0 10 20 30 40 50 60

Figure 26 9 Weight loss as a function of time for aluminium specimens subject


to ultrasonic cavitation. From Shima et al. [57].
247

Ten years after these early results, Shima et al. [57] conducted very similar
experiments, but used PEO instead of PAM and had the aluminium test
specimen attached to the vibrating rod. Their results are very much different
of those of Ashworth and Procter [56]. First, no incubation time seems to be
necessary either in water or polymer solution; second, a 100 ppm solution
shows a behavior similar to that of water with a slight increase of the weight
loss; third, for 500 and 1000 ppm solutions the weight loss increases for a
short time after initiation and decreases significantly thereafter, Figure 26.
When the 1000 ppm polymer solution is moderately degraded the weight loss
increased in the early stages and decreased afterwards while for a fully
degraded solution the results are very close to those for pure water. Again, a
water/glycerol mixture decreased the material removal. Tsujino [58] pursued
the research by Shima et al. [57] and investigated the effect of the peak-to-peak
amplitude. He showed that when the amplitude is decreased, from 38 to
25 ~trn, water and a 1000 ppm PEO solution have exactly the same behavior
during a very short duration (-10 min) after initiation and then differ, the
aggressiveness being much decreased in the polymer solution.

50 Weight loss, mg f = 19.5 kHz


t I = 295 K
40 - A= 38 l a m ~
A, vibratory amplitude

30 - ~ ~'0

A Shimaet al. [57]


20
- -

9 Tsujino[58]
_ / / / / ~ ' \ A = 25 gm O Ashworth and
10 Procter [56]

I Test ti~ne, min I


0
0 10 20 30 40 s0 60

Figure 27 9 Weight loss as a function of time in water. Adapted from


Ashworth and Procter [56], Shima et al. [57] and Tsujino [58].

As shown, the results obtained by Ashworth and Procter [56] on one side, and
Shima et al. [57] and Tsujino [58] on the other, are markedly different in spite
of the fact that the two polymers used in the tests are very effective drag
reducing agents, display analogous viscoelastic effects, and have, for equal
concentrations, comparable shear viscosities in solution. The major differences
248

have to do with the materials and with the position of the test specimens in the
ultrasonic pressure field. However, in the case of pure water, the weight loss
versus time remains nearly linear regardless of the specimen material, the
position of the sample and the amplitude of the ultrasound, Figure 27. In the
polymer solutions, a linear behavior has been shown to exist by Ashworth and
Procter, while a power law behavior, with a power of less than unity, has been
demonstrated by Shima et al. and Tsujino. In a quite recent work, Hoyt and
Morgan [59] have reported results analogous to those of the Japanese
researchers when a 200 ppm PAM was added to a 6% mixture of tomato
puree in water.
How then to interpret the changes in the case of the polymer solutions ? No
explanation has been proposed as yet, although the visualisation conducted by
Shima et al. and Tsujino show conclusively that the development of the
cavitation bubbles in the more concentrated PEO solutions is very different
from the one observed in pure water. In the latter case, a cloud of bubbles
develops in an orderly fashion on the surface of the horn with a thickness
increasing from the rim to the center and, away from this surface and around
the axis, a column of small bubbles extending quite far away (one to two times
the diameter of the horn). When the erosion progresses, the thickness of the
cloud seems to increase on the surface and, in particular, around the center of
the sample. In 500 and 1000 ppm polymer solutions, isolated bubbles, clouds
or filaments, occupying a cylinder of at least the diameter of the horn, occur
away the sample surface and on the side wall of it, in what can be said to be a
disorderly fashion (as compared to water). When time, and hence erosion,
progresses, bubble clouds become denser and thicker on the surface and on a
column extending away from it. Therefore, it seems that, in the polymer
solutions, low pressure regions, where cavitation can occur, are developed
because of macro scale (of the size of the sample diameter) changes in regions
which are not even supposed to be affected by the vibration, such as the side
wall of the sample and the column mentioned above. For the test conditions of
Shima et al. [57] and Tsu~ino [58], it seems reasonable to hypothesize that the
very dense and thick cavitation, developing on the test specimen surface when
erosion progresses, has a cushioning effect that reduces the damage due to
bubble (cloud) collapse. For the Ashworth and Procter test geometry, this
cushioning effect is reduced because of the distance between the sample and the
tip of the vibratory horn and the modification of the macro scale (which is in
this case of the order of magnitude of the space between the sample and the
tip). Ting [60] offered a completely different explanation of the Ashworth and
Procter results. He claimed that the enhanced erosion resulted from the "solid-
like" behavior of the reentering jet and the additional impact stress developed
in the extremely rapid deformation process undergone by the fluid. In view of
the results by Shima et al. [57], Tsu~ino [58] and Hoyt and Morgan [59] this
reentering jet erosion enhancement effect does not seem to be of primary
249

importance. In fact, because all of the single bubble behavior investigations


indicate that the effect of viscoelasticity is to keep the bubbles spherical much
longer than in the Newtonian solvent, it seems reasonable to expect, if no
macro scale changes of the flow occur, a reduction of cavitation erosion.
Only one paper (Shapoval and Shal'nev [61]) refers directly to erosion in a
flowing PAM solution. In their tests, cavitation was generated in the wake of a
cylinder attached to the wall of a rectangular section. They determined the
cavitation numbers corresponding to the onset of cavitation and to the situation
corresponding to a cavity having a given length. In the polymer solution onset
is quite significantly delayed as compared to water. For a cavity having a
length of the cavitation zone equal to the cylinder diameter, the cavitation
number in the polymer solution was also reduced but proportionally less than
for the onset condition. The removal of material from a sheet of rolled lead
placed on the wall of the test section behind the cylinder was reduced in the
case of the polymer solution when the length of the cavity was twice the
diameter of the obstacle. Assuming that the pattern of the eroded regions,
limited to the shear layer, reflects the structure of the flow behind the
cylinder, the authors note that cavitation in the polymer solution appears to be
analogous to that in water at larger values of the Reynolds number.
Again, in these Russian experiments it is impossible to conclude the reasons
for the apparent decrease of the erosion. This is more so considering that the
region where the erosion imprint was analyzed corresponds to the shear layer
in the wake and to conditions which seem to be far from those prevailing at the
closure of an attached cavity where a stagnation point occurs. This latter
situation was discussed by Ting [60] for travelling bubbles.

6. C O N C L U S I O N

A considerable number of investigations have been conducted towards the


understanding of non-Newtonian fluid effects on cavitation occurrence and its
consequences. The major conclusions can be summarized as follows 9
9 for single bubbles,
- non-dimensional time evolution of spherical bubbles in a stagnant and
unbounded domain is unaffected by the non-Newtonian properties of
the fluids,
- in a domain bounded by a solid or free surface and in a shear layer, the
non-spherical collapse of bubbles is affected by the non-Newtonian
properties which delay the departure from sphericity,
9 in moving liquids such as 9
- in submerged jets, results show that cavitation can be either delayed or
unaffected by the non-Newtonian fluids and that major changes seem to
be associated with modifications of the structure of the free-shear
layer,
250

- in flow around blunt bodies, major changes appear to be correlated


with the early occurrence of laminar to turbulent transition in drag
reducing polymer solutions, without any major contribution associated
with bubble dynamics,
- in tip-vortex flow situations with homogeneous polymer solutions, the
non-Newtonian effects seem to enhance, for equal foil lift coefficient,
the cavitation occurrence, while, in pure water and polymer solution
ejection from the wing tip, there is a systematic cavitation inhibition,
- in confined vortex flow in homogeneous solutions, the results show
either cavitation enhancement or inhibition, without an explanation for
the causes of such conflict,
- in lubricating type flows, cavitation is delayed as a result of the
development of a normal stress contribution to the pressure.
Erosion due to cavitation can also be either augmented or diminished
depending on the way the tests are performed.
In summary, besides for a few controlled flow situations, such as the ones
utilized for investigating single bubble behavior, there is not a definite answer
concerning the favorable or unfavorable effect of non-Newtonian fluid
properties on cavitation occurrence and the consequences of its development.
The major obstacle to a complete understanding of this problem lies on the
fact that polymer solutions have, at least for the range of flow parameters
encountered in laboratory test situations, strong effects on the boundary layers
developing along streamlined bodies, on the structure of large eddies in shear
layers of separated flows, on the lift of hydrofoils, etc. These large scale
effects are further complicated by the behavior of the individual cavitating
bubbles, essentially determined by the local pressure and the presence of solid
boundaries or fluid interfaces at distances smaller than the bubble diameter;
thus relatively small scales. It is our opinion that some of the many questions
that remain unanswered may receive an appropriate response if test can be
conducted at much larger Reynolds numbers than those achieved up to know
and with very dilute polymer concentrations. However, the degradation of the
polymer molecules may have then a negative impact on the results of the
cavitation tests.

ACKNOWLEDGEMENT

The authors wishes to thank the support he has received over many years by
the Ministry of Defence, France, to conduct research on polymer solutions
effects in hydrodynamics. He is deeply indebted to Professors Jack W. Hoyt
and William R. Schowalter for taking the time to read a drat version of this
chapter and provide the author with many correction and wise comments.
251

NOMENCLATURE
r concentration
D diameter of headforms
e minimum gap in lubrication tests
G non-dimensional injection rate of polymer fluid
k polytropic index
L distance between the center of a bubble and a wall
m,n power law parameters
P pressure
po pressure at infinity at t = 0
pc critical pressure
Pd pressure in a reservoir
Pg non-condensible gas pressure within the bubble
PgO non-condensible gas pressure at t = 0
p(R) pressure at the interface of a bubble on the liquid side
Pv vapour pressure
p~(t) pressure far from the bubble
Prez reference pressure
qi non-dimensional flow rate
Q volume flow rate of ejected solution
R radius of the rotating drum in lubrication tests
Ro radius of a nucleus
RA distance between the position of the bubble interface, opposite
to a wall, and the center
Re Reynolds number
Rc critical radius
Rch initial radius for a collapsing bubble
R(t) bubble radius at time t
(r, O, (p) spherical coordinate system
t time
U radial velocity component
U free stream velocity
vj bulk velocity of a jet
vo tangential velocity

l~ts boundary layer displacement thickness


Y surface tension and shear rate
7? apparent viscosity
r viscosity
252

elongational viscosity
P liquid density
O"d desinent cavitation number
O"i incipient cavitation number
"t'R Rayleigh time
Trr, TOO, T ~ normal components of the deviatoric stress tensor
CO angular velocity of the drum in lubrication tests
APe added pressure in lubrication tests

REFERENCES

~ R.T. Knapp, J.W. Daily and F.G. Hammit, Cavitation, McGraw-Hill, New
York, 1970.
. D.H. Trevena, Cavitation and tension in liquids, 1987, Adam Hilger,
Bristol and Philadelphia.
. W. Tillner et al., The avoidance of cavitation damage, Mechanical
Engineering Publications Limited, London, 1993.
. C.E. Brennen, Cavitation and bubble dynamics, Oxford University Press,
Oxford, 1995.
. A.J. Acosta and B.R. Parkin, J. of Ship Research, 19, 1975, pp. 193-205.
6. A. Shima and T. Tsujino, Chemical Engineering Science, 31, 1976, pp.
863-869.
~ H.S. Fogler and J.D. Goddard, Phys. Fluids, 13 (5), 1970, pp. 1135-1141.
8. H.S. Fogler and J.D. Goddard, J. Appl. Phys., 42 (1), 1971, pp. 259- 263.
9. A.T. Ellis and R.Y. Ting, NASA-SP 304, Pt. 1, 1974, pp. 403-420 (with
discussions).
10. W.J. Yang and M.L. Lawson, Journal of Applied Physics, 45 (2), 1974.
11. G.L. Chahine, 1974, Thesis, Univ. Paris VI, 1984, Rapport de recherche
ENSTA 042.
12. R.Y. Ting, Phys. Fluids, 20 (9), 1977, 1427-1431.
13. P.S. Kezios and W.R. Schowalter, Phys. Fluids, 29 (10), 1986, pp. 3172-
3181.
14. G. Riskin, J. Fluid. Mech., 218, 1990, pp. 239-263.
15. G.L. Chahine and D.H. Fruman, Phys. Fluids, 22 (7), 1979, pp. 1406-
1407.
16. E.A. Brujan, C.-D. Ohl, W. Lauterborn and A. Philipp, Acustica, 82,
1996, pp. 423-430.
17. G.L. Chahine and A.K. Morine, ASME Cavitation and Polyphase Flow
Forum, 1980, pp. 7-8.
18. A.K. Morine, 1982, Thesis, Univ. Paris VI, Rapport de recherche
ENSTA 179, 1983.
19. G.L. Chahine, Applied Scientific Research, 38, 1982, pp. 187-197.
253

20. P.T. Kezios, PhD dissertation, Princeton University, 1984.


21. P. Ligneul, Phys. Fluids, 30 (7), 1987, pp. 2280-2282.
22. J.W. Hoyt, ONR Drag Reduction Workshop, Boston, 1970.
23. J.W. Hoyt, 16th American Towing Tank Conference, Sao Paulo, Brazil,
1971.
24. J.W. Hoyt, ASME Cavitation and Multiphase Flow Forum, pp. 44-47,
1973.
25. J.W. Hoyt, ASME Journal of Fluids Engineering, 98, 1976, pp. 106-112.
26. C.B. Baker, J.W. Holl and R.E.A. Arndt, ASME Cavitation and
Polyphase Flow Forum, 1976, pp. 6-8.
27. J.W. Hoyt and J.J. Taylor, ASME Journal of Fluids Engineering, 103,
1981, pp. 14-18.
28. R.E.A. Arndt, ASME Cavitation and Polyphase Flow Forum, 1980, pp.
9-10.
29. R.Y. Ting, AIChE Journal, 20 (4), 1974, pp. 827-828.
30. J.W. Hoyt, 1l th International Towing Tank Conference, Tokyo, 1966.
31. A.T. Ellis, J.G. Waugh and R.Y. Ting, Journal of Basic Engineering, 92,
1970, pp. 459-466.
32. C.B. Baker, R.E.A. Arndt and J.W. Holl, 1973, App. Res. Lab. Tech.
Memo. 73-257, The Penn. State Univ.
33. J.H.J. van der Meulen, ASME Cavitation and Polyphase Flow Forum,
1973, pp. 48-50.
34. J.H.J. van der Meulen, 1976, Doctoral Thesis, Netherlands Ship Model
Basin.
35. E.M. Gates and A.J. Acosta, 12th Symposium on Naval Hydrodynamics,
National Academy of Sciences, Washington, 1978.
36. H. Reitzer, C. Gebel and O. Scrivener, J. of Non-Newtonian Fluid Mech.,
18, 1985, pp. 71-79.
37. J.H.J. van der Meulen, ASME Cavitation and Polyphase Flow Forum,
1976, pp. 4-5.
38. C.E. Brennen, J. Fluid Mech., 44, 1970, pp. 51-63.
39. V.A. Bazin, Y.E.N. Barabanova and A.F. Pokhil'ko, Fluid Mechanics
Soviet Research, 5, 1976, pp. 79-82.
40. R.Y. Ting, Phys. Fluids, 21 (6), 1978, pp. 898-901.
41. R.J. Gordon and C. Balakrishnan, Nature Physical Science, 231,1971, pp.
177-178.
42. C. Inge and G. Bark, TRITA-MECH-83-12, 1983, The Royal Institute of
Technology, Sweden.
43. D.H. Fruman, ASME Cavitation and Polyphase Flow Forum, 1984, pp.
73-76.
44. J. Wu, Lift reduction in additive solution, Journal of Hydronautics, 3 (4),
1969, pp. 188-200.
45. T. Kowalski, Journal of Hydronautics, 5 (1), 1971, pp. 11-14.
254

46. D.H. Fruman, D. Bismuth, S.S. Aflalo, in The influence of polymer


additives on velocity and temperature fields, B. Gampert ed., Springer-
Verlag, Berlin, 1985, pp. 399-4.
47. S.S. Aflalo, Thesis, Univ. Paris VI, 1987.
48. D.H. Fruman and S.S. Aflalo, ASME Journal of Fluids Engineering, 111,
1989, pp. 211-216.
49. T. Pichon, A. Pauchet, A. Astolfi, D.H. Fruman, J-Y. Billard, Journal of
Ship Research, 41 (1), 1997, pp. 1-9.
50. D.H. Fruman, T. Pichon, P. Cerrutti, J. Mar. Sci. Technol., 1, 1995, pp.
13-23.
51. G.L. Chahine, G.F. Frederick and R.D. Bateman, ASME Journal of
Fluids Engineering, 115, 1993, pp. 497-503.
52. C. Inge, TRITA-MECH-83-05, 1983, The Royal Institute of Technology,
Sweden.
53. J.W. Hoyt, ASME Cavitation and Polyphase Flow Forum, 1978, pp. 17-
18.
54. D. Bismuth, Thesis, Univ. Paris VI, 1987, also, Rapport de Recherche
ENSTA 215.
55 A. Ouibrahim, D.H. Fruman and R. Gaudemer, Phys. of Fluids, 8 (7),
1996, pp. 1964-1971.
56. V. Ashworth and R.P.M. Procter, Nature, 258, 1975, pp. 64-66.
57. A. Shima, T. Tsujino, H. Nanjo and N. Miura, ASME Journal of Fluids
Engineering, 107, 1985, pp. 134-138.
58. T. Tsujino, Ultrasonics, 25, 1987, pp. 67-72.
59. J.W. Hoyt and J. Morgan, ASME Cavitation and Multiphase Flow Forum,
1989.
60. R.Y. Ting, Nature, 262, 1976, pp. 572-573.
61. I.F. Shapoval and K.K. Shal'nev, Sov. Phys. Dokl. 22 (11), 1977, pp.
635-637.
255

LOW-DIMENSIONAL DESCRIPTION OF VISCOELASTIC TAYLOR-


VORTEX FLOW

Roger E. Khayat

Department of Mechanical & Materials Engineering,


The University of Western Ontario, London, Ontario, Canada N6A 5B9

1. INTRODUCTION

The influence of inertia and elasticity on the onset and stability of Taylor
vortex flow (TVF) is examined for an Oldroyd-B fluid. The rigid-flee (stick in
the azimuthal direction and slip along the axial direction) and rigid-rigid
boundary conditions are used. Both formulations lead to essentially the same
qualitative stability picture, but the former conditions allow a much simpler
formulation, more easily amenable to algebraic manipulation and analysis. The
resulting equations reduce to the Lorenz equations for a Newtonian fluid and
small number of modes. A truncated Fourier/Chandrasekhar representation of the
flow field, and the Galerkin projection, lead to a closed nonlinear dynamical
system with sixteen degrees of freedom. In contrast to the six-dimensional
system derived previously [Khayat, Phys. Fluids A 7 (1995) 2191 ], the present
model is capable of capturing the dynamical behavior observed experimentally
for a (Boger) fluid with negligible inertia. The improvement stems mainly from
the proper choice and inclusion of higher-order modes that account for normal
stress effects. For flow with dominant inertia, the stability picture is similar to
that for a Newtonian fluid: steady TVF sets in as the Reynolds number reaches a
critical value, Reo which decreases with fluid elasticity or normal stress effects,
and is strongly influenced by fluid retardation. As elasticity exceeds a critical
level, a subcritical bifurcation emerges at Re o similar to the one predicted by the
Landau-Ginzburg's equation and the previous six-dimensional model. It is found
that slip tends to be generally destabilizing. The coherence of the model is
established through comparison with existing linear stability analyses, and
experimental measurements and flow visualization.
256

Recent linear stability analyses and experiment indicate a dramatic departure


in the stability and bifurcation of the Taylor-Couette flow of viscoelastic fluids in
comparison to Newtonian fluids. The Taylor-Couette flow of Newtonian and
viscoelastic fluids involve a rich sequence of fundamental flow regimes that are
observed during experiment, coveting the range from laminar to turbulent
motion. For a Newtonian fluid, it is observed that at a sufficiently small
Reynolds number, Re, there is a unique stationary (Couette) flow, which is
globally stable, except perhaps near the ends of the cylinders. When Re is raised
in the range near a critical value, R e o the stationary flow loses its stability, and
develops a regular cellular vortex structure in which closed ring vortices
alternating in sign are wrapped round the axis of rotation. In this range of
Reynolds number, the flow remains essentially time independent and
axisymmetric. This is the well known Taylor vortex flow (TVF) after G. I.
Taylor, who was the first to examine this flow regime both theoretically and
experimentally [ 1]. At higher Re value, the stationary cellular structure loses its
stability in turn to another structure with a different number of vortices [2,3]. The
upper limit of the TVF regime coincides with the emergence of time-dependent
flow and the breaking of axisymmetry. Simultaneously, there appears a
uniformly rotating pattern, or wavy vortex flow (WVF), of tangential azimuthal
traveling waves superimposed on the cellular structure of TVF [4].
The TVF-WVF transition was carefully investigated by Fenstermacher et al.
[5] and Gorman & Swinney [6] through flow visualization and spectral
techniques. They obtained an accurate picture of the spatio-temporal flow
structure which was later theoretically examined by Rand [7]. As the WVF sets
in, the amplitude of the azimuthal waves is shown in the experiments to grow
continuously with Re. The velocity power spectrum indicates the appearance of a
single discrete frequency component with its harmonics. As Re is further
increased, a second fundamental appears, with a gradual increase in the small-
scale structure. At higher Re value, there is a broadening of the base in the power
spectrum, indicating the onset of aperiodicity or chaos.
For highly elastic (Boger) fluids, normal stress effects appear to prohibit the
onset of steady TVF. The experimental measurements of Muller et al. [8] clearly
demonstrate the existence of a purely elastic overstable mode when the Deborah
number exceeds a critical value for a fluid with vanishingly small inertia.
Periodicity appears to be lost as the level of elasticity is further increased. The
influence of inertia was later examined by Baumert & Muller [9] through flow
visualization. More extensive experiments are, however, needed that could
elucidate on the possible route(s) to instability and transition to turbulence. One
of the main objectives of the present paper is to actually predict the conditions
257

under which transition may occur in the Taylor-Couette flow of viscoelastic


fluids.
Fluid turbulence is among the most challenging areas in the physical sciences,
and has been the subject of intensive research over the past hundred years or so.
Similarly to any flow in the transition or turbulent regime, the Taylor-Couette
flow involves a continuous range of excited spatio-temporal scales. In order to
assess the effect of the motion of the arbitrarily many smaller length scales, one
would have to resolve in detail the motion of the small scales. This issue remains
an unresolved one since, despite the great advances in storage and speed of
modem computers, it will not be possible to resolve all of the continuous range
of lengths scales in the transition regime. It is by now well established that low-
order dynamical systems constitute an alternative to conventional numerical
methods as one strives to understand the nonlinear behavior of flow [ 10].
For the Rayleigh-Benard thermal convection problem, the simplicity of the
Lorenz equations [11], and the rich sequence of flow phenomena exhibited by
their solution, have been the major contributing factors to their widespread use as
a model for examining the onset of chaotic motion [12]. Despite the severe level
of truncation in the formulation of these equations, some of the basic qualitative
elements of the onset of thermal convection and the destabilization of the cellular
structure have been recovered through the model. More general problems in fluid
dynamics have also been treated using low-dimensional descriptions and the
theory of nonlinear dynamics [13]. These methods are based on the expansion of
the flow field in terms of a complete set of orthogonal functions, Fourier series or
other standard basis functions, and the Galerkin projection technique whereby
the initial set of partial differential equations is decomposed into an infinite set of
ordinary differential equations governing the time-dependent expansion
coefficients. The system is then truncated, leading to the finite-dimensional
dynamical system. These methods have mainly been applied to Newtonian
fluids, and the author appears to be the first to extend their application to non-
Newtonian fluids [ 14-19].
While the problem of Taylor-Couette flow has been extensively investigated
for Newtonian fluids, relatively little attention has been devoted to the case of
viscoelastic fluids. In addition to the relatively recent interest in polymeric flows,
the onset of chaos and turbulence is far less widespread in viscoelastic than in
Newtonian fluids. To our knowledge, there has been no experimental evidence of
the existence of (deterministic) chaos for highly elastic polymeric solutions
similar to the one found for Newtonian fluids [5,6]. Recently, however, there has
been a growing interest in the Taylor-Couette flow of viscoelastic fluids [20,21 ]
for a recent review). Some of the early experimental work on the stability of
rotating viscoelastic fluids is that of Giesekus [22,23]. Using polyacrylamide
258

solutions [23], he showed that the critical value of the Reynolds number at the
onset of Taylor vortices may be smaller or higher than the Newtonian value
depending on polymer concentration. At high polymer concentration, when
elastic effects become dominant, the critical Reynolds number was found to
decrease with the flow Deborah number, De. However, at high polymer
concentration, solutions may exhibit significant shear thinning. Thus, ideally,
both viscoelastic and shear thinning effects must be accounted for through a
suitable constitutive model. The later experiments of Haas & Biahler [24] tend to
show similar results. The experiments of Muller et al. [8] clearly demonstrate,
for the Taylor-Couette flow of a Boger fluid, the existence of a purely elastic
time-periodic instability at a critical rotation rate. The experiments were
conducted using Laser Doppler velocimetry (LDV) measurements of the axial
velocity component of a polyisobutylene-based fluid between two concentric
cylinders, with the outer cylinder being at rest and the inner cylinder rotating.
The results show an oscillatory flow at a vanishingly small Reynolds number (Re
-~ 7x 10-3). The flow appears to undergo a transition from the purely azimuthal
Couette flow to time periodic flow as the Deborah number exceeds a critical
value, which is in good agreement with the value based on earlier linear stability
analysis of an Oldroyd-B fluid [25]. The LDV measurements show that the
oscillatory behavior is not localized but appears to be spread throughout the
flow. As the Deborah number increases from the critical value, the amplitude of
oscillation increases monotonically with De. The corresponding power density
spectra show peaks, which are instrumentally sharp at the fundamental
frequency, the growth of harmonics, and eventually subharmonics, reflecting,
perhaps, the presence of period doubling or emergence of a second fundamental
frequency (quasiperiodicity). The influence of inertia (shear rate) on the
evolution of TVF was later examined by Baumert & Muller [9] who performed
flow visualization on Boger fluids using reflective mica platelet seeding.
Parallel theoretical work, for axisymmetric and non-axisymmetric flows, has
also been conducted on the linear stability of viscoelastic fluids [25-27] and
nonlinear calculations [28-30], with new phenomena being constantly attributed
to fluid elasticity [31,32]. Unlike Newtonian fluids, which obey Newton's law of
viscosity, viscoelastic fluids are not governed by a similar universal constitutive
law. The dependence of the predicted flow behavior on the particular choice of a
constitutive model adds another difficulty in our attempt to interpret an already
complex flow situation, particularly in the transition and turbulent regimes.
Larson [33] carried out a linear stability analysis in the narrow-gap limit,.using
the Doi-Edwards and K-BKZ constitutive equations. He found that the
dependence of the critical Reynolds number on De is generally non-monotonic,
but the flow is increasingly destabilized by fluid elasticity in the higher De range
259

[33]. Numerical calculations were carried out by Northey et al. [30] for the
Taylor-Couette flow of an upper-convected Maxwell (UCM) fluid neglecting
inertia. The non-axisymmetric flow solution was obtained by Avgousti & Beris
[28] and Avgousti et al. [29]. Elastic overstability was also proved through
various linear stability analyses [25,26], showing that the base (Couette) flow
can lose its stability via a Hopf bifurcation as the Deborah or Weissenberg
number exceeds a critical value. Most of existing analyses are, however, limited
to the linear regime.
In our quest to probe the nonlinear regime, we have undertaken the current
series of studies on the influence of fluid elasticity on Rayleigh-Benard thermal
convection and Taylor-Couette flow. Khayat [14-17] developed a four-
dimensional dynamical system for the thermal convection of strongly elastic
fluids of the Oldroyd-B type. Such a system constitutes a generalization of the
Lorenz equations [11] to include viscoelastic fluids. The critical Rayleigh
number at the onset of the convective cellular structure was found to be the same
as for Newtonian fluids. This is a direct consequence of the fact that the
nonlinear terms (at least to the degree of truncation adopted) are the same as
those in the Lorenz equations; no convective or upper-convective nonlinear
terms survive in the constitutive equations. The conductive state thus loses its
stability to the two steady convective branches C1 and C2, say, through a
supercritical bifurcation similarly to Newtonian fluids. The two convective
branches lose their stability in turn through a Hopf bifurcation as the Rayleigh
number exceeds a value that, this time, depends strongly on fluid elasticity and
retardation. It was also observed that fluid elasticity tends to precipitate the onset
of chaotic motion, while fluid retardation tends to delay it. Above a critical value
for the Deborah number, the flow behavior departs significantly from that of a
Newtonian fluid. All three fixed branches remain unstable in the supercritical
range for any value of the Rayleigh number. Thus, for D e > D e c, no steady
convection can set in, and the cellular structure is always periodic in time
(overstable), with the corresponding Fourier spectrum showing a sequence of
period doubling as D e is increased. The existence and stability of the Hopf
bifurcation corresponding to the onset of overstability were established in a later
paper [15] using center manifold theory. These findings have prompted us to
adopt a similar approach to the Taylor Couette flow of viscoelastic fluids in the
narrow gap limit [18,19].
Our earlier work [ 18] focused on the influence of elasticity and retardation on
the sfability and amplitude of the Taylor vortices with inertia dominating the
flow. A truncated Fourier representation with Galerkin projection of the flow
field similar to that of Kuhlmann and associates for a Newtonian fluid was
proposed. Kuhlmann [33], and later Kuhlmann et al. [35] examined the
260

stationary and time-periodic TVF, in the narrow-gap limit and arbitrary gap
width, respectively, with the inner cylinder rotating at constant and harmonically
modulated angular velocity. The solution to the full Navier-Stokes equations was
obtained by implementing a finite-difference scheme, and an approximate
approach based on the Galerkin projection. Comparison of flows based on the
two methods led to good agreement. In the approximate Galerkin projection [ 18],
the velocity and stress components assume a truncated Fourier representation in
space, with the expansion coefficients being functions of time alone. Since
elastic effects were assumed to be weak, higher-order normal stress terms were
neglected, thus leading to a six-dimensional nonlinear system, and reducing to
Kuhlmann's three-dimensional system for a Newtonian fluid [18]. In both
models, the velocity is assumed to adhere to the cylinder surfaces in the
azimuthal direction, while it slips in the axial direction (rigid-free boundary
conditions). The corresponding rigid-rigid boundary model leads to more
realistic values of the critical Taylor number and the corresponding wave
number, but to a less realistic value for the torque. Kuhlmann's three-dimensional
system turned out to be equivalent to the Lorenz system with the Prandtl number
equal to one. In this case, Kuhlmann's model cannot predict the destabilization of
the Taylor vortices, and therefore cannot account for the onset of chaotic
behavior. The situation is exactly similar for the Lorenz system when the Prandtl
number is less than or equal to one [36,37].
In the previous work [18], a similar level of truncation in the Fourier
representation for the velocity and stress was adopted for an Oldroyd-B fluid.
Similar levels of truncation have been widely used for the Navier-Stokes and
energy equations [13,38]. Examination of the influence of additional modes on
flow and temperature fields [39-42] indicates that many of the gross features
predicted by low level models are essentially recovered by the higher-order
models. In the viscoelastic model, there are two Fourier modes in the azimuthal
velocity and shear stress components that are expected to lead to nonlinear terms
in the stress equations. This is in contrast to thermal convection [14-17] where
the nonlinearities originate from the convective terms in the energy equation.
While the stability and bifurcation picture at the onset of the cellular structure in
the Rayleigh-Benard convection of viscoelastic fluids is similar to that of
Newtonian fluids, the presence of nonlinearities in the stress equations of the
Taylor-Couette flow of viscoelastic fluids has a drastic influence on the stability
of the Taylor vortices and the transition to chaos. While the two models suffer
from a similar level of truncation, the addition of viscoelastic effect, even to a
very weak extent, appears to alter dramatically the stability and bifurcation
picture. Most importantly, unlike the Newtonian model, elasticity tends to
destabilize the TVF leading to chaotic behavior and the emergence of a Lorenz
261

type attractor. As the Deborah number increases, the onset of TVF occurs at a
Reynolds number that decreases with De. Beyond a critical De value, the
exchange of stability takes place via a subcritical instead of a supercritical
bifurcation. Since the previous formulation [18] is not valid in the presence of
dominant elastic effects, most results from existing linear analyses, finite
amplitude numerical calculations, and experiments on the purely elastic
overstability could not be recovered.
The aim of this chapter is to investigate the onset and stability of finite
amplitude TVF with elastic effects dominating over inertia. It is shown that the
purely elastic overstability can only be predicted if the higher-order normal stress
terms, neglected in the previous formulation [18], are properly accounted for.
Particularly, the addition of the azimuthal normal stress component z00 leads to
additional coupling with higher-order eigenmodes that can no longer be
neglected. The resulting nonlinear dynamical system involves sixteen instead of
six degrees of freedom. Although more cumbersome, and therefore less
amenable to algebraic manipulations, the present expanded model is more
accurate in its predictions, and leads to good agreement with existing
formulations and experiments. Both the rigid-free and rigid-rigid boundary
conditions are used and the resulting flows are compared. It is found that the two
formulations lead essentially to similar qualitative flows. The former conditions
allow a much simpler formulation. It also allows the examination of the influence
of the higher-order normal stress terms when compared with the previous
formulation of Khayat [18]. Additional calculations are also carried out in the
absence of inertia, based on a dynamical system that accounts for yet higher
order normal stresses, in an attempt to reach a quantitative agreement with the
measurements of Muller et al. [8].
The chapter is organized as follows. In w the derivation of the nonlinear
dynamical system is discussed. The stationary solutions and their stability are
examined in w The influence of inertia for weakly and strongly elastic fluids is
examined through numerical calculations, which are presented in w Discussion
and concluding remarks are covered in w

2. DERIVATION OF THE NONLINEAR DYNAMICAL SYSTEM

The derivation of the nonlinear dynamical system for a viscoelastic fluid is


summarized in this section. The governing equations are solved by assuming an
infinite discrete Fourier/Chandrasekhar representation in space for the flow field,
which, upon truncation and application of the Galerkin projection, leads to the
sixteen-degree-of-freedom nonlinear dynamical system (16NDS) that governs
the expansion coefficients.
262

2.1. General equations and boundary conditions


Consider an incompressible viscoelastic fluid of density 9, relaxation time X1
and viscosity q. In this study, only fluids that can be reasonably represented by a
single relaxation time and constant viscosity are considered. The fluid is assumed
confined between two infinite and concentric cylinders of inner and outer radii R i
and R o , respectively, and the flow is axisymmetric. The inner cylinder is taken to
be rotating at an angular velocity fl, while the outer cylinder is at rest. The
conservation of mass and linear momentum equations are, respectively:

V . u = O, Re + u . Vu ) = - V p -/-Rv V 2 u - V. z" (1)

where V is the gradient operator, t is the time, u is the velocity vector and x is the
elastic part of the deviatoric stress tensor obeying the following constitutive
equation [43]"

De + u. V r- (17u) t . z" -
1
r. Vu = - r - Vu- ( V u ) t . (2)

The various dimensionless groups in the problem, namely the Reynolds number,
Re, the Deborah number, D e , the solvent-to-polymer viscosity ratio, R v , and the
gap-to-radius ratio, ~ (not explicitly appearing in the equations above), are given
by:

dRi.O 2 R .O 22 d
Re - De = Rv = - , c= , (3)
w ' d " A1 - / Z 2 tip Ri

where qs and rip are the solvent and polymer viscosities, respectively, and
22 (0 <_22 < 21) is the fluid retardation time. Note that in this case q = rls + qp
and v= r//p. Other related dimensionless groups are the Taylor number, Ta, the
Elasticity number, E, and polymer-to-solution viscosity ratio, a:

Ta cRe 2 E- De Zip 1
9 (4)
Re" r/ Rv + l
263

The boundary conditions on the cylinder walls are prescribed as follows.


Regardless of the nature of the two cylinders, the no-penetration condition must
apply at any time. The no-slip conditions in the azimuthal direction are assumed
to hold. There remains the conditions along the axial direction. Generally, stick
(rigid) conditions should be assumed. However, the use of the less realistic flee-
slip conditions simplifies greatly the formulation. Such conditions were
previously used by Kuhlmann [34], and more recently by Khayat [18]. Similar
unrealistic conditions have extensively been adopted in the Rayleigh-Benard
thermal convection of Newtonian fluids [44] and viscoelastic fluids [45]; the two
planes at which the temperature differential is imposed are assumed to be free
surfaces.
In the present work, the free and rigid conditions will be used. Thus, the term
rigid-flee (RF) boundary conditions will be referred to if slip along the axial
direction applies, and the term rigid-rigid (RR) boundary conditions will be
referred to if stick is used. It will be concluded that little qualitative difference
emerges between the two formulations. As mentioned above, the introduction of
free conditions is customary in the case of a Newtonian fluid [34,35], and should
be compared to the free-surface assumption usually adopted in the Rayleigh-
Benard convection problem [ 13,38,44,46]. In that case, both the lower and upper
planes, separated by the fluid, are assumed to be free surfaces. This is an
unrealistic assumption, but it makes the problem formulation much easier to
handle since trigonometric rather than hyperbolic (Chandrasekhar) functions can
be used in the expansion of the solution along the gap.
For Taylor-Couette flow, a slip along the Z direction tends to enhance the
onset of Taylor vortices given the absence of an axial friction force acting
between the fluid and the cylinder walls. In the narrow gap limit, linear stability
analysis for a Newtonian fluid gives the minimum value of the critical Taylor
number and corresponding wave number for the onset of TVF, respectively, as
Ta m = 1695 and k m = 3.12 if rigid-rigid boundaries are assumed [44], whereas
for rigid-free conditions, Tacm = 654 and k m = 2.23 [34]. The neutral stability
curves in both cases are qualitatively similar. This is in close analogy to the
lowering threshold in the Rayleigh-Benard problem by using free-free boundary
conditions [4]. While rigid-rigid boundary conditions tend to give more accurate
values for the critical Taylor number and wave number, the mixed (rigid-free)
boundary conditions lead to a torque value at the onset of TVF much closer to
what existing formulations [34,35,47,48] predict. As to the nonlinear range, both
rigid-rigid and rigid-free conditions have been used with the same level of
truncation in a three-dimensional dynamical system [34]. The two models do not
lead to any significant qualitative difference in behavior. Based on these
264

arguments, one may then suspect that not much further insight would be gained
if rigid-rigid boundary conditions were used in the case of a viscoelastic fluid, at
least at the qualitative level. This is indeed confirmed in the present study
through comparisons with existing linear stability analysis results, experimental
measurements and flow visualization.

2.2. Governing equations in the narrow-gap limit


Consider the flow between two concentric cylinders in the limit when the
radius R i / R o is very close to one. Suitable scales are sought for length, time,
velocity and stress. There are, in fact, several choices possible [49-51]. One
obvious choice for the length scale is d - R 0 - R i . In this case, one expects e =
d / R i to emerge as the natural perturbation parameter for the problem, no matter
what the remaining scales are. However, a careless choice of these scales can
easily lead to the reduction of the equations to those governing plane Couette
flow if all terms of O(e) and higher are neglected. Additional complexities are
also expected in the case of a viscoelastic fluid since it is difficult to properly
assess the magnitude of the non-Newtonian contribution of the stress
components [18]. The flow induced by relatively high rate of rotation of the
inner cylinder is of special interest here since it leads to the generation of large
normal stresses. Following Tabeling [51], one velocity scale is chosen for the
flow components along the azimuthal direction, and another for the components
along the radial and axial directions [ 18]. If x, y and z are taken as the reduced
coordinates in the radial, azimuthal and axial directions, then the problem
consists of finding the variables Ui and ~0" where i,j = x,y,z.

2.3. Elimination of the z dependence


The general solution of equations (1) and (2) is carried out by seeking a
periodic flow field along the cylinder axis. Thus, the velocity and stress fields are
represented by infinite Fourier series in the z direction, with the expansion
coefficients depending on x and t. The Fourier modes have a fundamental
wavelength ~/k (in units of d) in the z direction. After elimination of the pressure
in the momentum equations, the flow variables can be written as spectral sums:

O0 OO

Zum(x,t)e imkz
ui(x,z,t) - 9 (5)
m - -oo m ~ ~oo
265

Additional relations among the Fourier coefficients are obtained by examining


the physical symmetry of the flow. A symmetry allowed by equations (1)-(2),
similarly to the Navier-Stokes equations [53,54], occurs when the flow is
invariant under z ~ - z, and uz ~ - U z , rce ~ - r a e (~ = x, y). This amounts to
having invariance in flow when the Taylor-Couette apparatus is turned upside
down [54]. Thus, in addition to being axisymmetric, the velocity and stress fields
must satisfy additional symmetry conditions It is not difficult to see that this
symmetry is preserved under the nonlinear multiplication in both the convective
and upper-convective terms of equations (1)-(2). The symmetry leads to
additional simplification
Upon application of the Galerkin projection method, which consists of
multiplying equations (1)-(2) by the corresponding modes above and integrating
with respect to z from 0 to ~k, an infinite set of coupled partial differential
equations governing the expansion coefficients is obtained. If equations (1)-(2)
were linear, the various modes involved would have separated in a manner
similar to the case of a Newtonian fluid [44], and one would examine the first-
order z-dependent terms. In the present nonlinear context, some of the zeroth-
order terms are also retained in order to ensure that part of the nonlinear
convective terms in the momentum equations, and the upper-convective terms in
the constitutive equations do not vanish in the projection process.

2.4. The nonlinear dynamical system


The most crucial step in the problem formulation is seeking an orthogonal
representation form of the expansion coefficients in (5) and imposing a suitable
level of truncation leading to the final nonlinear dynamical system. A judicious
selection process must thus be applied for the choice of the various modes in
order to ensure the physical and mathematical coherence of the final model. Not
only does the approximate flow solution have to satisfy the imposed boundary
conditions at the inner and outer cylinders, it must also reduce to the
corresponding solution for linear flow: 1) in the limit of zero-Elasticity number
for a Newtonian fluid, and 2) in the limit of vanishing inertia for a viscoelastic
flow. Since the linear behavior of a Newtonian fluid is well understood [44],
attention is first focused on the linear stability of a viscoelastic fluid with no
inertia. In this case, the governing equations reduce to those of Larson et al. [25]
in the narrow-gap limit.
The discussion is limited to the case of an upper-convected Maxwell (UCM)
fluid. Thus, R v is set equal to zero, and inertia effects are neglected in the
momentum equations. If the nonlinear terms are neglected in the remaining
266

0 A
equations, u~and (3' become decoupled from the remaining variables. For

linear stability analysis, one sets u l ( x , t ) = U ( x ) e - i a ) t , with similar


expressions for the remaining variables, where co is generally complex. In this
case, the stress components are explicitly expressed in terms of the velocity and
velocity gradient through equations. Upon elimination of the z velocity
component from the continuity equation, a fourth-order equation is obtained for
U(x) as in Larson et al. [25]. In this case, the quantity A= 2 oDe2 / k ( 1 - i a)De)2
becomes the eigenvalue of the problem once the boundary conditions at the inner
and outer cylinders are imposed. The solution of the resulting eigenvalue
problem, with both the RR and RF conditions, can be obtained using the direct
method, similarly to Larson et al. [25]. Further details are given elsewhere [56].
The eigenvalues are determined using an iteration scheme on the initial guesses
and the secant method until the determinant is less than an imposed small
tolerance. The eigenvalues A obtained using this method are found to be all
imaginary, leading to the same dispersion relation as that of Larson et al. [25]. It
is found that the critical (minimum) value of the Deborah number and
corresponding wave number for the onset of the most unstable (overstable) mode
are, respectively, Dec = 5.77e -1/2 and kc = 4.44, compared to the values 5.92e -1/2
and 6.7 in the case of RR boundary conditions [25]. This critical value for the
onset of linear overstability, and, more generally, the marginal stability curve in
the (De, k)-plane, should serve as limits (corresponding to small departure from
the base flow) for the flow field predicted by the nonlinear theory, where only an
approximate solution can be found.
An approximate solution is first sought for the linearized problem, and is then
compared with the exact solution from the direct method. For RF boundary
conditions the following truncated Fourier representation is adopted:

U ( x ) = U 1 sin 1rx + U 2 sin 2 erx, x ~ [0, 1] (6a)

which satisfies the RF conditions, whereas the solution that satisfies the RR
conditions is expressed in terms of the even and odd Chandrasekhar functions
denoted, respectively, by @l(X) and @2(x) and defined in [55]:

U ( x ) = U 1 ~ l ( X ) + U 2 tl~2(x), x ~ [-1/2, +1/2] (6b)

The corresponding characteristic equation for A, for both the RF and RR


conditions, may be written in the form:
267

)22/2
k6CiA2-(k 2 C2 ( k - C 3 =0, (7)

where C1, C2 and C3 are constants involving the inner product of the orthogonal
functions and their derivatives. These constants take, of course, different values
depending on whether equation (6a) or (6b) is used. The exact solution, when
compared with (6a) or (6b), gives an estimate of the magnitude of the error
resulting from the t~ncation in solution. One should expect that the error
remains of the same order when such a truncation is applied to the solution of the
full nonlinear equations for small departure from the base flow. It is clear from
eq. (7) that the neutral stability curves based on the approximate solutions 6a and
b, those based on the exact solution using the RF boundary conditions, and the
exact solution from [25] using the RR conditions exhibit similar behavior. At
small wave number, the critical Deborah number for overstability to set in
decreases sharply with k, reaches a minimum at k = kc, and increases again for k
> k c. All curves indicate a general flattening around kc, confirming the
experimentally observed wide range of wave numbers at which overstability sets
in [9,20].
On the basis of these observations, and further analysis of the influence of
inertia and fluid retardation (see section 3), one can now proceed and impose the
type and number of modes that may adequately describe the nonlinear flow field.
Although the solutions and results for both the RF and RR formulations are
presented in this work, more details will be given concerning the RF problem.
The RF formulation presents advantages as (i) the RF boundary conditions lead
to a much simpler formulation than the RR conditions, thus allowing further
algebraic manipulation and analytical investigation using various tools from the
theory of nonlinear dynamical systems, (ii) the resulting equations reduce to the
six-dimensional system studied earlier [ 18], thus allowing the examination of the
influence of the higher-order modes, previously neglected, (iii) these equations,
in turn, reduce further to the Lorenz system for a Newtonian fluid, a system
extensively investigated in the past thirty years or so [ 12].
It is first observed that, if only one fundamental eigenmode was kept in
expressions (6), one would obtain the trivial solution to the eigenvalue problem.
This corresponds exactly to the level of approximation adopted in the previous
work [18], where the flow of highly elastic fluids (De >> Re) could not be
treated. It will be seen later that this amounts to neglecting important normal
stress terms in the limit Re ~ O. The general solution, must then be chosen so
that (i) linear behavior is recovered for small deviation from the base flow, (ii)
268

the Newtonian flow is recovered in the double limit De, Rv ~ 0, and (iii) the
boundary conditions are satisfied.
Only two terms tumed out to be sufficient in the expansion in x for all the
velocity and stress coefficients u m ( x , t ) and rn~ij(x,t). For the RF conditions,
trigonometric (Fourier) functions are used, similarly to (6a), except that the
expansion coefficients are time dependent. For the RR conditions, hyperbolic
(Chandrasekhar) functions are used to satisfy the stick conditions, similarly to
(6b). There are thus sixteen independent modes that are retained. Higher-order
terms are included in the radial and axial velocity, and the normal stress
coefficients as opposed to keeping only one term in the previous six-dimensional
system [18]. The truncation level adopted in the previous work [ 18] is similar to
the one used by Kuhlmann [34] and Kuhlmann & Lficke [35] for a Newtonian
fluid, where two terms are kept only for the velocity and stress in the y direction,
and one term is retained for the remaining variables. This level of truncation does
not leads to the trivial solution of the (linearized) eigenvalue problem in the limit
Re ~ 0 [25].
Upon projection of the Fourier modes for the RF formulation, or the
Chandrasekhar modes for the RR formulation, a set of sixteen ODEs are
obtained, which govern the time dependent expansion coefficients. The two sets
of equations are complicated and will be given explicitly elsewhere [56]. In the
present study, each set of equations will be referred to as the sixteen-degree-of-
freedom nonlinear dynamical system (16NDS). The projection consists of
multiplying equations (1)-(2), in the narrow-gap limit, by the appropriate mode
and integrating with respect to x over the intervals [0, 1] and [- 1/2, + 1/2] for the
RF and RR formulations, respectively.
Upon elimination of higher-order axial and radial velocity, and normal stress
coefficients, the system can be reduced to the six-dimensional system derived
previously [ 18,56], which is given here for reference"

(J- V + X-aRvU,
l? - - U W + r U + Y - a R v V ,
- uv + b(Z- RvW), (8)
)( - - 5(X + aU),
]" - - b UZ - ~ WX - rX - 8 ( Y + a V ),
2 = fl(gY- r 5(Z + a W),
269

where U, V, W , X, Y and Z are time dependent and are related to the expansion
coefficients [18]. Note that a dot denotes differentiation with time. The
parameters in system (8) are given by:

1 1"
r = k 2 z'3Ta, r= 2 k 2' ~- E
+
(9)
b - 41r 2 r, ,
17= mb ( )
qo- 31r2 - k 2 r

In the double limit E, R v --~ 0, that is, in the case of a Newtonian fluid, equations
(8) reduce to the three-dimensional system derived by Kuhlmann [34], or to the
Lorenz system with the Prandtl number equal to one [ 11 ]"

O-v-u, 12 - - U W + rU - V, t~- UV - bW. (10)

It is important to observe that the nonlinear terms in the viscoelastic 16NDS stem
from both the convective terms in the momentum equations and upper-
convective terms in the constitutive equations. This is in sharp contrast to the
case of Rayleigh-Benard convection of a viscoelastic fluid [14-16] where no
nonlinear terms survive in the stress equations. The consequence of such
nonlinear terms in stress, as we shall see, is drastic on the flow behavior, as
opposed, for instance, to the cases of a convected Maxwellian fluid with less
nonlinear terms, or linear Maxwellian fluid with no nonlinear terms in stress. It is
at this point, perhaps, that one may begin to appreciate the difficulty with
analyzing viscoelastic fluids given the marked difference in flow behavior
subject to one constitutive model or the other.

3. L I N E A R S T A B I L I T Y ANALYSIS

Before proceeding with the numerical solution of the 16NDS, subject to some
appropriate initial conditions, it is useful to carry out local stability analysis in an
attempt to unravel some of the fundamental differences between Newtonian and
viscoelastic fluids. In section 2.4 the linear stability of the Couette flow was
already examined in the absence of inertia for a Maxwell fluid (Re = R v = 0).
Consider the influence of inertia on the stability of a viscoelastic fluid. For small
values of E (or De) and Rv, one expects the behavior of the flow in phase space
to be similar, in both the Newtonian and viscoelastic regimes, at least around the
fixed points. As E increases, the stability picture changes, giving rise to a
270

periodic solution, or overstability, which is the result of normal stress effects that
are usually not present in the case of a Newtonian fluid.

3.1. Stability of the circular (Couette) flow


As in the case of a Newtonian fluid, one of the steady-state solutions of the
16NDS is the origin in phase space, which corresponds to purely circulatory
motion in the azimuthal direction (base flow). As the Taylor (or Reynolds)
number exceeds a critical value, T a c , additional fixed points emerge. For a
weakly elastic fluid the stability picture is expected to be similar to that for a
Newtonian fluid, and, in the limit of a small number of modes, to the stability of
the Lorenz system [11]. One can also refer to the more complicated set of
Newtonian equations resulting from taking the double limit of the 16NDS as D e ,
R v --> O, but the resulting stability picture is essentially the same as that based on
the Lorenz equations.
In the case of the Lorenz model (10), one easily deduces that at r = r c = 1, t w o
additional fixed branches emerge, corresponding to the onset of (axisymmetric)
Taylor vortices in opposite directions"

U s = G = + [ b ( r - 1)] 1 / 2 , Ws = r - 1. (11)

For the 16NDS, the critical value of r = r c (or T a c for the Taylor number) at the
emergence of the two fixed branches is not as transparent as for equations (8) or
(10). The value of r c tends to generally decrease as the level of elasticity
increases. The value of r c corresponding to system (8) gives a reasonable
estimate of this tendency [ 18]"

rc = ~
" a or Ta c =
k2+.21,i 1 i ,
k )2 1 + a rE (12)

where the term between square brackets constitutes the Newtonian contribution
from the Lorenz system. In this case, T a c has a minimmn at
!
. . . .

lr 2 n. 2 1] 1/2
k = km = ~[aE (aE + 1) + 1 Clearly, in t h e c a s e of a Newtonian fluid,
aE
rc = 1 and k m = h a l 2 . Similar conclusions are reached when the formulae are
cast in terms of the Reynolds number, R e , and Deborah number, D e . We shall
use whatever notations are convenient as the present results will be compared
271

with those based on existing linear stability analyses. Comparison will be made
with existing formulations that use exact solutions for the eigenvalue problem,
and with the previous model that involves a lower number of eigenmodes [ 18].
Joo & Shaqfeh [27] carried out the numerical solution of the linear stability
problem, using the RR boundary conditions, and examined the influence of
inertia and elasticity for a UCM fluid. It is thus helpful to assess the accuracy of
the present approximate method against their results. Fig. 1 displays the neutral

300 I I I I I I 'I I

De - 0.0
2.5 ..... /
5.0 .....
7.5 ..........
250 10.0 .....

200

~D
150

9f,,,,\ I ..-"
"('..', ,, I" ..'"
UNSTABLE _.."'I"" ...... . .... ""'""" . ....... .'"
100 ~ "r ,, , ...-'" ..- . .......... ..
"~ \ " ", ~ i-'" .- ~ 9........... .-'~'~"

9 "\'..... ".... ..--" ,,. ..-"" . ............ . ~ . ~ . ~ "


9 "\ "... " . ".-.. _.~-'" .i" . . . - o" ........... ~ . . ~ . ~ ' ~

50

" ' , ,

I I I I I , I

1 2 3 4 5 6 7 8

Fig. 1. Influence of fluid elasticity and marginal stability curves for a UCM fluid
(Rv = 0) based on the RR boundary conditions and e = 0.01.

stability curves in the (Re, k)-plane for the onset of steady TVF for 0 < De <_ 15
and e = 0.1. Note that these curves are based on the linear stability analysis of the
16NDS. For all De values, the critical Reynolds number Rec = Re(k = k )
decreases roughly like 1/k2 near k = 0, reaches a minimum, and then increases
roughly like k2 for large k. There is a shift in the value of km towards the fight as
272

De increases, thus reflecting the greater difficulty for axisymmetric Taylor


vortices to be observed in the case of highly elastic fluids. On the other hand, this
minimum becomes less localized with increasing fluid elasticity, resulting in a
wider range of wave numbers at which Taylor vortices set in. However, for De >
10, the neutral curves show again a very localized minimum, with an
increasingly narrower range of wave numbers for the Couette flow to lose its
stability to steady TVF. For a given k value, Rec decreases generally with De
roughly like 1~De. This is, perhaps, the most important result in fig. 1" fluid
elasticity tends to precipitate the onset of axisymmetric Taylor vortices at any
value of the wave number in the axial direction. The results of Joo & Shaqfeh
(reported in [27]) display the (Re, k) neutral stability curves for the same range of
the Deborah number. Their results should be regarded as exact. They show good
agreement with the curves based on the current approximate method. The

approximate curves tend to give slightly lower values for Re c and ktn .
Neutral stability curves in the (Re, k) plane were also obtained using the RF
conditions. This enabled the assessment of the influence of the boundary
conditions on the stability picture. The RF marginal curves lead to essentially no
qualitative disagreement with the curves in fig. 1. Both figures show a localized
minimum at k -- km that increases with De, a tendency of the curves to flatten
around the minimum for the more elastic fluids, a general decrease in the critical
Reynolds number as fluid elasticity increases, and the narrowing range of
practical k values for the onset of steady TVF.
Consider now the important question whether there is any (at least qualitative)
change in the stability picture due to the inclusion of higher-order modes in the
16NDS as opposed to keeping only six modes as in system (8) [18]. A direct
quantitative comparison was carried out by examining the influence of fluid
inertia, elasticity and retardation through Re, E and Rv, respectively. Comparison
shows that the stability picture is qualitatively the same in both cases. The
inclusion of higher-order modes appears to make little difference for Newtonian
and weakly elastic fluids. In fact, the curves corresponding to small values of E
(E = 0 and 0.01) were essentially identical. Deviation between the two sets of
curves became more evident for the larger E values. This is expected, as
discussed above, since higher-order modes, previously neglected in [ 18], become
important as the level of fluid elasticity increases. In this case, the neutral
stability curves appear to be flatter as a result of the inclusion of higher-order
modes.
Similar conclusions are reached as to the influence of fluid retardation on the
critical Taylor number Tac for an Oldroyd-B fluid with E = 0.5. Fluid retardation
273

tends to delay the onset of Taylor vortices. Note that in the limit Rv ~ 0% one
recovers the Newtonian marginal stability curve (E = 0). The influence of Rv on
the loss of stability of the base flow is to be expected since retardation tends to
delay the onset of instability. Comparison shows a growing deviation and
flattening of the curves for the lower Rv values considered. Recall that as
retardation decreases, elasticity becomes relatively more dominant, and the
influence of the higher-order (normal stress) modes is expected to be stronger. It
is interesting to note that the curves in fig. 1 are reminiscent of those
corresponding to the onset of overstability obtained from the linear analysis of
Larson et al. [25] (see also fig. III-11 by Larson [20]). It is clear in both cases
that fluid retardation tends to prohibit the onset o f steady and oscillatory Taylor
vortices. Although the mechanism of onset of instability is different in the two
situations (given the absence of inertia in the analysis of Larson et al. [25]), the
role of fluid elasticity (at least for an Oldroyd-B fluid) appears to be very similar
to that of inertia. This is particularly obvious from fig. III-11 in [20], which
clearly shows the destabilizing influence of both inertia and elasticity.
Additional calculations and comparisons with existing results were also
carried out to closely examine the influence of fluid elasticity and retardation.
The results will only be commented upon here but not shown for lack of space.
The behavior of the minimum value of the critical Taylor number
Tacm(E,Rv)-Tac(k=km,E,Rv) as function of the Elasticity number and
viscosity ratio reveals a very similar trend as established previously (fig. 3 in
[18]) on the basis of model (8). It is found that Ta m drops sharply with E at
small elasticity in the case of a UCM fluid, in a manner similar to that predicted
by earlier linear analyses (see fig. 111-6 in [20] and fig. 8 in [57]) for the case of
corotating cylinders). The decrease in Ta m is considerably attenuated at the
larger E values. As to the influence of fluid retardation, there is a flattening in the
(Ta m , k) curves as Rv increases. This behavior agrees with that obtained from
the linear analysis of Thomas & Walters [58] in the case of a Maxwell fluid, and
those of Ginn & Denn [59] for a second-order fluid corresponding to a second
normal stress of zero value. In this case, the formulation based on the second-
order fluid is expected to lead to the same stability picture as that based on a
Maxwell fluid with small E value. Similar agreement is reached, as to the
influence of fluid elasticity on the value of the wave number km at which the
critical value of the Taylor number has a minimum, upon comparison with earlier
formulations [ 18].
274

3.2. Stability of the Taylor vortex flow and bifurcation


The previous section (3.1) deals with the loss of stability of the (purely
circulatory) Couette flow to steady or oscillatory TVF depending on whether
inertia is significant or negligible. In this section, attention is focused on the
stability of the TVF itself. For a Newtonian fluid, the emergence of the two
nontrivial branches at r = 1 is accompanied by an exchange of stability between
the origin and the two fixed branches via a supercritical bifurcation. An exactly
similar situation is encountered in the Rayleigh-Benard convection of a
Newtonian fluid [37], and even in the convection of a viscoelastic fluid of the
Oldroyd-B type [ 18]. For the Taylor-Couette flow of a viscoelastic fluid, the two
solution branches correspond to the nontrivial steady solution of the 16NDS. A
closed form solution such as (11) is difficult to obtain in this case because of the
nonlinear coupling in the flow field and the large number of equations involved.
The solution is thus obtained numerically using the damped Newton-Raphson
method. On the other hand, the steady-state solution of system (8) is possible to
obtain in closed form, and is written here for reference as [ 18]:

(U 2 - 5b)(U 2 - 8 ) + ( a R v + a fa+ 5 ) ( b a R v - a f a + 5 ) U 2
r ~
a ( b R v - (a- 1)U 2 + b 5(a + 5 )

U Z + (a + 5 ) r - tY
Ws = baRv - a fa + tY ' Vs = Us, (13)

Xs = -aUs, Ys = ( a R v - r - W s ) U s , Z s = aRv Ws - b

It is thus clear that the solution of the first equation in (13) leads, similarly to the
Lorenz system, to two nontrivial solution branches corresponding the onset of
Taylor vortex flow. The steady-state solution of the 16NDS also leads to the
emergence of two nontrivial branches at r = rc(E, Rv, k), which will be referred
to as C 1 and C2, that are symmetric with respect to the r, and correspond to the
onset of Taylor vortices in the two opposite directions. Unlike the case of a
Newtonian fluid, the pitchfork bifurcation at r = rc is not always supercritical. In
fact, our previous study based on system (8) shows that the bifurcation is
supercritical only for weakly elastic fluids. When the Elasticity number E
exceeds a critical value, E = Esub, the bifurcation becomes subcritical. This
bifurcation picture is confirmed when higher-order modes are included as we
shall see next.
275

For weakly elastic flows (E < E s u b ) , the 16NDS leads to an exchange of


stability similar to the case of a Newtonian fluid, which takes place at r = rc(E,
Rv, k), except that the critical value r c now depends strongly on fluid elasticity
and retardation. As was established above, rc is different from one, and becomes
increasingly smaller as E increases. At r = r c, a supercritical bifurcation
emerges. The base flow, which is stable for r < r c, loses its stability to the two
steady branches C 1 and C2 as r exceeds r c. Thus, for r > r c, the solution evolves
to either one of the two branches depending on the initial conditions. As r
increases and reaches a critical value, the two branches lose their stability as will
be seen below. This stability and bifurcation pictures remain essentially
unchanged until E exceeds E s u b , when the supercritical bifurcation gives way to
a subcritical bifurcation at r = r c.
The influence of the Elasticity number on the bifurcating branches at r = r c is
depicted in fig. 2 for a UCM fluid ( R v = 0) and k = 3. These curves correspond
the RF conditions, and should be compared to the bifurcation diagrams based on
system (8) as shown in fig. 5 of [ 18], and included here in the inset of fig. 2 for
reference. Because of symmetry with respect to the r-axis, only one branch (C 1)
is shown. The curves in the figure are obtained numerically, contrary to those
based on the analytical expressions (11) and (13). There are several important
aspects to be observed from fig. 2, and the inset therein, in comparison to the
Taylor-Couette flow of a Newtonian fluid. It is first observed that there is a
strong dependence of r c on fluid elasticity. For small E, there is a supercritical
bifurcation at r = rc, with r c becoming increasingly smaller as E increases, and
decreasing all the way to zero. At the critical value E = E s u b ( R v , k), which in this
case ( R v = 0, k = 3) is equal to 0.018, the bifurcation changes from super- to
subcritical. This corresponds to a change in concavity at (r = r c, U] = 0). In the
supercritical regime, when r < r c, a small disturbance of the base flow decays
exponentially according to linear theory. The nonlinear terms in the 16NDS
remain small in this case. As r exceeds r c, linear theory predicts an exponential
growth of a small disturbance of the base flow. This growth is, however, halted
by the stabilizing nonlinear effects.
Although linear theory may well describe the onset of secondary flow, it fails
to give the magnitude of the steady disturbance. In the subcritical regime (E >
E s u b ) , when r < r g (shown here for the curve E = 0.05), the base flow is globally
asymptotically stable [44]. When r g < r < r c, t w o bifurcation solutions exist for
the disturbance from the base flow. In this case, any disturbance below a certain
threshold decays to the origin. Above the threshold value, the base flow is
destabilized. In this range of r values, the flow is commonly defined as
metastable [60]. Whether the bifurcation at rc is subcritical or supercritical,
276

I " I !
i I I
I I I I I I .- ; E = 0.000
...............:.;.::.-..................... 0.006 .....
1.6
1.8 0.018 ......
..., ........
0 . 0 4 0 ...........
_ :.. 9
9
....
...-.
0.050 .....
1.2 '.
// 0.100 .....
1.6
0.8
.... --- .......................
1.4 .~ o.o. /.
0.4
..-'" s/.
. . - o. j . ../
1.2 0
I i I i il

0 0.2 0.4 0.6 0.8 1 1.2 1.4 ,'" E sub = 0 . 0 1 8 .-"


,," ,,,./"
,' /1"

.................................................... - ......................................................?.':" ............................................


s
0.8 ~ _ .......................... i- . . . . . . . . . . . . . . . . . . . . . . . _/- ..................

l" ', ....... i .'


I'".,, I "".. : /
0.6 ,,.~,... ... ~ /
I I ~.~ "-. : t /
<-'- .................
-., I"I .... 11.... ~ "~.
..... -.-'.. .................. ; ................ )r / ...............................
" / ..........................
- -
0.4 "-.. , , x - /
9~" ~- ~ I I ~. ", t
-.i i \ - / /
",. , .~ ",.. /
0.2 ,,. , -~ ~. I / -
, ", i ~ I
I "~I i ~
i I il ~ i I I * I I .
, I
0.2 , 0.4 , 0.6 0.8 1 1.2 .4
i' I
I

rg rh rc r

Fig. 2. Bifurcation diagrams and influence of higher-order modes for a UCM


fluid (Rv = 0) with k = 3 and e = 0.01. Super- and subcritical bifurcations for
weakly and strongly elastic flows corresponding, respectively, to E < Esub and E
> Esub. For the subcritical curves, the stable regions are indicated by breaks in
the curves and are delimited by two arrows. Inset shows bifurcation diagrams
based on lower-order theory eqs. (8) with Esub = 0.014.

the TVF loses its stability at some critical r value r h > r c (shown here for the
curve E = 0.05)
Consider the stability of the steady branches C 1 and C2 as r is increased from
r c. It is helpful, however, to first recall the situation for a Newtonian fluid. The
discussion is focused on the Lorenz system (10) since the inclusion of higher-
order mode does not alter the qualitative picture. Linearization of the Lorenz
equations around the steady-state solution leads to a characteristic equation of
the third degree. The conditions for the loss of stability of the toroidal TVF is of
particular interest here. For a Newtonian fluid, the two nontrivial fixed points C 1
277

and C2 are always sinks for any value of r. At r = 1, a pitchfork bifurcation


occurs, while the origin remains a saddle point with a one-dimensional unstable
manifold. Note that this is exactly the same situation for the Lorenz equations
with the Prandtl number equal to one. In this case, no Hopf bifurcation occurs
since the characteristic equation does not possess a pair of purely imaginary roots
[61 ]. The roots have a real part that remains negative for any value of r. Thus, for
r > r c = 1, the three-dimensional Newtonian model cannot predict the
destabilization of the TVF (through a Hopf bifurcation, for instance), nor can it
entertain a chaotic solution. This situation remains unchanged when higher-order
modes are retained in the solution.
The presence of fluid elasticity in the constitutive equations appears to be a
sufficient condition for the existence of a Hopf bifurcation. Indeed, at r = r h ( E ,
R v , k) > r e, linear stability analysis around C 1 and C2 indicates that these two
branches lose their stability via a Hopf bifurcation emerging at r - r h. In fig. 2,
sub
the range of r values for which the branch C 1 is stable (for E > E ) is indicated
by breaks in the lines between two arrows. It is observed from the figure, that for
highly elastic fluids (E > 1), the range of stability of the C1 and C2 branches
decreases to zero, so that the (steady) TVF is unstable for any postcritical r value.
While the existence of the Hopf bifurcation may not be difficult to establish, the
investigation of its stability can be algebraically quite involved [62]. The
difficulty stems from the hyperbolicity of the fixed point at r = rh, and center
manifold theory must be applied in order to determine the stability of the fixed
point [63]. The situation is similar for r near r o The stability picture will be thus
established through the numerical solutions of the 16NDS.

4. FINITE A M P L I T U D E TVF AND C O M P A R I S O N W I T H E X P E R I M E N T

Linear stability analysis, such as the one presented in the previous section,
determines the flow field as it is slightly perturbed from the base flow or from
the (steady) TVF. However, it fails to give the flow structure for a large
disturbance. The influence of the nonlinear terms must thus be examined through
the numerical solution of the 16NDS. The influence of inertia and elasticity on
weakly and strongly elastic flows is examined in some detail. A weakly
(strongly) elastic flow is defined as one whose Elasticity number is small (large)
enough for it to undergo a supercritical (subcritical) bifurcation at the onset of
Taylor vortices. Particularly, the emergence of super- and subcritical Taylor
vortices, as well as the onset of chaos, are investigated by determining the flow
spatio-temporal structure in phase space, through power spectra and/or time
278

signatures, whichever representation is most insightful. The present nonlinear


formulation is assessed against experiments for a weakly elastic flow and a flow
with negligible inertia.

4.1. Influence of inertia on weakly elastic flow


The influence of inertia on a moderately weakly elastic Oldroyd-B fluid is
examined for a fixed value of E by varying the value of D e (or, equivalently, Re).
This is equivalent to fixing the level of fluid elasticity and increasing the shear
rate. Baumert & Muller [9] performed experiments on the flow visualization of
the Taylor-Couette flow of dilute solutions of high molecular weight
polyisobutylene in oligometrie polybutene. Rheologieal measurements of these
solutions confirmed that they are of the Boger fluid type (highly elastic with
constant viscosity). They monitored the temporal evolution of TVF over a range
of shear rates using reflective mica platelet seeding. The transition to steady TVF
was observed to occur with or without oscillatory flow depending on whether
elasticity was dominant or not. In this section, the calculations are carried out for
a weakly elastic flow of moderately small inertia (Re .~ 100), and the transient
flow behavior is compared with the observations of Baumert & Muller [9].
The aim of the present comparison is to reproduce some of the experimental
observations as the base flow loses its stability to TVF. Given the uncertainty in
the experimental values of the relaxation time and the wave number, and the
assumptions adopted in the derivation of the low-dimensional dynamical system,
only a qualitative agreement can be expected. The results are also compared with
the numerical calculations of Avgousti et al. [28] whenever possible. The
calculations performed in this section are based on the rigid-flee boundary
conditions. For a fixed level of fluid elasticity (E fixed), experiment suggests that
transition from the purely azimuthal flow to steady TVF occurs once the shear
rate reaches a critical value, i.e., D e = D e c (or Re = R e ) . The transition from
simple sheafing to the large toroidal axisymmetrie vortices, which span the gap,
is monotonic with time as D e first exceeds D e c. At higher D e value, an
oscillatory secondary flow is observed prior to the formation of steady vortices.
Axially migrating mica alignment bands begin to appear, but the band formation
ceases with steady vortices setting in after some time. At this point, steady TVF
is observed to persist for a long time.
Consider the ease of a moderately weakly elastic fluid with E = 0.16. This
corresponds roughly to the low-viscosity Boger fluid used in the experiment of
Baumert & Muller [9], with p = 0.856 g/c m3, rls = 2.6 poise and I"1 = 3.0 poise.
279

The value of the relaxation time ~1 was found to be equal to Ls = 0.021 s or ~t--
0.11 s (with the corresponding Elasticity number equal to E = 0.164 or E t =
0.94), depending on whether steady shear or transient experiments were used [9].
The value of the viscosity ratio is set equal to R v = 6.5. Figure 3 shows the time
evolution of the (dimensional) axial velocity component U z at z = k/2, where it is
maximum, and x = 0.5, midway through the gap. The evolution of the velocity is
displayed for various values of the shear rate: 13 < D e < 14 (81.25 < R e < 87.5),
above the critical shear rate, D e c ~ 13 (Re c ~ 81.25). The curves in the figure are
similar to the results reported in fig. 1 of Avgousti et al. [28], and the trend is in
agreement with the observations of Baumert & Muller [9] for low and medium

0.01

0.009 -

9 ..../ ....................................................................................................
0.008 -

0.007- "''".... ................................................................................................................................................................................................

0.006-

0.005 -
: /,'- ....
: i
i /
0.004-
: tI

0.003 -

De = 13.09
13.18 .....
0.002
13.27 .....
i 9 , ,i 13.49 ..........
i :; ,' / 13.70 .....
ii ," ,'
0.001 J i / ,'" /

o I !
0 5;0 10b0 15;0 2000 25'00 30'00 3500

Time (s)

Fig. 3. Monotonic time evolution of the axial velocity for different shear rates
(13 < D e < 14) and small initial perturbation of the base flow (E = 0.16, R v = 6.5,
k = 3.5 and e = 0.093).
280

viscosity fluids. The initial condition taken for all the curves in fig. 3 is U1 = 0.01
with the remaining variables equal to zero. In this case, the transition from the
base flow to TVF occurs via a supercritical bifurcation similar to the branches E
< 0.018 in fig. 2 for a UCM fluid. In the precritical range (De < Dec), the origin
in phase space (base flow) is asymptotically stable to any perturbation, and the
numerical integration of the 16NDS, in the vicinity of the origin, follows simply
an oscillatory decay to the origin. As De (Re) exceeds Dec (Rec), an exchange of
stability occurs between the origin and the steady-state branch C1 (or C2),
coinciding with the emergence of steady TVF, with increasing amplitude as De
(Re) increases.

100

Theory +

+ +
+
...., -i-
[-.,

Experiment

10 100

De

Fig. 4. Time after shear initiation of appearance of steady vortices as function of


the shear rate (De). The values in the figure are based on the curves in fig. 3. The
inset shows the experimental measurements of Baumart & Muller [9] for a
medium viscosity fluid.
281

Typically, the flow evolves monotonically from the base flow to steady TVF
over a period of time that decreases with increasing shear rate as depicted from
fig. 3. This appears to be in agreement with the observations of Baumert &
Muller [9] who monitored the dependence of onset time on the shear rate for a
fluid with medium viscosity. Figure 4 displays the time after shear initiation of
appearance of steady vortices as a function of the Deborah number (shear rate).
An inset showing the values from experiment is also included for comparison,
and reveals good qualitative agreement with theoretical predictions. The values
in the inset are based on measurements performed for a medium viscosity fluid,
with E = 44. As will be seen next, at higher shear rates, however, theory predicts
that the time TVF takes to set in begins to increase with De, a trend in
disagreement with what experiment suggests.

0.o45[ '~ . . . . . . . .

o li ,

0.0as
1 '11 " !

o.o i ii
0.0254 i1!/ it i :V~i :.\ i ,,:' '.. ..-,,
7h | 1 i ! l, i i :,i" :", ".,,':'.2.~,vc.':-.-~'c:- "----:---- ..... " ......... -............ ---"- .....

~ o.o2 I

1 i. i De = 17.47
0 9005 ' ".,.,i 17 81
17194

o 5~o lo'oo ~;oo 2000

Time (s)

Fig. 5. Oscillatory time evolution of the axial velocity for different shear rates
(17 < D e < 18) and small initial perturbation of the base flow. The parameters
used are the same as in fig. 3.
282

0.028 -
(a)

0.026: 18.13
_ J

0.024-
16.78
0.022

0.02 De = 16.07

0.0 8

. , . -- ,. -

0.028
(b)
0.027 -
19.14
q7 0.026- [hh. 18.89
De = 18.77 11111#~r162 .... -

0.025 -

0.024-

.... 19.38 . . . . . . i
0.08- (c)
0.04-

0-

-0.04 -

-0.08 - i
..... 1 .... 010 "'! I I " I !
-0 2000 4000 6 0 8000 10000 12000 14000 16000 18000

Time (s)

Fig. 6. Time evolution of the axial velocity for 16 < De < 21. The flow for each
shear rate level is obtained using the flow of the previous (lower) level as initial
conditions. The parameters used are the same as in fig. 3. The figure is
subdivided into three segments of 18,000 s period each. Steady TVF is always
reached during the first period (a) for the range 16 < De < 18.7. Oscillatory
instability develops towards the end of the second period (b) for the range 18.6 <
De < 19.27. Transition to periodic behavior some time during the third period (c)
for the range 19.26 < De < 21.
283

The evolution of the flow loses its monotonicity as shearing exceeds a certain
level. This is depicted in fig. 3 for the curves corresponding to De > 13.2, which
show a local depression in the flow amplitude before steady TVF sets in. In fact,
as De is further increased, an oscillatory behavior emerges before the onset of
steady vortices as shown in fig. 5 for 17 < De < 18. The secondary flow gains
strength initially at a much faster rate than for the lower range of shear rates (fig.
3). The amplitude of oscillation decays with time until the onset of steady of
TVF, but the frequency appears to remain constant. As De increases, the overall
amplitude of oscillation increases, accompanied by a phase shift and delay in the
onset of steady TVF. Eventually, oscillatory behavior is sustained for longer time
for higher De values; in the limit, no steady TVF is found.
Before examining the onset of sustained oscillations, it is desirable to address
the issue of the influence of the initial conditions on the ensuing flow behavior.
Baumert & Muller [9] report that "flow behavior was not found to be particularly
sensitive to the type of velocity ramp imposed" initially. They indicate that the
flow behavior commencing at the start of the final velocity was indistinguishable
from the case when a plateau was imposed over a certain period. This
observation is presently confirmed through additional calculations carried out by
taking as initial conditions the steady TVF reached before increasing the shear
rate, instead of starting again from near the origin as in figs. 3 and 5. The
resulting sequence of initial flow transients, after the imposition of the higher
(new) shear rate, and subsequent final (steady or oscillatory) TVF are depicted in
fig. 6 for the range 16 < De < 21. The figure displays the evolution of flow over a
period of 54,000 s, subdivided into three equal periods of 18,000 s.
The first segment is shown in fig. 6a for the range of shear rates corresponding
to 16 < De < 19. Comparison between the curves in fig. 5 and those in fig. 6a
indicates that the flow behavior is essentially unaffected by initial conditions. It
is interesting to note the similarity in the emergence and decay of oscillatory
behavior between one steady level and the other. The figure clearly shows that
the time it takes steady TVF to set in increases with shear rate. This is a result of
the oscillatory transition that takes a longer time for the higher shear rate. This is
indeed even more evident from fig. 6b, which shows, for the range 18 < De < 20,
that steady TVF does not even set in over the time period a given shear rate is
applied. The oscillations tend to decay at a slower rate as De increases, and
eventually start to grow with time as indicated in the last portion of fig. 6b,
corresponding to De = 19.26. This growth, which is linear upon onset, does not
remain unstable over time as linear stability analysis suggests. At some time,
when the signal amplitude is large enough for nonlinear effects to become
significant and halt the (exponential) growth.
284

The bifurcation diagrams in fig. 2 suggest that steady TVF is reached as long
as r is below rh(E, Rv, k). This corresponds to the onset of a Hopf bifurcation at
De - Deh when steady TVF loses its stability to periodic behavior. For the
present flow in fig. 6, Deh is found to be approximately equal to 19. The
existence of a Hopf bifurcation at De = Deh is easily confirmed through linear
stability analysis around the nontrivial steady-state solution(s) when two
eigenvalues are purely imaginary, and no other eigenvalue possesses a zero real
part. The stability of the Hopf bifurcation, however, is difficult to establish for a
complex system like the 16NDS. Such an analysis is algebraically involved, but
a similar analysis, based on center manifold theory, was carried out previously
for the thermal convection of viscoelastic fluids [15]. Numerical integration, on
the other hand, shows that the Hopf bifurcation indeed exists, and is stable as
will be seen below. This situation is in sharp contrast to the case of the Lorenz
system (10) whereby no loss of stability of the steady TVF is predicted.
In figs. 6a and b, and for the range 16 < De < 19.15, the computation has been
limited to flows with final state, after transients have died out, that corresponds
to a standing wave. This also corresponds to most of the range of shear rates
examined in the experiment of Baumert & Muller [9] that led to steady TVF for
the low and medium viscosity fluids. However, they reported the observation of
coexistence of migrating bands and distorted vortices at the highest shear rate
considered. It is thus desirable to numerically examine the transition from steady
to oscillatory TVF. The sequence of flows displayed in fig. 6c is obtained for De
above Deh subject to the initial condition corresponding to the steady TVF at De
slightly below Deh. Avgousti et al. [28] reported having numerical difficulties
for a UCM fluid when they took the slightly perturbed standing wave as their
initial condition. Figure 6a shows the transition from steady to oscillatory TVF,
that is, from the flow at De = 19.26 to that at De = 19.38. The figure also shows
the transition to another oscillatory state at De = 20.56. The first transition is
preceded by a destabilization of the steady TVF to a traveling wave of amplitude
(exponentially) growing with time, which is reminiscent of the behavior reported
in figs. 6 and 9 of Avgousti et al. [28]. The exponential growth (instability)
ceases immediately upon increase of the shear rate, giving way to a regular and
stable oscillatory vortex structure. This structure is sustained for as long as the
shear rate is maintained at De = 19.38. Upon further increase of De, another
regular oscillatory structure sets in with a larger amplitude. The modulation of
the oscillation indicates the emergence of period doubling. Additional theoretical
and experimental investigations are obviously needed if the flow at still high
shear rate is to be explored. It is possible that the flow may not remain
285

axisymmetric, a fact that does not seem to be suggested by the experiments of


Baumert & Muller [9].

4.2. Weakly inertial flow and influence of boundary conditions


The nonlinear behavior is further examined by focusing the calculations on the
influence of fluid elasticity in the presence of very weak inertial effect. One of
the most interesting phenomena encountered in the Taylor-Couette flow of a
viscoelastic fluid is the emergence of overstability that is attributed to fluid
elasticity, which is otherwise absent in the case of a Newtonian fluid. The
existence of such a purely elastic overstable mode was proved through linear
stability analysis [25] and numerical calculations [30] carried out in the absence
of inertia. Recently, Muller et al. [8] conducted Laser Doppler velocimetry
(LDV) measurements of the axial velocity component of a Boger fluid between
two concentric cylinders (~ = 0.0625), with the outer cylinder being at rest, and
the inner cylinder rotating at constant angular velocity. The measurements show
an oscillatory flow at a vanishingly small Reynolds number. The flow appears to
undergo a transition from the purely azimuthal Couette flow to time periodic
flow as the Deborah number De exceeds a critical value, Dec, which is in good
agreement with the value predicted by earlier linear stability analysis of an
Oldroyd-B fluid [26]. The LDV measurements show that the oscillatory behavior
is not localized but appears to be spread throughout the flow. As the Deborah
number increases from the critical value, the amplitude of oscillation increases
like (De - Dec) 1/2. The corresponding power density spectra show peaks, which
are instrumentally sharp at the fundamental frequency, the growth of harmonics,
and eventually subharmonics, reflecting, perhaps, the presence of a period
doubling or quasiperiodic motion.
A similar sequence of flow behaviors is also obtained from the finite-element
calculations (with inertia neglected) of Northey et al. [30] for a UCM fluid.
These authors, however, report having numerical difficulties in obtaining the
solution at the higher Deborah numbers (possibly coinciding with the onset of
period doubling). Their calculations are thus limited to an extremely small range
of postcritical Deborah numbers. As discussed earlier, the system (8) used in our
previous study [18] could not possibly reproduce any of the reported
experimental results for vanishingly low-Reynolds number. This fact is easily
verified if the inertia terms are dropped from system (8). In this case, the
acceleration terms on the left-hand side, together with the curvature term (V) on
the fight-hand side of the first equation, are neglected. It is then easy to deduce
that one recovers the trivial solution or the purely azimuthal flow. Thus, the level
of truncation adopted in the previous model makes the formulation inadequate
286

for the investigation of the purely elastic overstability or even a flow at very
small Reynolds number. In general, however, the presence of inertia, no matter
how small it may be, prohibits the base flow from losing its stability to the
overstable mode. Instead, the base flow loses its stability first to steady (and not
oscillatory) TVF since there is always a finite range of r values over which the
branches C 1 and C2 are stable (see fig. 2).
The influence of fluid elasticity is now examined for vanishingly small inertia
in the postcritical range of the Deborah number, in an attempt to recover the flow
sequence observed in the experiment of Muller et al. [8]. One cannot expect full
quantitative agreement between theory and experiment given the level of
approximation in the 16NDS. Moreover, the wave number k needs to be imposed
in the present calculation. This quantity is not known from experiment; the wave
number is difficult to establish under unsteady conditions of flow [9]. The
comparison between theory and experiment is covered next for the RF and RR
formulations.
Consider the flow with negligible inertia, and set r = 10-6. Also let Rv = 3.75
and e = 0.0625 corresponding, respectively, to the viscosity ratio of the fluid and
the gap-to-radius ratio used in the experiment [8]. Although the wave number is
not specified in the experiment, it is set k = 4 based on wave numbers reported in
other experiments on Taylor-Couette flow of viscoelastic fluids [9]. Thus, only
the Deborah number will be varied. The flow is examined as De is increased
from zero, that is from the Newtonian level. Referring to fig. 2, it is seen that for
De = E = 0 the base flow is unconditionally stable to any perturbation since r <<
1 in the present case. This situation remains practically unchanged until De
reaches the theoretically predicted critical value, Dec = 29.7 or 29.3, depending
on whether the RF or RR boundary conditions are used. These values are based
on the purely elastic linear stability analysis. At this point, the base flow is
supposed to lose its stability, but this is numerically detected at slightly higher
De values. Since inertia effects are small (r = 10-6), one observes an exchange of
stability between the base flow and oscillatory TVF, since no steady TVF can set
in in the absence of inertia. As the Deborah number is increased beyond the
critical value, the amplitude of oscillation increases. The resulting sequence of
flows is identified through the Fourier spectra shown in figs. 7a and b for the
calculation based on the RF and RR boundary conditions, respectively. The
range of Deborah numbers studied is the same as in the experiment: 32 < De <
53, and should be compared with that reported in figs. 7-10 of Muller et al. [8],
where it is noted that De = 0. 911DeM~lle~.
288

Theoretically, oscillatory TVF is expected to emerge at De = 29.3 for the RF


formulation; in practice, however, it is detected at a higher Deborah number. At
De = 33, the corresponding Fourier spectrum shown in fig. 7a clearly displays
periodic motion after the base flow becomes unstable. This periodic behavior
persists as De increases. The flow oscillates around the origin (base flow), with
an amplitude that increases as De increases. The spectrum shows the
fundamental frequency and six odd harmonics. Additional harmonics emerge
upon further increase of the Deborah number as shown for De = 39.7. Further
increase in the elasticity level (De = 41.2) leads (most apparently) to a doubling
in the period as depicted from the figure. It is not easy, in fact, to establish
whether one is dealing with a bifurcation into period-2 motion or quasiperiodic
behavior on a toms. At De = 43.57, the power spectrum indicate clearly the
presence of chaotic motion. A very similar sequence of flows is obtained on the
basis of the RR formulation. The qualitative similarity is obvious from fig. 7b.
Although the two formulations predict that the oscillatory TVF sets in at
essentially the same critical Deborah number (29.3 vs. 29.7, for the RR and RF
conditions, respectively), the emergence of higher harmonics, subharmonics and
chaos occurs at higher De values for the RR formulations (fig. 7b). This
confirms, once again, that stick conditions tend to delay destabilization.
The sequence of flows predicted by the present model is clearly comparable to
that reported by Muller et al. [8]. Experiment predicts a loss of stability of the
origin at a critical Deborah equal to 32.3 while the present formulation gives Dec
= 29.7. Direct quantitative comparison of the actual velocity amplitude is
impossible since not all experimental parameters are available. It is possible that
the experiment was in fact not conducted at fixed Reynolds number, and that the
Deborah number was increased by increasing f2. Both theory and experiment
predict the increase in amplitude of the velocity signal, the emergence of higher
harmonics in the Fourier spectrum, and eventually that of subharmonics
(corresponding to the onset of period doubling) as De increases.
Figure 8 shows the bifurcation diagrams for the square of the velocity
amplitude based on the present model and the measurements (inset) from [8]. In
the figure and inset, numerical calculation and the experimental measurements
show the onset of periodic motion as the Deborah number exceeds the critical
value. Both sets of data show that the amplitude of oscillation grows like (De -
Dec) 1/2, in agreement with the prediction based on asymptotic analysis in the
limit De --~ Dec. Although the qualitative agreement between theory and
experiment is obvious, the quantitative comparison is also encouraging,
especially for the RR results, given the assumptions adopted in the present
formulation: level of truncation involved, axisymmetric flow, narrow-gap
289

0.14- 0.01
Experiment
0.008

0.006
0.12
O
0.004

0.002
0.1-
0
20 2; 3;

0.08 -

0.06-

0.04-

0.02-

!
020 2'5 30 35 40 45 5'o 55

De

Fig. 8. Bifurcation diagrams based on the present model (RF and RR


formulations) and the square of the axial velocity component amplitude. Inset
showing the experimental measurements of Muller et al. [8].

approximation, the constitutive model, and, very importantly, the uncertainty in


the value of the wave number. The subject of elastic overstability is an important
phenomenon in viscoelastic fluids, and remains somewhat unexplored in the
nonlinear regime.

4.4. Purely elastic overstability


The experimental measurements of Muller et al. [8] were conducted under
conditions of vanishingly small inertia. A closer quantitative comparison
between theory and experiment is obtained when a modified twenty-mode
dynamical system (20NDS) is used, which accounts for even higher normal stress
effects than in the 16NDS, with the Reynolds number set equal to zero [ 19]. The
most influential normal stress modes are carefully selected to ensure that the
290

relevant dynamics is captured by the approximate model and solution. This is


first done by referring to the results from linear stability analysis as above.
Comparison between the approximate and exact solutions leads to good
agreement, especially in the lower wave number range. The model takes into
account more effectively the influence of normal stresses (which lead to the
Weissenberg rod-climbing phenomenon), and is thus adequate for the flow of a
highly elastic fluid (with negligible inertia), thus allowing direct comparison with
the experiment of Muller et al. [8].
Not all needed experimental flow parameters were explicitly reported in ref.
[8]. The test fluid used in the experiment has a (constant) viscosity 11 = 162 Pa s,
and consists of 1000 ppm of a high molecular weight polyisobutylene, dissolved
in a viscous, low molecular weight polybutene of viscosity rls = 128 Pa s, so that
the solvent-to-polymer viscosity ratio Rv - 3.76. The fluid relaxation time )~
varies depending on the rheological technique used to measure it. Steady shear
flow data give )~ = 3.3 s, while transient relaxation experiments lead to 10.9 s [8,
9]. The inner and outer cylinder radii were 8 and 8.5 cm, respectively, so that ~ =
0.0625. Although the inner cylinder angular velocity f~ was not explicitly given
in the experiment, its value can still be inferred from the values of the
experimental Deborah number, DeM, which was introduced by Muller et al. [4]
2 (1 +
as: DeM - (1 + ~2 _ 1 . Note that De M = De. It appears that there was only one
~--~0

fluid used throughout the experiment, and DeM was probably varied by varying
only the inner cylinder speed, f~. Hence, from the range of DeM values reported,
the corresponding value of the inner cylinder speed can be given by f~ =
DeM/77.06 for ~. - 4.4 s. Muller et al. [8] reported that the highest Reynolds
number, Re, reached in the experiment was of the order 7x 10-3. Indeed, if one
considers the value of f~ corresponding to the highest Deborah number reported
(De M = 54.5), one finds that Re = 2.86 x 10-3 (assuming the density 9 ~ 1
g/cm3). The experimental wave number, k (in units of d), at which overstability
is first observed, was also not reported by Mullet et al. [8]; its measurement may
have been difficult under transient conditions. Its exact value, however, is not
crucial in this case since the critical Deborah number for the onset of
overstability does not depend strongly on the wave number, over a wide range of
practical values: k e [4, 8] for Rv = 3.76 as linear analysis suggests [20]; this is
reflected by the flattening of the corresponding neutral stability curve in the (De,
k) plane around the critical value Dec. The wave number will be fixed to k = 4.85
for all subsequent calculations. This is the minimum value of the wave number
291

that corresponds to De c = 32 as predicted by the linear stability analysis based of


the twenty-mode model. This value is also close to wave numbers reported in
other experiments on TC flow of viscoelastic fluids [20]. Thus, as in the
experiment of Muller et al. [8], only the Deborah number will be varied (by
varying f2) in the following calculations and results.
Consider the flow as De is increased from zero, that is from the Newtonian
level. Linear stability analysis indicates that the Couette flow is unconditionally
stable for De < Dec = 32. In the absence of inertia, an exchange of stability takes
place between the circular Couette flow and oscillatory TVF at the critical
Deborah number, since no steady TVF can set in. This is in contrast to a
Newtonian fluid in which case only steady TVF sets in at the critical Reynolds
number. The main variable of interest is the axial velocity component, Uz(r,z,t),
which is obtained from the solution of the twenty-mode nonlinear dynamical
system. Here r and z are, respectively, the radial and axial coordinates. In the
experiment [8], Uz was measured at the point located midway through the gap,
and, probably, where it is maximum over a wavelength in the axial direction.

Thus, the amplitude of uz(r= -Ri- - +


- ~Ro
,z d ) will be monitored next.
= -~-,t
The experimental critical value of the Deborah number, at which oscillatory
motion was first detected, was reported to be equal to 32.3, and happens to be
slightly larger than the theoretical value De c = 32 predicted by the present linear
stability analysis. As De is increased beyond the critical value, the amplitude of
oscillation increases, confirming the existence and the stability of the Hopf
bifurcation, in agreement with experiment [8]. Calculations are carried out for
the same range of Deborah numbers as in the experiment: 32 < De < 50. At De =
32.5, the velocity signature and corresponding Fourier spectrum display periodic
motion after the purely circular (Couette) flow becomes unstable. The amplitude
of oscillation remains relatively small (0.008 cm/s). The power spectrum
indicates the presence of a dominant frequency of 0.02 Hz and a weak second
harmonics. This periodic behavior persists as De increases, with the flow
oscillating around the origin (Couette flow). At De = 43.57, the motion remains
periodic around the origin, with an increase in amplitude to 0.052 cm/s. There is
an increase in the fundamental frequency to 0.0298 Hz and the emergence of four
significant even and odd harmonics. This trend persists as De is further increased
with the eventual emergence of additional harmonics.
Figure 9 shows the Hopf bifurcation for the square of the velocity amplitude
based on the twenty-mode model 20NDS, and the measurements from [8]. Both
experiment and theory suggest that the amplitude of oscillation grows like (De -
292

0.01 9 ! !

0.009 -
/t
Theory ,

0.008 -

,s
/
/'
0.007 -
9

O 9
0.006
O

0.005 -
,,'" Experiment
II
eq
ii s
i
i
I
it"
0.004 - i
i
,/

t
9"
0.003 - i

0.002 -
/

0.001 - r
,/

0 I I 9 I I I I ..... I

20 25 30 35 40 45 50 55 60

De

Fig. 9. Bifurcation diagram and comparison between theory (dashed line) and
experimental measurements (diamonds) of Muller et al. [8].

Dec) 1/2 in agreement with the prediction based on asymptotic analysis in the
limit D e --+ D e c. Figure 10 displays the dependence of the dominant frequency
and its harmonics on the Deborah number. The frequency tends to increase with
De almost linearly. Unlike the amplitude, the frequency exhibits a jump at the
critical Deborah number. This means that any initial weak velocity amplitude at
the onset of oscillatory TVF has a dominant frequency that is relatively easy to
detect. The agreement between the computed and measured frequencies is
obvious from the figure. The apparent growing disagreement for the higher
harmonics is to be expected. Any initial discrepancy at the dominant frequency
level is simply amplified as it is multiplied by two for the second harmonics, by
three for the third harmonics and so on.
A closer quantitative agreement between theory and experiment can hardly
be envisaged given, on the one hand, the uncertainty surrounding experimental
293

0.25 ' I I I I =' I I I

fl ~
f2 + .....
f3 [] ......
f4 x ...........
0.2 -

E T
x h X ....................
p e
e o X ...............
r r X .....................
l y
0.15 - m X ..........
e
"4" n
t ............................................................. []o.OO...o- . . I'. r.l. . . ..oo[] o o.o.--'"~176
b . o.O
.o~o.-OO~176
o
0.1 -
oO~176

..o" ........

0.05 -

I , I ,, l l_ I _. I I
0 "'
20 25 30 35 40 45 50 55 60

De

Fig. 10. Frequency diagram and comparison between theory (dashed line) and
experimental measurements (diamonds) of Muller et al. [8]. The figure shows
the square of the amplitude of axial velocity at x = 88 as a function of De.

conditions, and, on the other, the lack of a universal and accurate constitutive
model for viscoelastic fluids. The sources of discrepancy between theory and
experiment are related to limitations of both Newtonian and viscoelastic flow
formulations. The lack of a theory capable of predicting the value of the axial
wave number, k, constitutes a major difficulty. The prediction of the value of k
remains an unresolved issue (in a given formulation, it is usually simply imposed
from experimental observation). In the case of viscoelastic Taylor-Couette flow,
however, the measurement of k is difficult under transient flow conditions [9].
Other parameters and variables are also difficult to obtain from the experiment of
Muller et aL [9] and had to be deduced. Additional uncertainty originates from
the type of constitutive model used. Although the Oldroyd-B equation predicts
the behavior of constant viscosity highly elastic fluids, it does not incorporate the
spectrum of relaxation times that is characteristic of real fluids. More
294

complicated constitutive equations, accounting for the nonlinear dependence of


the transport coefficients on the rate-of-strain tensor, may also be examined. The
present formulation accounts for nonlinearities stemming from the upper-
convective terms in the constitutive equation. Another source of discrepancy can
come from end effects in the Taylor-Couette apparatus that have been neglected
in the present formulation. The narrow-gap approximation is also a limiting
assumption. Inertia effects can also play an influential role despite the fact that
the experiment was conducted at a vanishingly small Reynolds number (Re < 1O
2). In general [18], the presence of inertia, no matter how small it may be,
prohibits the base flow from losing its stability to the overstable mode. Instead,
the base flow loses its stability first to steady (and not oscillatory) TVF since
there is always a finite range of Re values over which the branches corresponding
to steady TVF are stable.

5. DISCUSSION AND CONCLUDING REMARKS

A low-dimensional dynamical system approach is proposed for the simulation


of the Taylor-Couette flow in the narrow-gap limit of a viscoelastic fluid of the
Oldroyd-B type. Among the clear advantages of introducing such a severely
truncated representation for the flow field are the lower cost involved in the
computation, and, more importantly, the increased possibility of obtaining the
actual flow field under almost any condition without encountering numerical
instability.
The rather elementary Oldroyd-B constitutive model is adopted in the study
for two main reasons. First, since the aim of the study is to examine the influence
of elasticity on the onset and stability of TVF, the use of the Oldroyd-B model
becomes justified for a class of the so-called Boger fluids, for which the viscosity
remains sensibly constant over a wide range of shear rates, with the
corresponding normal stress levels nevertheless substantial. A polyacrilamide
solution in a maltose syrup/water mixture typically constitutes such a fluid [63].
Second, the Oldroyd-B constitutive equation is one of the simplest viscoelastic
laws that accounts for normal stress effects (leading to the Weissenberg rod-
climbing phenomenon). Other more realistic phenomenological [43,64] or
molecular-theory based models [65-67] are more difficult to handle, and are
likely to lead to a stability picture different from the one predicted by the present
analysis. For instance, the presence of shear thinning, not accounted for by the
Oldroyd-B equation, will likely have a destabilizing effect since the effective
Reynolds number increases as the viscosity decreases with increasing shear rate
[17,33]. Other more complicated constitutive equations, accounting for the
295

nonlinear dependence of the transport coefficients on the rate-of-strain tensor,


may also be examined. The present formulation accounts for nonlinearities
stemming from inertia in the momentum equation, and the convective and upper-
convective terms in the constitutive equation.
The nonlinear dynamical system is derived by expanding the flow field in
double Fourier series in space for the rigid-free (RF) boundary conditions, and
Fourier/Chandrasekhar series for the rigid-rigid (RR) conditions. The set of
ordinary differential equations, governing the time-dependent coefficients, is
obtained after applying the Galerkin projection onto the various modes, and
adopting a suitable truncation to close the hierarchy of equations. The more
severe truncation level used in the previous work [ 18] led to system (8), derived
by neglecting higher-order normal stress terms. These terms, however, tend to
become significant for highly elastic flows or in the limit Re ~ O.
The present 16NDS takes into account more effectively the influence of
normal stresses; thus being more adequate for a flow in the limit of vanishingly
small Reynolds number, and allowing direct comparison with the experiment of
Muller et al. [8]. This also allows the assessment of the influence of the higher-
order modes on the solution by comparing the solutions based on the two models
in the linear and nonlinear regimes (see below). If the higher-order terms are
neglected, the resulting system (8) reduces, in the double limit E, Re ~ O, to the
three-dimensional system (10) derived by Kuhlmann et al. [34,35] for the
Taylor-Couette flow of a Newtonian fluid, or, equivalently, to the Lorenz
equations with the Prandtl number set equal to one. The presence of the higher-
order modes is found to make little difference for weakly elastic and Newtonian
fluids, but is significant for highly elastic fluids.
Regarding the boundary conditions, rigid boundaries are assumed in the
azimuthal direction, while free (slip) boundary conditions are taken along the
axial direction for the RF formulation, and rigid (stick) conditions along the axial
direction for the RR formulation. The RF conditions give a somewhat less
realistic value for the critical Reynolds number in the case of a Newtonian fluid,
but a more accurate estimate for the torque [34]. The use of one type of boundary
conditions or the other does not seem to alter qualitatively the stability picture.
Despite the severity of the level of truncation, both formulations yield good
agreement with existing linear stability analyses and experiments in the nonlinear
regime.
A judicious selection process of the most influential modes is carried out to
ensure that the relevant dynamics is captured by the approximate model 16NDS
and solution. This is first done by referring to the results based on linear stability
analysis. The exact and approximate solutions are obtained for the linear
296

problem, in the absence of inertia, using the direct method. In this case, the
eigenvalue problem is relatively simple to solve since only constant coefficients
are involved. Only two modes, given by (6), are retained in the approximate
solution, leading to good agreement with the exact solution, especially in the
lower wave number range. Comparison between the approximate solutions based
on rigid-rigid and rigid-flee boundary conditions indicates that the type of
boundary conditions used does not influence the overall stability picture. This
confirms the earlier conclusions reached in the case of a Newtonian fluid [34]. In
fact, most earlier results based on existing linear analyses are recovered (see fig.
2).
It is found that the critical Taylor number, Tao at the onset of axisymmetric
Taylor vortices, not only depends on the disturbance wave number, but also
decreases with increasing level of elasticity. At Ta - Ta c (r = rc), two
bifurcating steady-state solution branches C1 and C2 emerge, corresponding to
TVF in two opposite directions. These solutions lose their stability when the
Taylor number exceeds a critical value Tah (r = rh), coinciding with the birth of
a Hopf bifurcation. It is found that Tah also decreases with E. Thus, at least in
the absence of shear thinning, fluid elasticity appears to destabilize the base
flow, precipitating the onset o f Taylor vortices, which, in turn, tend to get
destabilized, leading (eventually) to chaotic motion, earlier in comparison with
the ease of a Newtonian fluid. The stability and bifurcation pictures are
summarized in fig. 2 for a UCM fluid.
The validity of the model is further assessed in the nonlinear regime by
examining the influence of higher-order modes, neglected in the previous model
[ 18], on the nontrivial steady-state solution branches C 1 and C2. The additional
modes stem from normal stress effects and are thus found to have most of their
influence on highly elastic flows, and are rather insignificant in the high
Reynolds number range or for a Newtonian fluid. Indeed, for a Newtonian or
weakly elastic fluid, the inclusion of the higher-order modes does not alter
significantly the results in both the linear and nonlinear (fig. 2) regimes.
From the analysis and numerical results based on the present model, fluid
elasticity appears to alter dramatically the flow behavior in comparison with the
flow of a Newtonian fluid. One of the most striking differences in flow behavior
between the two flows is the destabilization of the TVF for a viscoelastic fluid
that cannot be predicted on the basis of the (axisymmetric) Newtonian equations.
Thus, unlike the Newtonian model, the Oldroyd-B model predicts the onset of a
Hopf bifurcation (at Re = Reh) in the postcritical range of Reynolds or Taylor
number (Re > Rec). Hence, to the degree of the present truncation, the
Newtonian equations predict that the TVF is always stable regardless of the
297

value of the Reynolds or Taylor number. This is somewhat expected since, for a
Newtonian fluid, the destabilization of TVF coincides, in practice, with the loss
of axisymmetry. This coincidence does not seem to hold for viscoelastic fluids
[9]. If fluid elasticity is accounted for in the constitutive equation, a different
stability picture appears to emerge. In fact, linear stability analysis of the 20NDS
around the TVF leads to a characteristic equation (not presented in this chapter)
that happens to admit a pair of purely imaginary roots, coinciding with the birth
of a Hopf bifurcation at Re = Re h as discussed earlier. The value of Reh depends
strongly on fluid elasticity, fluid retardation and wave number as illustrated in
fig. 2.
In an effort to examine the role of inertia on the onset of TVF and its finite-
amplitude instability, numerical calculations are carried out at intermediate
elasticity level (E = 0.16) by varying the Deborah (or, equivalently, the
Reynolds) number. The results (shown in figs. 3-6) indicate that, for weakly
elastic flows, there is an exchange of stability at De = De c (Re = Rec) between
the base flow and TVF, through a supercritical bifurcation similar to that in
Newtonian fluids. At some point in the postcritical range (De > Dec), the steady
TVF loses its stability, in turn, as De (Re) is further increased to De = Deh (Re =
Reh) when a Hopf bifurcation sets in. The influence of shear rate (De) was
investigated experimentally by Baumert & Muller [9]. Their flow visualization
focused essentially on the early transient behavior once a shear rate is imposed.
They showed that steady TVF sets in after a period of monotonic or oscillatory
vortex growth depending on whether the imposed shear rate is low or high. The
results based on the present formulation lead to good qualitative agreement with
experiment and the nonlinear calculations of Avgousti et al. [28].
For a highly elastic fluid, of Elasticity number greater than the critical value,
Esub, the steady-state solution undergoes a subcritical bifurcation at Re = Rec
(fig. 2) similar to that predicted by the Landau-Ginzburg's equation [44]. In this
case, the range of stability of the TVF becomes increasingly narrower as E
increases, giving way to (time) oscillatory vortex structure. Indeed, beyond a
critical level of elasticity (E > 0.1 in fig. 2) and for vanishingly small Reynolds
number, the base flow loses its stability to oscillatory and not to steady TVF.
This prediction confirms the experimental observation of Muller et al. [8].
Unlike the previous model [18], the present system led to favorable comparison
with experiment (figs. 7 and 8). The present theory predicts the sequence of
periodic behaviors observed as the Deborah number is increased: (1) loss of
stability of the base flow to an oscillatory flow at a critical Deborah number (Dec
= 29.3 and 29.7 for the RF and RR formulations, respectively, as predicted by the
model vs. 32 from experiment), (2) growth of amplitude of the velocity signature
298

like (De - Dec) 1/2, in agreement with asymptotic analysis, with the
emergence of higher harmonics in the Fourier spectrum, and (3) emergence
of subharmonics as De is further increased, reflecting (most likely) the
bifurcation into period doubling and, eventually, chaos.
Finally, an augmented twenty-dimensional dynamical system,
20NDS, is also used to describe highly elastic TVF. The model is derived
without inertia effect, and higher normal stress terms are added. This
approach constitutes a first systematic and accurate theoretical prediction
for the purely elastic overstability observed by Muller et al. [8]. It is shown
that the finite amplitude TVF can be effectively described if higher-order
normal stress terms are properly accounted for. Particularly, the addition of
the azimuthal normal stress component leads to additional coupling with
higher-order eigenmodes that are of O(eDe). The resulting nonlinear
dynamical system involves only twenty degrees of freedom. The sequence
of flows predicted by the present model is comparable to that reported by
Muller et al. [8]. The model predicts the sequence of periodic behaviors
observed as the Deborah number is increased: (1) loss of stability of the
base flow to an oscillatory flow at a critical Deborah number (Dec = 32 as
predicted by the model vs. 32.3 from experiment), (2) growth of amplitude
of the velocity signature like (De - Dec) 1/2, in agreement with asymptotic
analysis, and (3) the emergence of higher harmonics in the Fourier spectrum
as De is further increased [ 19].

REFERENCES

~ G. I. Taylor. Phil. Trans. Roy. Soc. A 223 (1923) 289.


2. T. B. Benjamin. Proc. Roy. Soc. Lond. A 239 (1978) 1.
3. D. E. Shaeffer. Math. Proc. Camb. Phil. Soc. 87 (1980) 307.
4. D. Coles. J. Fluid Mech. 21 (1965) 385.
5. P. R. Fenstermacher, H. L. Swinney & J. P. Gollub. J. Fluid Mech. 94
(1979) 103.
. M. Gorman & H. L. Swinney. J. Fluid Mech. 117 (1982) 123.
7. D. Rand. Arch. Ration. Mech. Anal. 79 (1982) 1.
8. S. J. Muller, E. S. J. Shaqfeh & R. G. Larson. J. Non-Newt. Fluid
Mech. 46 (1993) 315.
. B. Baumert & S. J. Muller. Rheol. Acta 34 (1995) 147
10. G. R.Sell, C. Foias and R. Temam Turbulence in Fluid Flows: A
Dynamical Systems Approach (Springer-Verlag, 1993).
299

11. E. N. Lorenz. J. Atmos. Sci. 20 (1963) 130.


12. C. Sparrow. The Lorenz Equations, (Springer-Verlag, New York
1983).
13. H. N. Shirer & R. Wells. Mathematical Structure of the Singularities
at the Transitions Between Steady States in Hydrodynamic Systems
(Springer-Verlag, Heidelberg 1980).
14. R. E. Khayat. J. Non-Newt. Fluid Mech. 53 (1994) 227.
15. R. E. Khayat. J. Non-Newt. Fluid Mech. 58 (1995) 331.
16. R. E. Khayat. Phys. Rev. E 51 (1995) 380.
17. R. E. Khayat. J. Non-Newt. Fluid Mech. 60 (1996)
18. R. E. Khayat. Phys. Fluids A 7 (1995) 2191.
19. R. E. Khayat. Phys. Rev. Letts. 78 (1997) 4918.
20. R. G. Larson. Rheol. Acta 31 (1992) 213.
21. E. G. S. Shaqfeh. Ann. Rev. Fluid Mech. 28 (1996) 129.
22. H. Giesekus. Rheol. Acta 5 (1966) 39.
23. H. Giesekus. Prog. Heat Mass Transfer 5 (1972) 187.
24. R. Haas & K. Btihler. Rheol. Acta 28 (1989) 402.
25. R. G. Larson, E. S. G. Shaqfeh & S. J. Muller. J. Fluid Mech. 218
(1990) 573.
26. E. S. G. Shaqfeh, S. J. Muller & R. G. Larson. J. Fluid Mech. 235
(1992) 285.
27. Y. L. Joo & E. S. G. Shaqfeh. Phys. Fluids A 4 (1995) 2415
28. M. Avgousti, B. Liu & A. N. Beris. Int. J. Num. Meth. Fluids 17
(1993) 49.
29. M. Avgousti & A. N. Beris. J. Non-Newt. Fluid Mech. 50 (1993) 225.
30. P. J. Northey, R. C. Armstrong & R. A. Brown. J. Non-Newt. Fluid
Mech. 42 (1992) 117.
31. A. Groisman & V. Steinberg. Phys. Rev. Letts. 77 (1996) 1480.
32. M. K. Yi 7& C. Kim. J. Non-Newt. Fluid Mech. 72 (1997) 113.
33. R. G. Larson. Rheol. Acta 28 (1989) 504.
34. H. Kuhlmann. Phys. Rev. A 32 (1985) 1703.
35. H. Kuhlmann, D. Roth & M. LOcke. Phys. Rev. A 39 (1988) 745
36. E. Ott. Chaos in Dynamical Systems (Cambridge University Press,
Cambridge 1993)
37. J. Guckenheimer and P. Holmes. Nonlinear Oscillations, Dynamical
Systems, and Bifurcations of Vector Fields (Springer-Verlag, New
York 1983).
38. G. Veronis. J. Fluid Mech. 24 (1966) 545.
39. J. H. Curry. Commun. Math. Phys. 60 (1978) 193.
40. H. Yahata. Prog. Theor. Phys. 59 (1978) 1755.
300

41. H. Yahata. Prog. Theor. Phys. 61 (1979) 791.


42. H. Yahata. Prog. Theor. Phys. 59 (1978) 1755.
43. R. B. Bird, R. C. Armstrong & O. Hassager. Dynamics of Polymeric
Liquids, vol. 1, 2nd Ed. (John Wiley & Sons, New York 1987).
44. P. G. Drazin & W. H. Reid. Hydrodynamic Stability (Cambridge
University Press, Cambridge 1981).
45. H. Harder. J. non-Newtonian Fluid Mech. 36 (1991) 67.
46. J. B. McLaughlin & P. C. Martin. Phys. Rev. A 12 (1975) 186.
47. A. Davey. J. Fluid Mech. 14 (1962) 336.
48. J. T. Smart. J. Fluid Mech. 4 (1958) 1.
4.9. P. Chossat & G. Iooss. The Couette-Taylor Problem. Springer-Verlag
(1991).
50. M. Nagata. J. Fluid Mech. 169 (1986) 229.
51. P. Tabeling. J. Phys. Lett. 44 (1983) 665.
52. D. V. Boger. J. Non-Newt. Fluid Mech. 3 (1977/78) 97.
53. R. D. Richtmyer. NCAR Tech. Note TN-176+STR (1981).
54. P. S. Marcus. J. Fluid Mech. 146 (1984) 65.
55. S. Chandrasekhar. Hydrodynamic and Magnetohydrodynamic
Stability, Dover (1961).
56. R. E. Khayat. J. Fluid Mech. (in press).
57. D. W. Beard, M. H. Davies & K. Walters. J. Fluid Mech. 24 (1966)
321.
58. R. H. Thomas & K. Walters. J. Fluid Mech. 18 (1964) 33.
59. R. F. Ginn & M. M. Denn. AIChE J. 15 (1969) 450.
60. C. Normand, & Y. Pomeau. Rev. Mod. Phys. 49 (1977) 581.
61. S. Wiggins. Introduction to Applied Nonlinear Dynamical Systems
and Chaos (Springer-Verlag, New York 1990).
62. J. E. Marsden & M. McCracken. The HopfBifurcation and Its
Applications (Springer-Verlag, New York 1976).
63. J. Carr. Applications of Center Manifold Theory (Springer Verlag,
New York 1981).
63. K. Walters. Rheometry: Industrial Applications. Research Studies
Press (1980).
64. R. I. Tanner. Engineering Rheology. Oxford University Press (983).
65. R. B. Bird, R. C. Armstrong, C. F. Curtis & O. Hassager. Dynamics
of Polymeric Liquids, vol. 2, 2nd Ed. John Wiley & Sons (1987)..
66. M. Doi & S. F. Edwards. The Theory of Polymer Dynamics. Oxford
University Press (988).
67. B. C. Eu & R. E. Khayat. Rheol. Acta 30 (1991) 204.
301

NON-NEWTONIAN MIXING WITH HELICAL RIBBON


IMPELLERS AND PLANETARY MIXERS

Philippe A. Tanguy ~ and Edmundo Brito-De La Fuente 2

1 Department of Chemical Engineering, Ecole Polytechnique Montreal


P.O. Box 6079, Station Centre-ville, Montreal H3C 3A7, CANADA

2 Food Science and Biotechnology Department, Chemistry Faculty "E",


National Autonomous University of Mexico, UNAM, 04510 Mexico, D.F.,
MEXICO

1. INTRODUCTION

Mixing is a very common processing operation, which accounts for


about 15 % of all unit operations in the chemical and food industries. The
range of possible mixing duties is extremely wide and may involve a single
phase (in which case agitation is a more appropriate wording) or several
phases. Moreover, the phases may be of a different nature, solid, liquid and
gas. All combinations may be found depending on the process. Table 1
summarizes some important practical mixing applications.

Table 1. Typical mixing applications

Kneading of Pastes Crystallization


Dispersion of Agglomerates Emulsification
Flocculation Flotation
Gas-Liquid Dispersion Heat Transfer
Liquid-Liquid Dispersion Liquid Blending
Reaction Enhancement Polymerization
Slurrying Solids Suspension
Storage Blending Uniformity Maintenance
302

A mechanical mixer is a relatively simple device. It comprises a


vessel, an impeller mounted on a shaft and a drive assembly. In mixing
process development, a number of issues must be addressed for a given
mixing application, namely
9 the selection of mixing equipment (impeller, drive) and operating
conditions (speed, temperature, mixing duration in batch mode)
9 the design of the vessel (shape, aspect ratio, type of bottom, position of
feed and side streams) and the internals (baffles, draught tube, heat
exchangers)
9 the evaluation of performance (power consumption, mixing time,
circulation patterns, circulation rate, distribution of shear rate)
9 the control of composition non-uniformity (segregation).

Turbulence is the main physical mechanism responsible for mixing


with low viscosity fluids. The impeller power is split into the power
dissipated by the circulating fluid and the power dissipated by the sheafing
action of the impeller in the vicinity of the blade tip. The respective amount
of these two power components depends on the impeller type. When the
viscosity is high, turbulence can hardly be achieved and when it is, the
power drawn by the impeller is extremely large, yielding high operating
costs and a significant temperature increase in the fluid bulk.
Viscous mixing is a laminar process based on the "stretching-folding-
breaking" principles described in [1 ]. Mixing is obtained by the mechanical
decrease of the striation thickness by shear and/or extensional forces up to a
scale where molecular diffusion may fully play its role. This is a rather
"gentle" flow process, which usually requires time.
As with other fluid mechanics problems, the Reynolds number is the
relevant number to characterize the significance of the viscosity effects. The
Reynolds number is defined as:

Re -
-
pND2 (1)
77

where p and 1] are the fluid density and dynamic viscosity respectively, N is
the rotational speed of the impeller and D its diameter. Below Re = 10, the
viscous effects predominate, yielding a quasi-creeping flow. The mixing
regime is laminar. At the other end of the Reynolds number scale, the
mixing regime is turbulent. Depending on the impeller, the turbulence
303

threshold is comprised between Re = 104 and Re = 3.104. The intermediate


mixing region above the laminar regime and below the turbulent regime is
called the transition regime.
In principle, mixers can be operated in any of this regime, provided
adequate power is available. In practice, for viscous fluid, as turbulence
cannot be achieved in the vessel, the mixing regime is at best in the
transition region and often in the laminar region. Experience shows that the
mixing regime is always laminar when the fluid viscosity is above 50 Pa.s.
High viscosity fluids usually exhibit non-Newtonian properties. Non-
Newtonian flows are very common in the process industry. Typical
examples include the manufacturing of rubber, plastics, petroleum,
detergents, cosmetics, pharmaceuticals, cement, food, paper pulp, paints to
name a few. A wide variety of flow situations are encountered in practice
(flow in transfer lines, in processing equipment). Our discussion in this
chapter will focus on non-Newtonian mixing and we refer the reader to the
literature [2] for the treatment of the general principles governing non-
Newtonian fluid mechanics.
The most common non-Newtonian behavior is shear-thinning, i.e. the
viscosity decreases with an increasing shear rate. This behavior is very often
encountered in polymer manufacturing or in fermentation. Yield stress is
often present especially with gels and highly concentrated suspensions. The
yield stress phenomenon is associated with the presence of structures in the
fluid that requires a minimum amount of deformation energy to break up.
The mixing of yield stress fluids (or B ingham fluids as they are often called)
is really difficult due to the particularly complex fluid mechanics in the
vessel. A well-mixed cavern forms in the fluid bulk around the impeller [3]
due to the shearing action of the blades. Far from the impeller, the rate of
deformation is not sufficiently high to break the structure and the fluid is at
rest. The same phenomenon may be encountered with shear-thinning fluids
although in a less severe manner [4].
Another particular behavior is thixotropy, which is manifested by a
time-dependent shear-thinning. From a physical standpoint, it is believed
that thixotropy is related to the presence of organized microstructures whose
conformation changes under an applied stress. Examples of such fluids are
chocolate and paper coating colors. Thixotropy can be described in
engineering terms as substances with several yield stresses [5][6]. Likewise
B ingham fluids, thixotropic fluids may be difficult to mix.
Shear-thickening is the opposite behavior of shear-thinning, i.e. the
viscosity increases with the shear rate. This behavior is observed with
304

colloidal suspensions at very high concentration like high solids mineral


slurries and with starch. Fluids exhibiting this property must be handled
very carefully. Indeed, as the power consumption is proportional to the
viscosity (as it will be seen later in this chapter), if shear-thickening
properties develop during the process, the impeller may stall in the vessel
or, worse, the motor may pull out.
Viscoelasticity may add another level of complexity in the mixing of
non-Newtonian viscous materials. Like shear-thinning, viscoelasticity is
encountered in polymerization and fermentation. Apart from the well-
known shaft climbing phenomenon by the fluid (known as the Weissenberg
effect in rheology) viscoelasticity increases significantly the power
consumption in the vessel and strongly alters the circulation patterns [7].

2. S U M M A R Y OF MIXER DESIGN PRINCIPLES

The mixer performance, expressed in terms of energetic, dispersing


sheafing or blending efficiency, is directly related to the impeller
performance. The selection of the appropriate device requires a very careful
appraisal of the mixing duty, and a systematic stepwise approach based on:
9 Identification of the operations to be carried out
9 Detailed physical characteristics of the phases (Table 2)
9 General features of the impeller (shear rate, circulation rate, pumping
pattern)
9 Selection of impeller candidates
9 Final selection

Table 2. Phase characteristics

Liquid Solid Gas


specific gravity Specific gravity Flowrate
viscosity curve Particle size distribution Pressure
yield stress Settling velocity Solubility
temperature range Wettability- solubility
wt. % wt. %

There are two possible ways of classifying an impeller, according to


the discharge flow it produces, or according to its size with respect to the
305

vessel. In the first category, there are radial discharge impellers (flat
turbine), axial discharge impellers (propellers) and tangential flow impellers
(anchors). If the diameter of an impeller is significantly smaller than the
diameter of the vessel, this impeller is an open impeller. On the contrary, if
the impeller exerts a scraping action at the vessel wall, it is called a close-
clearance impeller. We show in Figure 1 typical examples of open and
close-clearance impellers.

Figure 1: Flat blade turbine (open impeller) and helical ribbon screw (close-
clearance impeller)

In viscous and non-Newtonian mixing, close-clearance impellers are


by far the best choice due to their superior top-to-bottom pumping capacity
and wall scraping action as compared with open impellers. If the medium is
a paste, multiple impellers comprising high speed tools for intense shearing
in the bulk (emulsification head, dispersing disk) and low speed wall
scraping blades may be advantageously used.
We show in Figure 2 an example of kneading equipment. Sometimes,
the various impellers are mounted on a rotating carousel (planetary mixers).
With this technology the kinematics of the impellers is such that all the
vessel volume is swept at regular intervals. There is therefore no dead zone.
There is not a unique choice of mixing technology for a given
application and several impellers may likely satisfy the mixing
requirements. The final selection will be based on both process operating
conditions and economics. Design parameters to consider include operating
306

mode (batch or continuous), volume to be processed, mixing time, and


pressure in the vessel. The design of the vessel and the internals is also
critical from mechanical and economic reasons: shape, dimensions, open or
closed tanks, positions of feeds and outlets, baffles, heating coils are all
important technical decisions that will directly impact the overall mixing
system efficiency.

Figure 2: a multiple tool kneading head

In industry, the use of empirical correlations has been the standard for
the design of mixing systems. Nowadays, stringent process efficiency
requirements make the use of pilot experiments in association with
computer modelling an essential step of the development phase. From the
formulation developed in the laboratory by the chemists, the process is first
developed at pilot scale (typically 200 to 500 1) and then some scale-up
guidelines are used to design the industrial unit. Computer modeling can be
used at both steps, for the pilot scale and for the eventual production line.
The selection of the mixing scale-up criteria to be used for the design
of industrial facilities is an open issue. The fundamental question to address
is how a process changes when the scale is changed, the objective being
obviously to maintain process similarity irrespective of the scale.
There are two broad classes of similarity: flow and transport
phenomena. Flow similarity may involve the geometry (shape similarity),
the kinematics (speed similarity) and the dynamics (force similarity). As for
307

the transport phenomena, one may have thermal similarity (same


temperature gradients), chemical similarity (same concentration gradients)
and rheological similarity. In the latter case, the same distribution of
deformation rates (shear and extension) is sought between the pilot process
and the scaled-up installation. In practice, when the process scale is
changed, the new design will be based on keeping some parameters equal.
The most common scale-up parameters are:
9 mixing times (in laminar regime, the mixing time is inversely
proportional to the impeller speed)
9 power per unit volume
9 impeller tip velocity
9 effective shear rate or process viscosity.

An approximate similarity can be obtained if the same geometry and


the same kinematics is used, while using proportionality rules. For instance,
if we want to keep the impeller tip speed ND constant, if D is increased, N
should be diminished proportionally. An improved similarity would
consider identical energy dissipation scales. The main difficulty is the effect
of walls and bottom surface whose influence decreases drastically as the
vessel size is increased. This becomes a serious constraint when heat
transfer is involved during the mixing.

3. NON-NEWTONIAN MIXING F L O W S

From an historical perspective, non-Newtonian mixing flows in


agitated vessels have been first dealt with simple models based on simple
physics principles and correlations, and then with computer fluid dynamics
(CFD) approach. These two approaches will be described here, the scope
being restricted to close clearance impeller mixers and planetary mixers.

3.1 Flow models in a mixing vessel


Several flow models have been proposed in the literature to predict
power consumption with close clearance impellers. They fall in two main
groups, namely flow between two coaxial cylinders and drag-based analysis
As it was mentioned in Section 1, the mixing of highly viscous and
rheologically complex fluids is a slow process both at the macro and micro
scale. As the flow regime is laminar, the theoretical notions given in the
following paragraphs are valid mainly for this regime.
308

3.1.1 Coaxialflow model


The basic idea is to depict the impeller as a solid cylinder rotating
inside another cylinder (the vessel). This geometry is known as the Couette
geometry.
Let us consider the case of a shear-thinning fluid obeying the power
law model:

O~rl)-- 2(n-1)12Klrl n-' (2)


where K is the consistency index, n the shear-thinning index and T the shear
rate. It is easy to show that in the Couette geometry [8], equation (3)
expresses the variation of the shear rate with the radius, namely:

[rd(Vo/r)l_[ 2~]n 1(1)2


dr r1-27n --- r7-21n
/n
(3)

Equation (4) gives the shear rate at the inner rotating cylinder (or
impeller):

I 1-(rl[r2)a/n2)1[ 2or2]
nO-(r, l r 2 r22-rl 2
(4)

In these equations, f~ is the rotating speed (in rad/s), V0 the angular


component of the fluid velocity, and r~ and r2 are the inner and outer radius
respectively.
From a mixing perspective, two cases of practical interest are
included in equation (4). The first one occurs when the impeller shear rates
are only contained in a small volume very close to the blade. This situation
is common in laminar mixing, where the flow generated by the agitator does
not reach the tank wall as if the impeller was rotating in an infinite medium.
For this case, the usual procedure is to assume that rl/r2 essentially
approaches zero. Then, equation (4) becomes:

7, - 4n:N 1 (5)
n
309

According to equation (5), the shear rate is linearly related to the


impeller rotational speed, N, and the rheology.
The other interesting case arises when r~/r2 approaches 1. This
situation can be found when using very small clearances in the mixing
geometry. Then, equation (4) becomes:

2 ~2 r 2 I (6)
~/ ~ 2 2
r 2 -- r1

The flow patterns around mixing impellers are fairly complex and the
controlling factors of the drag on an impeller are difficult to appraise. In
practice, simplified flow situations are used to estimate the viscous energy
dissipation, and then deduce the drag or power requirements. In non-
Newtonian mixing technology, the flow model around an impeller blade
must obviously include the shearing rates or shearing stresses. This is a
direct consequence of the fact that the viscosity of these fluids to an
imposed stress is not constant but depends on the magnitude of the
deformation rate.
The knowledge of shear rates in stirred mixing tanks is central to
power input calculations and equipment scale-up. Empirical and theoretical
approaches have been developed in the literature to estimate shear rates. In
the following paragraphs, these approaches are discussed and compared.

3.1.2 Metzner and Otto Approach


In late 50's, Metzner and Otto [9] introduced the following rule to
define an apparent viscosity in the mixer. If a Newtonian fluid and a non-
Newtonian fluid are agitated in the laminar regime under the same operating
conditions and in the same equipment (so that the power input measured is
the same), then, because all variables are identical, the average viscosity is
the same for both fluids. This quite important result for mixing practice was
expressed by equation (7):

y'- KsN (7)

where Ks is a constant of proportionality which has to be determined


experimentally for each impeller geometry of interest. In other words, the
average fluid shear rate is related only to the impeller speed.
310

It is worth noting here that equation (7) is only valid for laminar
mixing, although it has been used for the transition regime. The value of Ks
is typically estimated by performing duplicate power input measurements,
one with a Newtonian fluid and one with a non-Newtonian fluid, in the
same mixing system. Thus, if the rheological behavior of the fluid is
expressed by a power-law model, then, from the experimental Newtonian
power input curve, at fixed N, an apparent or effective viscosity defined as:

~a -- K(Yav )n-1 (8)

is calculated by:

(9)
770-77

where 1"1is the Newtonian viscosity, PnN and PN are the non-Newtonian and
Newtonian power respectively.
From the apparent viscosity calculated in equation (9), the
corresponding value of )'av is estimated by equation (8). Upon repeating this
procedure for pairs of (N, ~/av)values, the value of Ks is determined.
It is worth nothing here that as n tends towards the limiting
Newtonian case (n=l), equation (5) gives 7 = 41-IN = 12.6N or Ks = 12.6.
This is very close to the universal value of Ks = 13, originally proposed in
the literature [9] for several fiat-blade turbines.
The Metzner-Otto correlation has been extensively used to
characterize the average shear rates of non-Newtonian inelastic fluids [ 10],
and it is routinely recommended as a standard procedure in mixing and unit
operations textbooks. This algorithm is considered a useful approach for
the estimate of hydrodynamic similarities on scale-up although it might not
represent the rheological properties existing in the mixing vessel. It is a
current practice to consider Ks as a constant for given impeller geometry.
However, some authors have reported that this constant might depend on the
power law index [11][12][13], in particular for highly shear thinning fluids.
A compilation of experimental Ks values for helical ribbon (HR) and helical
ribbon screw (HRS) impellers is shown in Table 3.
A detailed analysis on the correlations proposed for Ks clearly shows
that the effect of wall clearance might be the most important one. The value
of Ks increases as the wall clearance decreases and this becomes more
pronounced as the ratio r~/r2 approaches the limiting value of 1.0.
311

Table 3. Ks data available in the literature (HR and HRS impellers)

i
Type , D/d n Ks Comments

HR 1.02- 1.12 0 . 4 - 1.0 66.06 a Weak Ks(n) [14]


Ks(geometry)

HR 1.059 0 . 2 - 1.0 25 Ks = constant [15]

HR 1.10- 1.11 0 . 3 5 - 1.0 27 Ks - constant [16]

HRS 1.03 0 . 3 5 - 1.0 24.58 Weak Ks(n) [17]


Ks(geometry)

HR 1.053 0 . 5 - 1.0 36.73 Ks = constant [18]

HR 1.056 0 . 2 7 - 1.0 30 Ks = constant [19]

HR 1.042- 1.19 0 . 5 - 1.0 27.60 a Ks(geometry) [20]

HRS 1.056 - 1.118 0 . 2 6 - 1.0 30.6 Ks = constant [21]

HR 1.11 - 1.37 0.17 -0.65 79.85 b Strong Ks(n) [22]


Ks(geometry)

HR 1.05 - 1.163 0.35 -0.75 24.68 a Ks(geometry) [23]

HR 1.05- 1.33 26.80 a Ks(geometry) [24]

HR 1.11 0 . 1 8 - 1.0 17-40 Weak Ks (n) [13]


Ks (geometry)

HR & 1.135 0 . 0 9 - 1.0 7 - 36 Ks (n) for n < 0.5 [12]


HRS Ks (geometry)
This value is calculated for D/d = 1.10.
Value estimated for n = 0.6; D/d = 1.135 and 1/d - 3.24, where 1 = length
of impeller blade.

Regarding the functional dependence of Ks on the fluid flow


properties, for most authors, Ks may be considered as independent of the
rheological properties. This is particularly true for those authors who have
strictly followed the Metzner and Otto assumptions. However, for authors
who developed correlations based on the Couette flow analogy principle,
the resulting Ks was found to be a function of the shear-thinning level.
312

3.1.3 Extended Couette Flow Approach


Bourne and Butler [25] investigated the mixing of viscous Newtonian
and shear thinning fluids with helical ribbons, using flow pattern
visualization and liquid velocity measurements. They proposed the
existence of three main flow zones:
9 the core flow surrounding the impeller shaft
9 the region between the blades of the impeller
9 a highly sheared zone located between the tank walls and the impeller
blades.

The authors used the Couette flow analogy to describe the highly
sheared zone substituting the impeller by a solid cylinder. Setting as a
characteristic fluid velocity the tip speed velocity ND, a particular form of
the Reynolds number for power law fluids was proposed:
pN2-nD2
R e p~ = K (10)

Thus, the familiar expression for the dimensionless power input is


given by:
P
Np Re pl KNn+ ] D3 =Kp(n) (11)

In the above equations, no Ks values must be known a priori to


handle experimental power input data. Equations (10) and (11) have been
the basis for the work of Chavan and Ulbrecht [ 17], and Rieger and Novak
[ 18]. Unlike in [25], Chavan and Ulbrecht considered only the geometry of
the helical ribbon impeller to define an equivalent cylinder diameter to be
used in the Couette flow analogy. More recently, the use of an equivalent
diameter based on the ribbon geometry and the fluid rheological properties
has been suggested as improvements over previous studies [26][13].
Figure 3 is a typical representation of power input experimental data
obtained with a HR impeller for a wide range of shear thinning fluids, using
equations (10) and (11). Details regarding the experimental setup and
geometry of the HR impeller are given elsewhere [27].
As suggested by the results of this figure, the power consumption
decreases as the fluid becomes more shear-thinning (decreasing values of n)
at the same Reynolds number. It is also clear that as n tends towards the
limiting Newtonian value (n=l), the power data tends to the Newtonian
power correlation.
313

10 8 -- i. l I I I I I I I ' ' ' ' ' ' ' 1 ' ' ' ' ' ' ' ' 1 ' ' ' ' ' ' ' ' 1

_
-

_ Q
-

10 5 - "'..

.%
_
_
_

10 4 _--
"O
_
o ~ e o
Z _

10 3 _-
_
_
[] n = 1.0 (Newtonian) "'-.O.'.-.+ :-:
_ * n = 0.77 ~-~"
-, . . _

_
/x n = 0.70 ' O . ~I k . ' & ..

i
0 n = 0.5 3 " "',II~"Lx "J,
10 2 "O "'', " .
_ 9 n = 0.28 6A-
_
I I I I I I 'll I I I ' ' ' I l l I I I , , , , , I I i n i i i n i l -

1 0 .4 1 0 .3 1 0 .2 1 0 ~ 1 0 ~

Repl

Figure 3. HR impeller power curves

Figure 4 is the plot of the Kp(n) function corresponding to the data


shown in Figure 3. The Kp(n) function can be represented by a non-linear
model [11][ 12], b and c being fitting parameters, namely:
n-1

Kp (n) - Kp(n=l) bn-1 c n


(12)

It should be noted that the representation of the non-Newtonian


power input as shown in Figure 3 is not very convenient since at least two
dimensionless parameters, n and Re, influence the power number. Extensive
experimentation would thus be necessary to include the effect of other
variables. Another drawback is that this type of representation is not general
and unique.
A unifying principle can, however, be established by comparing the
Metzner and Otto approach with equation (11). Indeed, it can be shown [ 11 ]
that:

_ P
Np Repl - - KN,,+~ 9 3 "- K p (n) - Kp(n=l) Ks n-1 (13)
314

or

Ks- mN--2-+]d3
1

II- Kp(n)
gp
n-1
1

(14)

G e l l a n 1.5 %
Xanthan 3%
180 X a n t h a n 2 % - C M C 1% J '
Xanthan 1.5%-CMC 1.5%
160 - Xanthan 1.75%-CMC 1.25%
Q Xanthan 1.65%-CMC 1.35% //
/
140 - 9 Xanthan I%-CMC 2% /
/
[] CMC 2.5% / /
120 - A CMC 1.5% /
Newtonian fluids /
100 -
Kp - -
a=
80 -
/

60

40

20 -

0 , I , I , I = I ,
0.0 0.2 0.4 0.6 0.8 1.0

Figure 4. Variation of the power constant with the power law index

As suggested by equation (14), Ks plays the role of a shift factor


which should coalesce all the non-Newtonian power input results into the
Newtonian curve, yielding a unique power input (master curve). We show in
Figure 5 the result master curve corresponding to the data of Figure 3. As it
can be observed, the whole set of non-Newtonian power input data are
superimposed on the Newtonian curve when using the experimental Kp(n)
function as shown in Figure 4 and equation (14).
315

10 6

dl
10 5 - 1
Np Re = 162.55 -!
[exp. Newtonian correlation]-
U3
"=t
10 4 -
Z
Q.

II 10 3 -

Z
9 Xanthan 3 % "~1~
-i
9 Gellan 1.5% ' ,I~~,
10 2 - 9 Xanthan 2%-CMC 1%
9 Xanthan 1.5%-CMC 1.5% Orn
9 Xanthan 1.75%-CMC 1.25%
[] CMC 2.5%
I I I I I I I
10 ~
10-3 1 0 -2 10 1 10 ~ 10 ~

Re = p N 2"n d 2/In Ks n'l

Figure 5. Power master curve for HR impeller

It is worth noting that the functional form of equation (14) makes the
Ks function very sensitive to small variations in power consumption
measurements, particularly in the highly shear thinning region (0.5 < n).
This is illustrated in Figure 6 for a power measurement uncertainty of 10%.
It can be seen that the interpretation of the results on the actual value
of Ks is delicate. With the measurement uncertainty, all conclusions are
possible: Ks may be taken either as a constant (independent of the fluid
rheological properties), or as a decreasing or even an increasing function of
the flow index. The most significant variations of Ks are found in the highly
shear thinning region, 0.1 < n < 0.5. Undoubtedly, these results shed some
light on the apparent opposite findings shown in Table 3. In the opinion of
the present authors, the differences found in the literature could be easily
explained if the impact of experimental errors on power measurements had
been fully accounted for.
316

40 ''''1''''~'''1''''1''''1''''1''''1''''1''''1'''

35 \
c=l.1
30 -.
~Oo
c=l.O
25

20
/ c = 0.913 (exp. data)

15

10

5 i ,,,I,,,,I,,,,I,,,,I,,,,I,,,,I,,,,I,,,,I,,,,I,,,,
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 6. Ks sensitivity to power consumption uncertainty

3.1.4 Drag force-based analysis


This approach considers that the torque exerted on the rotating
impeller is due to the drag force exerted by the fluid flow around the
impeller blade. The forces acting on the blade are namely [28]:
9 the dynamic pressure due to the normal velocity at the blade (known as
normal drag or form drag),
9 the tangential stresses caused by the fluid friction (skin drag) along the
surface of the blade.

If the impeller tip speed is taken as the characteristic velocity, the


mixing power can then readily obtained with the following equation:

P - z ( 2 ~ N ) - 2~rbNFdo (15)

where rb is the radius of the impeller blade, and Fd0 the drag force in the 9
direction. The drag force can be written in terms of the drag coefficient, Cd,
as"
317

( l pVo2~
F d - C d -~ (16)

where A is a characteristic blade area and V0 is the angular component of


fluid velocity.
In the literature, only a small number of correlations based on the
drag flow analogy has been proposed. The existing ones differ mainly in the
expression used for Cd. It is worth noting that most of these correlations are
limited to Newtonian fluids. For non-Newtonian fluids, one correlation is
presented in Table 3 [22]. For other works on this topic, the reader is
referred to [29].
In the following section, we now turn to the computer modelling
approach of mixing flows.

3.2 Numerical simulations


We consider the flow of a viscous fluid generated by an impeller
rotating in the center of a cylindrical vessel. Using symmetry of revolution
arguments, the velocity and the pressure fields can be readily obtained by
resolving the standard equations of motion in the non-Galilean frame of
reference of the impeller; in other words by considering a stationary
impeller and a rotating vessel. In this Lagrangian viewpoint, the equations
of change read as:

- Vp + divT - p(v.Vv + co o (co o r ) + 2o9 o v )

fdivv = 0

-, (Irl)r
(17)

where v is the velocity vector, p the pressure, r the radial coordinate and
~0the angular velocity. The symbol o denotes the standard vector product.
This equation is written without considering the unsteady term. This
approach is valid provided the flow in the Eulerian viewpoint is periodic.
For completeness, the viscosity function in equation (17) must be made
explicit with a rheological model. Typical models used in mixing
simulations include the Newtonian model (constant viscosity), shear-
thinning (power law and Carreau models) and yield stress (Herschel-
Bulkey) fluids.
318

Power law model

O~'l)-- 2(n-1)/2 girl'-' (2)

where K is the consistency index and the n the shear-thinning index. These
parameters can be readily obtained by performing a logarithmic regression
on the flow curve.

Carreau model

r/~?'l) r/= + (770 r/oo)(1+ 2(1t,I ~s


_ _ 2 (18)

where 7/0 is the zero-shear viscosity, r/_ the high-shear viscosity, t, a fluid
characteristics time and s the slope of the viscosity curve. This slope is
related to the shear-thinning index of the power-law model by s = (1- n)/2.

Herschel-Bulkley model

TO 2 (n-l)/2 (19)
r/~'l)- 21/217,--~+ Klrl" '

where ~:0 is the yield stress, K the consistency index and n the shear-
thinning index. When the viscosity of the fluid is constant, the familiar
B ingham model is obtained:

TO
rTOy])-~21,2it
+ r ]
I (20)

For mathematical well-posedness, the equations of change must be


completed with appropriate boundary conditions. In the Lagrangian frame
of reference chosen, the boundary conditions read as [30]:
9 at the vessel wall and bottom: v = Xwallo r
9 on the impeller: v=0
9 at the free surface: Vz = 0

The last boundary condition implies that the free surface in the vessel
is fiat. Experience shows that this assumption is valid in viscous mixing.
319

The above equations are solved using the Galerkine finite element
method. Let us recall briefly the principle of this method. We first rewrite
the equations of motion using tensor notations, namely:

f div FI = f (21)
div V = 0
II = - P 5 + 271)'
With help of the variational calculus, these equations may be written
as a saddle-point problem, namely:

Inf Sup ~ ~t (r(v)) 2 d ~ - ~a P div v d f ~ - ~ f v d ~ (22)

The equilibrium equations of this saddle-point problem are:

a(v, lg)-b(lg, p) = ( f , ~ ) , V ~ ~ [H~ (f~)] 3

fb(v, O) - O, V O e L 2 ( a )
a (v, N) - ~a/1 grad v grad N d ~

b(v, ~) - fa ~ div v df~

where H and L are appropriate Sobolev spaces and

(u, v) - ~~ uvdf~
(23)

(24)

The principle of the finite element method is the transformation of the


above variational problem into a set of algebraic equations that can be
readily solved with the tools of linear algebra. For this purpose, suitable
approximations are used for the velocity, the pressure and the test functions.
In order to guarantee stability and convergence of the numerical solution,
these approximations must be carefully chosen.
In three-dimensional mixing simulations, the super-linear tetrahedral
element PI+-Po and the tetrahedral quadratic element P2+-P1 are state-of-the-
320

art elements [31]. Tetrahedral elements are a much better choice than
quadrilateral elements for several reasons:
9 they are more flexible for unstructured grids
9 their accuracy suffers less from distortion
9 they are less costly to use that their equal-order quadrilateral counterpart.

After discretization, and using the penalty-Uzawa algorithm for the


treatment of incompressibility, the matrix problem to solve is of the form:

A
rv
i+1
--
BT p i = F

A r = A + rB rB (25)
pi+l _ p i _ r n v i + l

where A and B are the elliptic (viscous) matrix and the divergence matrix
respectively and r is the penalty parameter. The reader is referred to [32] or
an advanced textbook on finite element methods in fluid flow problems for
a detailed exposure of the above principles.
For very viscous fluids and pastes, a single impeller may not be
sufficient to mix the media and stagnant regions may occupy most of the
vessel volume. In such a case, multiple impeller kneaders and planetary
mixers must be used.
The fluid mechanics of mixing inside a planetary mixer is much more
complex than with a single impeller. From a simulation standpoint, the
complexity arises from the fact that there is no symmetry of revolution
usable to simplify the flow description. The most convenient viewpoint for
flow description is the classical Eulerian viewpoint of the laboratory and the
problem must be dealt as fully transient.
In order to tackle kneading flows, a new approach has been
developed called the virtual finite element method [33]. This method is
nothing but a subset of the broad category of domain imbedding finite
element method. In short, the idea is to represent in the volumic finite
element mesh, the impeller as kinematics constraints. For this purpose, the
impeller is discretized with control points on which the kinematics is
known. Then these additional constraints are implemented into the finite
element discretization. Mathematical optimization methods (penalty method
and Lagrange multipliers) are used to enforce these constrainnts. The final
matrix problem is modified accordingly and an extra step is required to
update the Lagrange multiplier iteratively.
321

The flowfield is obtained after resolution of the set of governing


equations. This field adequately post-processed may yield very useful
information for design purposes (power consumption, shear and extensional
force profiles, mixing segregation) and process understanding (dispersion
and distribution mechanism, chaotic features). We propose to illustrate
some of these capabilities in the next section.

4. R E S U L T S A N D D I S C U S S I O N

4.1 Validation and Couette flow analogy


The validation of the above numerical method is carried out for the
mixer shown in Figure 7. The impeller is a single flight helical ribbon close-
clearance impeller.

D s = 0.013 rn
W=0.03 m

C = 0.0125 m

P = 0 . 1 8 5 rn H = 0.24 m

D = 0.210 m ~I

Figure 7: Single flight helical ribbon impeller mixer

We first show in Figure 8 the variation of Np vs Reg for Newtonian


and shear-thinning power law fluids (CMC solution, 1.5 wt. %) at room
temperature up to a Reg value of about 2. It can be seen that the agreement
between the numerical predictions and the experimental data is excellent,
which confirms the findings of an earlier study [30].
322

1111111
I ::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
. . . . . . . . . 9...............
.............. .~
~ ...........
...................
.:....... ,~~....
..... ~~-.:---i--!"-i/
........ :...,......
.................. ' ........... ' ....... ~. . . . . . : . . . . . ~ . . . . i . . . ~ . . x..:" ................... : ........... c. . . . . . . .~ . . . . . . . . . . . ~ . . . ~ . . . . ~ . . . : . .
. . . . . . . . . . . . .9. . . ..:~ . .. . ... . ... .: . . . . . . . . ,. .~. ., .: : ~ . . . . 1 .~. . *: . ..' . :. i .!. ' :~ . . . . . -. . . . . . . . . . . . . . :i . . . . . . . . . . . ,~: . . . . . . :" . . . . 2~. . . d . , . .:~ . . . ~ .: . J . - : "
................. ~........ ..~.-.~ ..... i--..;...,..,..--~. .................. .:. ....... ~........ --...*...+.--t---~--.,--

I ................i........~-...".....~-.4-.~--~-!--i-..........4 ......~ ........i.....~'----i'"'-'"'!"


..............................:.-....... ...................~...........i.......~....~-----~---!---J-i--
. !.....i---i----i---!-i-;
.................. ~.......... -~ ...... ~-.... "...~-'..i-*-i
: i i ii.l~'
.............. ~ ......... i ....... .~ ..... ~-4 ...... .~-~--
! i i ! i:~! 9 num. n--1.00 [
9 num. n=0.651
Z" 100 . ..................
....................... ; ........... , ~i .......
....... :.. -L....~-~..,..'.~...-'--
"..,.~....:-.1..~,-~ . . . . x. . . ......
. i ...........
....... ~~........
. . . . . . . ,~
i ,.....
- . . . . : . i-.--:------i--.*-
---,~ ........... ...-
...... exp. n-l.00 ]
................... . ......... r .... :~ ~".'-*.'~"..'~:::::~2..~- ........ ~ .... .~ ..... ~....-...,-;..,.
" ~ ~ . - - ' ~ . " - . . _ ~ , . :-~..~_
................. ... ......... ; ...... ~....~....~__..'_...:.
,exp. n~O.esJ
. : : : ! : 9 i ' : ~. ~ :. ~ 3.-.:.,
.................. : ........... *. . . . . . . .

................. ~........... ~. ..... - . ~ ' ~ - - . ~ - ~


" ...... "'"'~"'T": ~": ................... ~ ........... ~ .......

................ ~ ........... ! ..... i ..... i--.-~ .......


: ..... :'""~ : ." :

i---i-

10 tf
0.10
i
1.00
i! iii i 1 10.00

Re.
F i g u r e 8: P o w e r c u r v e in the l a m i n a r m i x i n g r e g i m e

W e p r e s e n t in F i g u r e 9 the i s o - s u r f a c e ( s u r f a c e o f c o n s t a n t v a l u e s ) o f
the p r o c e s s v i s c o s i t y for the s a m e C M C s o l u t i o n a n d a r o t a t i o n a l s p e e d o f
10 R P M . In this case, a v a l u e o f 3.37 s -~ w a s o b t a i n e d for the e f f e c t i v e rate-
o f - d e f o r m a t i o n y i e l d i n g a p r o c e s s v i s c o s i t y v a l u e o f 4 . 8 6 Pa.s.

F i g u r e 9: L o c a t i o n o f the p r o c e s s v i s c o s i t y
323

It can be seen that the process viscosity is located along the impeller
blade and not in the gap between the ribbon and the wall, as it could be
expected with the Couette flow analogy. This is to our knowledge the first
time such a result is presented.
The numerical prediction of the variation of Kp(n) vs the shear-
thinning index is compared with the experimental results in Figure 10.

150

-.-e- numedcal 1
-..a-- experimental

100

:Z

50

.... ~ ..... i .... :, . . . . . . .


0 0.2 0.4 0.6 0.8 1

Figure 10: Variation of the power constant with the power law index

It can be seen that Kp(n) does not vary linearly with n. Indeed,
although the slope is constant above n=0.6 as already noted in [ 17], the
slope tends to decrease for higher shear-thinning fluids. The same trend is
observed numerically and experimentally, although to a lesser extent in the
latter case. In Figure 10, there is a discrepancy between the numerical
predictions and the experimental determination of Kp for the lower values
of n. The difference could be imputable to the larger numerical errors
inherent to the evaluation of stiff velocity gradients in the simulation.
We present in Figure 1 l a a plot of the normalized angular velocity
Vz/(Vr 2 q- Vz2)1/2. It corresponds to an isometric view of the cross section
plane intersecting the central shaft. Figure 1 lb shows a top view of the same
field at mid-height. The spectrum goes from dark grey (small angular
velocity) to light grey (high angular velocity). These two figures were
obtained for the CMC solution (1.5 wt.%, room temperature) at N= 10 RPM.
In Figure 11 a, the downward pumping zone appears clearly in the vicinity of
the shaft, where the rotational speed is negligible.
324

//
,jJ
Figure 11" Pumping zone of the HR impeller - a) side view; b) top view

In Figure 11, two pumping regions can be observed" a (downward)


pumping zone in the center and an (upward) pumping zone near the helical
ribbon impeller. The downward pumping zone is bounded by an off-
centered circular envelope which can be regarded qualitatively as the
Couette equivalent cylinder.

4. 2 Mixing flows in Dual helical ribbon impellers


We now consider the flow in a typical polymerization reactor
provided with a dual helical ribbon impeller (Figure 12). Usually, in
industry, these impellers are preferred over single helical ribbons due to
their improved performance. The simulation has been carried out using the
Galerkine finite element method described before and the Lagrangian frame
of reference approach. The surface finite element mesh is shown on Figure 6
as well.
We show in Figure 13 a typical circulation profile. The axial velocity
has been plotted at two horizontal cross sections. The scale is from-0.12
m/s (pumping down velocity) to 0.13 m/s (pumping up velocity). This
picture gives a fairly good representation of the circulation in the vessel.
The helical ribbon blades push the fluid at the top in the vessel
periphery while scraping the wall. The downward circulation region is
located along the shaft, the flow being mainly horizontal at the top (pointing
inward) and at the bottom (pointing outward).
325

Figure 12: Double HR impeller Figure 13" Vertical velocity profiles

Figures 14a and 14b represent a typical dispersion pattern


obtained with such an impeller. These results are obtained by tracking with
time the position of particles initially clustered and injected in the bulk in
the vicinity of the linking rod between the impeller and the shaft. The
trajectories are computed using an adaptive 4th order Runge-Kutta method
with time-variable steps. In Figure 14a, the "stretching-folding-breaking"
mechanism can be clearly identified. The particles initially agglomerated
gradually separate one from the next (like a widening ribbon) while turning
around the shaft during their travel towards the vessel bottom. In the bottom
region, they break apart and disperse in the bulk, circulating outward.
Figure 14b shows a similar behavior after a longer tracking time. The
dispersing-homogenizing capability of the impeller appears very clearly.,

Figure 14: Dispersion pattern- a) initial stage; b) homogenizing stage


326

Figure 15 shows the pumping pattern in the vessel from the top. This
figure is another way of looking at the circulation pattern. The pumping
down region (on the left) covers most of the region confined within the
helical ribbons while the pumping up phenomenon is restricted to the ribbon
region only. The "Couette" cylinder appears remarkably.

Figure 15: Downward (left) and upward (fight) circulation region

4.3 Mixing flows in planetary kneaders


Planetary kneaders can be seen as a special type of close-
clearance mixers. They are mixing systems of choice for pastes, thickener-
based fluids and for cross-linked polymers. The modeling of the kneading
flow is a new application of CFD which is receiving an increasing attention.
In this work, we consider the flow in a twin-blade planetary kneader from
APV. The system comprises a slow plain blade (twisted paddle) and a fast
hollow blade mounted off-centered on a rotating carousel (Figure 16).

Figure 16: Planetary kneader computer model


327

The carousel to plain blade speed ratio and plain blade to hollow
blade speed ratio are both 0.5 by construction. The results shown hereafter
were obtained with a carousel speed of 15 RPM, a plain blade speed of 30
RPM and an hollow blade speed of 60 RPM.
Contrary to the previous close-clearance impellers, the kinematics of
the blades is such that no convenient simplifying frame of reference can be
used. The virtual finite element method is then required to tackle the flow
simulation.
We show in Figures 17a and 17b the typical flowfield in the kneading
vessel for an industrial cross-linked polymer. In Figure 17a, a detailed flow
pattern in the region of the blade is shown. Although the blade has a twisted
shape similar to a screw, the circulation appears rather horizontal, an
indication of a poor top-to-bottom pumping. Figure 17b shows the general
circulation pattern at mid-height in the kneader. Intense shearing appears
between the blades and in the clearance between the blade tips and the wall.
Far from the blade, the flow is slow and almost stagnant near the wall.

. . . . . . . . . . . . .

"~ .'~ -.'~ .-, c ~. ,-- . . . . . . . ~ . . . .

Figure 17" Velocity profile in the kneading vessel

We show in figure 18 the dispersion pattern in the vessel. This result


was obtained in the same way as with the double helical ribbon impeller.
The clustered particles are initially injected at mid-height in the bulk close
to the region where the blades mutually interact in an intense shearing
region. It can be observed that the axial dispersion is limited to the central
core of the kneader, confirming the deficiency of the top-to-bottom
circulation observed in Figure 17. Radial dispersion appears, however, very
328

good as shown by the very chaotic trajectory patterns that covers most of the
cross-section surface of the vessel. This proves if necessary the relevance of
using more than one impeller to break the flow symmetry in viscous mixing,
and therefore enhance the process efficiency.

Figure 18: Dispersion pattern; left: side view; fight: top view.

5. CONCLUSION

In this chapter, we have shown how the design of mixers and


kneading equipment for non-Newtonian viscous fluids can be carried out
using published correlations and/or numerical techniques. It is now possible
to predict mixing performance during the process development phase,
before building up equipment. It is worth mentioning that this field is still
knowing significant developments, and it is expected that more complex
problems involving multiphase systems and more complex rheology will be
tackled in the near future.

ACKNOWLEDGMENTS

The technical contribution of Dr. F. Bertrand and Mr F. Thibault from Ecole


Polytechnique is gratefully acknowledged. Thanks are also directed to
DGAPA (UNAM) and NSERC for the financial support that made possible
several contributions in this chapter.
329

REFERENCES

~ Ottino J., 1989, The Kinematics of Mixing, Cambridge Univ. Press.


2. Skelland A.H.P., 1967, Non-Newtonian Flow and Heat Transfer,
Wiley.
. Eklund D.E. and J.E. Teirfolk, 1981, TAPPI J., 64, 63.
4. Solomon J., T.P. Elson, A.W. Nienow and G.W. Pace, 1981, Chem.
Eng. Comm., 11, 43.
. Barnes H.A., 1997, J. Non-Newt. Fluid Mech., 70, 1.
6. Rauline D., P.A. Tanguy and P.J. Carreau, Mixing of Thixotropic
Fluids, in preparation.
. Hocker H., G. Langer and U. Werner, 1981, German Chem. Eng., 4,
133.
~ Brodkey R.S., 1967, The Phenomena of Fluid Motions, The Ohio
State University Press, Columbus.
. Metzner A.B. and R.E. Otto, 1957, AIChE J, 3, 3.
10. Doraiswamy D., R.K. Grenville and A.W. Etchells, Ind. Eng. Chem.
Res., 1994, 33, 2253.
11. Brito-de la Fuente E., J.C. Leuliet, L. Choplin and P.A. Tanguy,
1992, AIChE Symp. Ser., 286, 28.
12. Brito-de la Fuente E., L. Choplin and P.A. Tanguy, 1997, Trans.
IChem. E., 75, 45.
13. Carreau P.J., R.P. Chhabra and J. Cheng, 1993, AIChE J., 39, 1421.
14. Bourne J.R. and H. Buffer, 1969, Trans. IChem. E., 47, T263.
15. Coyle C.K., H.E. Hirschland, B.J. Michel and J.Y. Oldshue, 1970,
AIChE J., 16, 903.
16. Hall K.R. and J.C. Godfrey, 1970, Trans. IChem. E., 48, T201.
17. Chavan V.V. and J. Ulbrecht, 1973, Ind. Eng. Chem. Process Des.
Develop., 12, 472.
18. Rieger F. and V. Novak, 1973, Trans. IChem. E., 51,105.
19. Nagata, S., 1975, Mixing" Principles and Applications, Kodansha and
Wiley.
20. Sawinski J., G. Havas and A. Deak, 1976, Chem. Eng. Sci., 31,507.
21. Chowdhury R. and K.K. Tiwari, 1979, Ind. Eng. Chem. Process Des.
Dev., 18, 227.
22. Yap, C.Y., W.I. Patterson and P.J. Carreau, 1979, AIChE J., 25, 516.
23. Kuriyama M., K. Arai and S. Saito, 1983, J. Chem. Eng. Japan, 16,
489.
24. Shamlou P.A. and M.F. Edwards, 1985, Chem. Eng. Sci., 40, 1773.
330

25. Bourne J.R. and H. Butler, 1969, Trans. IChem. E., 47, T11.
26. Brito-de la Fuente E., L. Choplin, A. Tecante, and P.A. Tanguy, 1994,
in Progress and Trends in Rheology IV, C. Gallegos, J. Munoz and
M. Berjano (Eds.), Darmstad-Steinkopff, Germany, 272.
27. Brito-De La Fuente E., J.A. Nava, L. Medina, G. Ascanio and P.A.
Tanguy, 1996, Proc. XII International Congress on Rheology, A. Ait-
Kadi, J.M. Dealy, D.F. James and M.C. Williams (eds.), Canadian
Rheology Group, 672.
28. Patterson W.I., P.J. Carreau and C.Y. Yap, 1979, AIChE J., 25, 508.
29. Tatterson G.B., 1991, Fluid Mixing and Gas Dispersion in Agitated
Tanks, McGraw-Hill.
30. Tanguy P.A., R. Lacroix, F. Bertrand, L. Choplin and E. B rito-De La
Fuente, 1992, AIChE J., 38, 939.
31. Bertrand F., M. Gadbois and P.A. Tanguy, 1992, Int. J. Num. Meth.
Eng., 33, 1251.
32. Tanguy P., L. Choplin and M. Fortin, 1984, Int. J. Num. Meth. Fluids,
4, 441.
33. Bertrand F, P. A. Tanguy and F. Thibault, 1997, Int. J. Num. Meth.
Fluids, 25, 719.
331

VISCOELASTIC FINITE VOLUME METHOD


N . P h a n - T h i e n a n d R.I. T a n n e r

Department of Mechanical and Mechatronic Engineering


The University of Sydney
NSW 2006, Australia

1. I N T R O D U C T I O N

Analytic solutions to non-trivial viscoelastic flow problems are rare due


to the complexities of the constitutive equations and the nonlinearities of
the conservation equations. To make any progress, we have to abandon
the search for the analytic solution and seek an approximate solution via a
numerical method, which is either a finite difference (FDM), a finite volume
(FVM), a finite element (FEM) or a boundary element method (BEM);
see for example, Crochet, Davies and Walters [1], Reddy and Gartling
[2], Phan-Thien and Kim [3] and Patankar [4]. Earlier numerical schemes
failed to converge at a relatively low level of flow elasticity when elastic
effects become comparable with viscous effects [5]. The level of the flow
elasticity is characterised by either the Weissenberg number Wi (product
of a characteristic fluid relaxation time and a typical shear rate), or the
Deborah number De (the ratio of a characteristic fluid relaxation time
to a characteristic time scale for the flow, which is usually taken to be the
reciprocal of the wall shear rate in a fully developed flow region).
Recent progress in FEM, such as the streamline-upwind (SU) [6], the
explicitly elliptic momentum equation (EEME) [7], the elastic-viscous split
stress (EVSS) formulation [8], and the adaptive viscoelastic stress splitting
(AVSS) scheme [9], have been reviewed elsewhere in this book. Therefore
we will be concerned with the Finite Volume Method here, which has been
very popular in high-Reynolds number Newtonian flows [4], but only made
its presence felt in computational viscoelastic fluid mechanics recently. Our
review begins with the formulation, the implementation, and some two and
three-dimensional problems that we have had recent success with, including
the channel flow past a cylinder, and the three-dimensional entry flow.
332

2. F O R M U L A T I O N

2.1 Governing Equations


We are concerned with a general time dependent and isothermal flow
of an incompressible viscoelastic fluid where the governing equations take
the form:
C o n t i n u i t y e q u a t i o n (conservation of mass)
V-u--O, (1)
M o m e n t u m e q u a t i o n s (conservation of linear momentum in absence of
body forces)

p ( ~ - + V- (uu)) - V . a , (2)

where t is the time, u is the velocity vector with components {u, v, w}, and
a the Cauchy stress tensor, given by
a = - p l + S, (3)
with p being the hydrostatic pressure, 1 the unit tensor, and S the "ex-
tra" stress tensor (not necessarily traceless), which is related to kinematic
quantities by an appropriate constitutive equation.
C o n s t i t u t i v e equations
In most studies, either a differential (Maxwell-type) or an integral consti-
tutive equation is used to model the fluid rheology. Integral constitutive
equations are less popular because of the need of particle tracking. A
review of constitutive equations, covering both microstructural and con-
tinuum view points is given in Huilgol and Phan-Thien [10]. Here, we will
be mainly concerned with the P T T model [10, 11], where the extra stress
tensor is written as
S - 2~ND + 7", (4)
where ~N is the Newtonian-contribution viscosity, D - (Vu + Vu T)/2 is
the strain rate tensor, with T denoting the transpose operation, and 7- is
the polymer_contribution stress, which evolves according to the following
constitutive equation

g r + A v = 2r/m0D' (5)
333

where A is the fluid relaxation time, r/m0 is the "polymer-contributed vis-


cosity", (.v) represents the following convected derivative:

v Or
v - ~ + V - ( u ~ ' ) - (L - ~D)7" - v (L - ~D) T , (6)

and
Ae
g = 1 +--tr (r), (7)
r/m0

where L - V u T is the velocity gradient tensor, and ~, e the material


parameters for the P T T model [11].
By introducing the retardation ratio defined as/3 - ~,n0/~0 with r/0 =
7/N + 7/m0 being the total viscosity, the Oldroyd fluid B [12] (when ~ and e
are set to zero) is recovered with the retardation time being A2 - (1 - / 3 ) A;
in addition, the Upper Convected Maxwell (UCM) model is recovered with
/3=1.
The Reynolds number R~ defined as R~ - p U h / ~ o , and the Weissenberg
number W~, defined as We = A U / h , are used to characterise the flow and
the fluid elasticity, where h and U are typical length and velocity scale
in the flow, respectively. To compare numerical to experimental results, a
zero shear rate relaxation time, defined by
N1 ('~)
2r
(8)
is used for the experimental fluid. The Deborah number D e - A ~ based
on the maximum wall shear rate "~w in the full developed downstream
channel is sometimes used instead of W~. As an example, for the planar
contraction flow (constant viscosity elastic fluids [13]), this is thrice We
when the upstream aspect ratio of the duct is sufficiently large.

2.2 Split stresses


Our numerical experimentation leads us to adopt the EVSS formulation
of Rajagopalan et el. [8]. That is, the extra stress (Eqn. (4)) is re-written
as:

Sij - 271odij + Y]ij. (9)


Substituting (9)into (3) and (5)yields
OUi 0 ( OUi~ OP OY]ik
334

gEij + A ~ i j - 2~7/0 [(1 - g)dij - A ~ij (11)

Thus the dependent variables are now ui, p and Eij.


With this coupling between the kinematics and the stresses, we find
that the kinematics calculations are less sensitive to the gradual loading of
elasticity via the pseudo-body force term OEik/OXk, thereby improving the
stability of the calculations.

2.3 B o u n d a r y C o n d i t i o n s
To the set of governing equations one must add a set of relevant bound-
ary and initial conditions.
Slip or no-slip on a solid surface
The no-slip boundary condition at a solid surface is usually adopted in
most studies, where the fluid velocity assumes the velocity of the solid
surface. This assumption works well for viscous fluids, but there is a large
amount of experimental data suggesting that it may not be relevant for
polymeric liquids in some circumstances. There are extrusion experiments
with polymer melts [14-18], which suggest that wall slip may be responsible
for melt fracture. In these experiments, the occurrence of the extrudate
irregularities occurs above a critical wall shear stress, which is accompanied
by a fluctuation in the pressure drop. A phenomenological approach to the
slip boundary condition has been proposed by Pearson and Petrie [19]
where the slip velocity is taken as an empirical function of the wall shear
stress. A polymer network model has been proposed recently to account
for the dynamic slip velocity [20]. Here, we just simply adopted the no-slip
boundary condition at a solid surface, and note that real progress in this
area will be made by a careful consideration of the microstructure near a
solid surface. With this (Dirichlet) boundary condition, the velocity field
is prescribed as
u=u0, OnSu,
where u0 is known on the boundary S~.
Traction boundary conditions
Sometimes the traction vector is given on a part of the boundary, say St.
This type of (Neumann) boundary conditions take the form
t=er-n=t0, on St,
where t is the traction vector. This yields the normal traction,
tn " - - n - O ' - n -- --p + S- n = n - to, o n St, (12)
335

and the tangential tractions


t-t-nn=t0-t0-nn, on St.
Note that we use Su to denote parts of the boundary where velocity is
prescribed, and St, parts of the boundary where the traction is prescribed.
These surfaces need not be singly connected, but may consist of several
patches. Indeed both velocity and traction boundary conditions can be
prescribed at a given location, but in different directions, as in the case of
the extrusion problem. Robbins boundary conditions arise in some slip-
stick problems, where the slippage velocity is determined by the tangential
traction (shear stress) at the wall.
Free surface boundary conditions
For problems with a free surface, e.g., bubbles, extrusion, etc., the location
of the free surface is not known and must be found as part of the solution
procedure. Here, a kinematic constraint can be used as the condition
to locate (implicitly) the free surface. In a steady flow, the kinematic
constraint for a free surface is
u-n=0. (13)
For unsteady flow, if r t) = 0 is the location of the free surface, then
always remains zero on the free surface and its material derivative must
also be zero there. The free-surface kinematic constraint thus takes the
form
ar
0--7 + u - V r - 0. (14)

Since n - + V r is a normal unit vector on the free surface (the sign


can be chosen so that n is the outward normal unit vector), this constraint
also takes the form
1 0r
+ u.. = 0. (15)
IVr a t
In addition to this, the traction on the free surface is known from the
physics of the problem. If there is no surface tension, for example, then the
traction vector is zero on the free surface. For the case where the surface
tension is not negligible, we recall that the surface tension 9/is postulated
to be the force per unit length acting along the edges of the free surface.
The equilibrium condition on an arbitrary surface element A S reads

fAS [er]. n dS + fAC "~q dl -- 0,


336

where [er] 9n - (er + - er-) 9n is the jump in the normal traction, with er +
being the stress on the positive side of n, and or- the stress on the negative
side of n, and q is the unit vector normal to the boundary curve AC, but
tangential to the interface AS. This is the mathematical statement of zero
force on AS. From Stokes' theorem, the surface integral over A C can be
done by parts,

/~c7 q dl - /~sV7 d S - /As7 (V-n)n dS,


leading to

(p- - p + ) n + (S + - S - ) . n + V 7 - 7 ( V . n ) n - 0. (16)
Note that the mean curvature of the surface is given by [21]
1 1
V n . . . ~~
. . (17)
R1 R2
Symmetry boundary conditions
Very often the computational domain can be reduced with the help of
symmetry, these sylmnetry boundary conditions are quite easy to state: a
plane of symmetry has no normal velocity component, and no tangential
stress (tangential tractions or vorticity are zero), i.e.,
u-n=O, t-t.nn=O, (18)
where t is the surface traction.
Inflow outflow boundary conditions
At the inflow to the solution domain, we know something about the flow,
usually the velocity field. For viscoelastic fluids, the stresses are also re-
quired there (in essence, these represent the information carried with the
fluid from its previous deformation history). Therefore, the boundary con-
ditions at the inlet are usually Dirichlet boundary conditions. Note that all
components of the stress cannot be arbitrarily prescribed. First, the stress
components should be consistent with the specified kinematics (but need
not be), otherwise steep stress boundary layers will be set up. Secondly, if
traction boundary conditions are also given there, then the traction com-
ponent tangential to the boundary would involve the stress components
alone without the pressure terms, and therefore these stress components
cannot be prescribed arbitrarily. Finally, Renardy [22] has pointed out
that specifying all the stress components at the inflow may lead to an
over-determined mathematical problem. It may be prudent in practice to
retain the time derivative terms in the (differential) constitutive equations,
337

even though we may be dealing with steady-state solutions, and integrate


the equations in time through an initial stress state until a steady state
solution is reached, including the stress components at the inflow. This
ensures that a physical solution is obtained.
At the outlet, the flow is usually well developed and arranged so that a
uni-directional flow results. Outflow boundary conditions therefore usually
take the form of no transverse velocity and no axial traction.
Boundary Conditions for the Pressure
Boundary conditions on the traction vector arise naturally from a weighted
residual method. However, in the finite volume method the equation for
the pressure is usually solved separately from the velocity field, by using
the continuity constraint on the linear momentum equations, and therefore
boundary conditions for the pressure may be required.
The correct boundary conditions for the pressure can be derived either
from the traction boundary condition, or from the momentum equations.
Thus, a traction boundary condition (to) will turn into a Dirichlet bound-
ary condition for the pressure:
p=S:nn-t0.n, on St. (19)
Otherwise, a Neumann boundary condition for the pressure will result from
the momentum equation, if the kinematics are fully prescribed,

= '1 -p + u. Vu + v. s + pb , on (20)

Note that a traction boundary condition at one point on the boundary


will implicitly set the pressure. If there is no traction boundary condition,
then the pressure can only be determined up to an arbitrary constant. It
is important to keep in mind that a traction boundary condition is not the
same as the pressure boundary condition; the former is a physical quantity
that we can actually impose on the fluid, the latter is a derived quantity,
arising only because we are interested in solving the Poisson's equation for
p in isolation. If the set of equations, continuity, momentum, are solved
jointly, then p would inherit the correct boundary conditions from the
boundary conditions for u and t, and there is no need to impose anything
on p at all.
Initial Conditions
For time-dependent flows, a set of initial conditions is required. The initial
velocity prescribed must be divergence free. For viscoelastic fluids, a set
338

of initial values for the stress components is also needed. Note that the in-
compressibility constraint also forbids an impulsive start and stop, normal
to the boundary, since

fs u.n dS-O.

3. F I N I T E V O L U M E M E T H O D (FVM)

The finite volume method has been very popular in computational fluid
dynamics dealing with Euler and Navier-Stokes flow problems. There are
two main approaches in the finite volume method.

3.1 Chorin-Type Methods


The first approach uses an artificial compressibility condition to satisfy
the continuity equation. A pseudo-transient formulation is then adopted in
the momentum equations and the steady state solution is considered as the
asymptotic solution of a time-dependent problem with time-independent
boundary conditions (if need be), and these are computed by a time-
marching scheme. The method is due to Chorin [23] and is very com-
monly used in computational fluid dynamics dealing with high Reynolds
number flows. Integration in time can be carried out either implicitly
or explicitly. In implicit schemes, it is necessary to solve a system of
equations at each time step, making the scheme very expensive for large
three-dimensional problems. However, such schemes can be made uncon-
ditionally stable and relatively large time steps can be used. On the other
hand, in explicit schemes, it may be necessary to solve only a simple mass
matrix, or it may not be necessary to solve any system of equations at
all. The main disadvantage of explicit schemes is that the time step is
restricted by the Courant-Friedrichs-Levy stability condition. However,
there are several ways to accelerate convergence such as residual averag-
ing, local time-stepping and multigrid schemes. Phelan et al. [24] use a
similar scheme, but with a modification to suit hyperbolic types of con-
stitutive equations, for solving cavity-driven flow problems for the UCM
model. The same method has also been used to solve the slip-stick problem
of Maxwell-type constitutive equations [25].

3.2 S I M P L E R - T y p e Methods
The second approach in finite volume method is much more popular and
forms the basis of some commercial packages dealing with Navier-Stokes
equations. The resulting algorithms are known as SIMPLE (Semi-Implicit
Method for Pressure-Linked Equations), SIMPLER (SIMPLE Revised) [4],
339

SIMPLEC (SIMPLE Consistent) [26], or PISO (Pressure-Implicit with


Splitting of Operators) [27], to name a few. In these schemes, the mo-
mentum equations are solved sequentially, and the pressure is corrected
from the continuity, equation and the linearised momentum equations.
Applications of the SIMPLER algorithm to viscoelastic flows include the
4-to-1 entry flow and the die entry problems [28-30], the non-isothermal
flow past a cylinder [31], the flow past an eccentric cylinder problem [32].
The PISO method has been applied to the non-circular pipe flow of the
Criminale-Ericksen-Filbey model to study the effects of the weak secondary
flow on the pressure drop [33]. A variant of the SIMPLER method, termed
SIMPLEST (SIMPLE with Splitting Technique) has been used recently
to study secondary flows in non-circular pipe flows, and the 4-to-1 three-
dimensional flows [30, 34].
Formulation
In the FVM, the flow domain is subdivided into a set of non-overlapping
control volumes (CV). The grid nodes (storage locations for the dependent
variables) are located at the centre of each of the control volumes. The
discretised equations of the dependent variables are obtained by integrating
the mass, momentum conservations over each control volume as follows:

1. First, all of the governing equations are cast into the form of a general
transport equation:
a
AO(I)
-b-/- + 0 (Auk,I)) -
k +
where (I) is the dependent variable which can be a component of a
vector or a tensor and even a constant. The coefficients A, F have
different meanings for each different dependent variable, and Sr is
called the source term, which lumps all the terms that cannot be
accommodated in the convective and diffusion terms, and is specific
to the particular variable (I).
2. Next, the general transport equation is integrated over the control
volume surrounding the node, namely node P, in the flow domain and
the time interval St, using the divergence theorem whenever possible,
340

Here, A V is the volume of the control volume and ~ v the correspond-


ing face area with n k being the unit outward normal to the face.
3. Finally, a proper discretisation scheme (the control volume-based in-
terpolation functions) is adopted for the temporal and spatial approx-
imations of the dependent variable.

Generally, the first-order backward Euler implicit formula is used for


temporal differences because of its simplicity and its unconditional stability
for numerical calculations.
As to the volume integral of the source term S~, the value at the central
node of the control volume, namely S~, is assumed to prevail over the
whole control volume, and it can be linearised in term of the nodal value
Op as usually assumed in FVM.
In this way, the final algebraic approximation equation which relates Op
to its neighbouring nodal values can be written as
ap~p -- ~ anb~nb -4- Sc
- --[-ape~p,
o - (23)
nb
where
o_ A o -
ap ~-~AV, ap -- E anb + ap -- Sp. (24)
nb
Here, the subscript p refers to the central node P, and the summation is
to be taken over all of the neighbor nodes nb of the node P. An overbar
means the applied values are evaluated using the known fields from the
previous time (or iteration) level, and the coefficients anb are the functions
of the dependent variables, and their structures depend on both the form
of the control volume chosen and the approximation scheme used. It is
these coefficients that determine the spatial accuracy of the final solution.
In our calculations, the power law scheme proposed by Patankar [4] is em-
ployed for the formulation because it covers central difference and upwind
difference, and gives an excellent approximation to the exact exponential
solution for the one-dimensional convective-diffusive equation. In addition,
FVM with power law scheme, instead of central or upwind difference, has
better conservational properties and thus produces physically realistic so-
lutions even for coarse meshes. In the power law scheme, the coefficients
are chosen as follows:

anb -- D n b f ([Pnbl) + [sign(nb)Fnb, 0~ , (25)


where Dnb -- (FA/ Xi)nb is the local diffusion conductance; Fnb -- (Au~A)nb
the "mass" flux passing thought the corresponding face A normal to i
341

direction of the control volume, sign(rib) is +1 for upstream face and - 1


for downstream face, and f (I Pnb I) is the function of the local Peclet number
defined by Pub -- Fnb/Dnb, which is given by

f ([Pub[) -- [0, (1 -- 0.11Pnbl)5], (26)


where the symbol [a, b~ means the greater of a and b.
Constitutive equation discretisation
The viscoelastic model is also in the form (21), without the diffusion term
(F - 0). To ensure numerical stability a first-order upwind difference is
used for spatial discretisation. Thus, the discretised constitutive equation
will take the similar form to (23) (thereafter we use the symbol "1- instead
of E for the elastic stress tensor Eij 2flTIodij to avoid confusing with
-- 7"ij --

the summation symbol E)"


ij ij ij -
ap Tp -- E anbTnb -~- Sc ~, (27)
nb

where the superscripts ij refer to tensor components while reserving sub-


scripts for the grid node and the overbar for the values from the previous
time (or iteration) level. The constant part of the source term, in which
the stresses and deformation are approximately piecewise-constant in each
control volume, takes the form (no sum in i and j)"

S:~ - ap~'p~ + A V {2/3~0 (1 - g)dp j + )~ (1 - ~)[(dpi + d jj) ,0] CpJ


+ )~ [~k (1 - 5kj)+ 2/3~0d~k] [(OuJ I (28)
\Oxk/p

\Oxk]p
0

while the coefficients ai.j and ap0 take a form like (24) with F - 0, A - A,
and
SpJ - - A V [g )~ (1 - ~ ) [ - (dpi + dJpJ), 0]] - E sign(nb)Fnb, (29)
nb

where Fnb -- ()~uiA)nb can b e thought of being the "mass" flux passing
through the corresponding face A normal to i direction of the control vol-
ume. The term Esign(nb)Fnb in 5'~J is actually the continuity constraint
nb
which should approach zero when the solution converges.
342

We note that there is an extra convective term in the source term in


the constitutive equation; this extra source term should be consistent with
that of the convective term of the stresses. Thus, upwind difference should
be applied to its discretisation. This leads to
0

With upwind difference, the resulting coefficients anb will be


anb -- [sig~(nb)Fn~, 0]. (31)
Therefore, the sufficient condition for convergence for the adopted nu-
merical algorithm-TDMA (typically alpj > E anb) is satisfied even for the
nb
steady-state calculation with a ~ being zero. Some work [35] used a pseudo-
transient constitutive equation for the steady-state calculation to ensure
a resulting diagonally dominant matrix, and so obtained convergent so-
lutions for the inertialess flow of UCM fluid (g - 1) through a 4 9 1
abrupt axi-symmetric contraction with a much higher value of the Weis-
senberg number (up to W i - 6.25). The pseudo-transient method for the
steady-state calculation is actually equivalent to a local under relaxation:
the positive inertia coefficient ap0 damps out possible oscillation during the
stress iteration.
In our previous work [32, 34, 36], an artificial diffusion term -g~-2~k\ ox~]
with c~ being an artificial diffusion coefficient is introduced on both sides
of the constitutive equation, and in discretisation, the current values are
taken for Tij on the left hand side; while the known values from previous
iteration level for that on the right hand side. As a result, depending on
the spatial discretisation scheme used, the coefficients anb in (27) will take
the form

anb- Dnbf ~ -4- [-sign(nb)Fnb, 0], (32)

where Dub -- ((~A/6xi)nb denotes the local diffusion conductance, and func-
tion f ( I D ~ b l ) h a s the different form for different schemes. For example,
\1 I/

when the central-difference scheme is adopted, we have

1 05, ,
343

Thus, with the scheme adopted, the convective term of the equation
can be thought of being always discretised using upwind difference, and
the diffusion term is discretised using central difference, but the resulting
local diffusion conductances are different for different schemes used, that
is
' + [sign(nb)Fnb, O~ - anb
anb -- Dub d + anb
~ , (34)

with D'nb -- D ~ b - 0.5 IFnbl for central difference and D'nb -- Dub for upwind
difference. Here, we use anb d to denote the contribution of the artificial
diffusion term discretisation, and a~b the contribution of the convective
term discretisation, that is:
d _ D n' b ~ anb
anb c _ [sign(nb)Fn~, O~ (35)
To ensure all of the contributions to be positive, we need
!

Dnb > O. (36)


Obviously, this is always the case when upwind difference is used. However,
for the case when central difference is employed, we require that
Dub ~ 0.5 IFnbl. (37)
Thus, we have
(
Dub-- 1-- ~
1) Dub, (38)

with Dub -- IFnbl and ~; is an arbitrary constant satisfying ~; _> 0.5. This can
be thought of being the criterion for the proper selection of the artificial
diffusion coefficient c~.
Similarly, the source term will contain an extra artificial diffusion term
_ f z x u _ ~ko \(c~~176 dV, and central difference is applied for its discretisation.
The discretised form for the extra source term will have the form:
~dj -- _ y~ DnbTnb
_ij + "fp3 ~ Dub. (39)
nb nb
Thus, the final discretised constitutive equation can be cast into the
form
(ap T ad T apO_ ~ipj) ~ "" _ E (anb
c + adnb) Tub
ij + s~J 2t- ~d
-"j, (40)
nb
where
apC :~-~ anb
C
and a d :~--~ anb,d (41)
nb nb
344

which can be re-arranged to yield


I v " _c ij __ O~J
.... (ap + a ~ - ~ipj) n~ ttnbTnb-1- ~c
- +A, (42)
+ [(a; + aO - + )

where
~ d _.i j ~.d _--dj q_ ~adJ
Y~ t.nb.nb ~ t~p'i p
/~ __ nb (43)
(a~ + a ~ - ~ J ) + a d

will tend to zero when the solution converges.


Therefore, by introducing an artificial diffusion term, the spatial accu-
racy of the solution is less than second-order, but an effective local under-
relaxation factor
ap + ap0 _ ~ j (44)
OiPJ -- (a~ + a ~ - ~pJ) + ~ D'nb

is introduced for the stress calculations so that possible oscillation during


iteration is alleviated.
Structured and Unstructured M e s h

W w_ , ' ..O" / _ G e E
- O- - ,-s- =0- - - I - @
""'P [

Figure 1" The control volume for node P.

In the S I M P L E R method and most of its invariants, a staggered mesh is


invariably used, where the pressure nodes (and the nodes for the extra
stress components), are positioned along x, y, and z directions first, then
the control volume faces, at the centroids of which the velocity components
are calculated, are placed midway between neighbouring grid points. Thus
the x-direction velocity u is calculated at the faces that are normal to the
x-direction. The same rule is applied to v and w. An example of a control
volume for node P, with neighbouring nodes U, D, N, S, E and W is
illustrated in Fig. 1.
345

Figure 2" An unstructured finite volume mesh for a flow past a cylinder
in a channel.
In the figure, the lower case letter refers to the interface between two
neighbouring cells, e.g., e refers to the interface between the control vol-
umes for P and for E. One ends up having four different types of control
volumes, three for the velocity components, and one for the pressure. The
staggered grid is designed to eliminate the checker-board pattern in the
pressure field [4]. With structured mesh, a linear interpolation scheme is
often used, resulting in an O(h) accuracy, where h is a typical linear di-
mension of the control volume. Although one ends up with four different
types of control volumes (three for the three velocity components and one
for the pressure), structured mesh has been very popular because of the
ease in implementation; in addition, iterative solvers like the tri-diagonal
Thomas algorithm (TDMA) can be readily applied along the coordinate
lines defining the mesh. Structured mesh is not very suitable for complex
geometry, however.
To provide the flexibility in fitting complex computational domains, un-
structured non-staggered triangle mesh have been developed by a number
of authors, for example, Prakash and Patankar [37], Masson et al. [38],
and Davidson [39]. An example for such an unstructured mesh for the
two-dimensional flow past a cylinder in a channel is given in Fig. 2. To
retain the basic structure of the iterative line-by-line tri-diagonal Thomas
solver (TDMA), lines of nodes from the entry to the exit boundaries are
introduced (Huang et al. [32]). If this is not possible, then the line is ter-
minated in the domain. The nodal information (neighbouring ID's, etc.) is
stored in a sweeping line array. Similar to a re-ordering of the nodes in the
finite element method to minimise the bandwidth of the system matrix,
this array stores all the information needed for the line-by-line tridiagonal-
matrix algorithm, which only requires computer storage and computer time
346

of O(N), where N is the number of unknowns. The method has been used
successfully on viscoelastic flows between two eccentric cylinders [32], and
past a cylinder in a channel [36].
Solver
In the S I M P L E R method and most of its variants, the kinematics are
determined by solving the momentum equations, assuming that the pres-
sure field in known. The pressure correction is then applied, by enforcing
mass conservation [4]. For the viscoelastic flow computations, the source
terms containing the extra stress OEik/OXk in the momentum equations are
treated as pseudo-body forces with the known dynamics field obtained from
the previous time (or iteration) level by solving the discretised constitutive
equations.
In each cycle of the algorithm, no system matrix needs to be solved, and
the coupled discretised equations for the dependent variables are solved se-
quentially from an initial guess for all field variables (typically quiescent
field for Newtonian computations). For the momentum equations, two
T D M A sweeps are performed, and for pressure correction and stress equa-
tions, four T D M A sweeps. In the iterative procedure, the calculations of
velocities and stresses are under-relaxed by a global factor of 0.85 ~ 0.5
depending on the elastic level, but no relaxation is need for the pressure
calculation.
The convergence criterion for terminating the calculation is that the
integral residual of the discretised equations over all control volumes for
any dependent variable is less than the input tolerance, of the order 10 -6
10 -s, and the relative changes in the values of flow field (typically the
velocities) near the solid wall from one iteration to the next are of the
order 10 -5 ~ 10 -7.
For viscoelastic computations, the corresponding Newtonian result (A - 0)
is used as the initial guessed field, and Wi is increased gradually by in-
creasing A. The under-relaxation mechanism due to adding artificial dif-
fusion is very effective in the stabilisation of the calculations at high elas-
ticity. To speed up the convergence rate and save some C P U time, a
three-dimensional block T D M A solver can be used, see [30].
The two-dimensional implementation was done with both structured
and unstructured mesh. The three-dimensional version was implemented
with structured mesh only, as three-dimensional automatic unstructured
mesh generation is still an active area of research. Both versions have
been well tested with simple flow problems where analytical solutions exist,
including Couette and Poiseuille flows, and proved to be very robust. In
the 2D Poiseuille flow, there seems to be no upper limiting Weissenberg
347

number, and an agreement to four significant figures with analytical results


was demonstrated at We = 40 in [36].
The 3D version has been tested for the Poiseuille flow of the Oldroyd-B
fluid in a straight pipe with square cross section. Numerical results show
that no upper limit in the Weissenberg number is encountered, and the con-
vergence is improved with mesh refinement. It is verified that adding the
artificial diffusion has a similar function to increasing the under relaxation
factor [34].
We now consider the flow past a cylinder in a channel, and the three-
dimensional entry flow.

4. F L O W P A S T A C Y L I N D E R

The flow past a cylinder in a channel has been well investigated, both
experimentally and numerically, and therefore is a good candidate to test
the finite volume method. Here, the flow past a cylinder, of radius R,
placed on the centreline, or offset from the centreline by a distance e in a
channel of width 2L, is considered. No-slip boundary conditions are applied
at the walls. Along the centreline symmetry conditions are applied. The
length of the downstream section is 9L, and that of the upstream section
is 6L. All distances are normalised by cylinder radius R. Having fixed
the model parameters and the problem geometry, the only parameter left
to vary is the average velocity U of the fluid at the channel entry. The
Deborah number (De) and Reynolds number (Re) are defined as
De = AU/R, Re = pRU/Vo. (45)
Dhahir and Walters [40] reported some experiments and finite element
calculations (using Polyflow TM) on the effects of viscoelasticity and of the
wall on the flow of non-Newtonian liquids past a cylinder confined by two
plane walls, with particular attention to the drag force on the cylinder. In-
ertia effects were neglected in the calculations, and they found the stream-
lines nearly independent of the mean flow rate. In addition, they also found
the streamlines to be essentially independent of the Weissenberg number,
at low values of the latter.
Flows past an asymmetrically confined cylinder have also been investi-
gated and the effect of the eccentricity on the drag force has been obtained
both numerically and experimentally. Baaijens et al. [41] investigated the
same flow of a shear thinning solution PIB/C14 both numerically and ex-
perimentally. A number of constitutive equations were used in their FEM
simulations, including the PTT, the Giesekus and the UCM models. The
field variables were compared to experimental values, along the centre line
348

and over cross sections of the channel. In general, there was a good agree-
ment between measured and computed field variables. Measured and com-
puted birefringence was also presented in [42] and a qualitative agreement
was found. The wall effects were considered by placing the cylinder off
the centreline, and it was found that the elongational thickening of the
viscoelastic fluid causes a significantly larger flow rate through the broader
gap compared with inelastic fluids. Barakos and Mitsoulis [43] investigated
viscoelastic flow past a cylinder symmetrically confined by two parallel
plates using a K-BKZ integral constitutive equation, using FEM with a
path tracking scheme. A good agreement between their results and those
of [41] was found. Huang and Feng [44] also reported some numerical re-
sults, using PolyflowTM, for the flow of the Oldroyd-B fluid past a confined
and non-confined cylinder, retaining inertial terms.
In the finite volume simulation, the computational domain is divided
into non-overlapping polygonal control volumes, as shown in Fig. 2, and
the SIMPLER algorithm is adopted for the UCM model.
First, the accuracy of the Newtonian calculations is assessed by com-
paring the drag coefficient defined by

K=
4~0UR'
where F~ is the axial force acting on the cylinder, with mesh refinement.
The extensive mesh refinement studies are performed with different meshes
by reducing both the polar mesh size hp and the maximum mesh size hm
according to Table 1. By comparing the results within and across mesh
groups, the refinement of the polar mesh size is found to play a more
important role in producing a more accurate solution. In the table, the
errors are calculated relative to the result of the finest mesh L4.
Polar Maximum
Mesh Nodes Drag Force Error
Mesh Size Mesh Size
M2 0.069813 0.4 1067 10.0634 0.03740
M4 0.0349065 0.2222 3434 10.3178 0.01306
M6 0.02617 0.1538 6504 10.3060 0.01419
M7 0.02416 0.1818 7711 10.2927 0.01546
N1 0.021 0.2566 2392 10.3260 0.01228
N2 0.0175 0.2566 2806 10.3313 0.01177
N3 0.00873 0.2566 4296 10.3513 9.86e-3
N4 0.00714 0.2566 4996 10.4172 3.55e-3
L1 0.021 0.127 6991 10.4306 2.27e-3
L3 0.00873 0.127 9695 10.4543 8.19e-7
L4 0.00714 0.127 10977 10.4543 0
349

Table 1" Drag force and relative error computed for a Newtonian fluid.

3000

2500 ..... FEM . 0 6 s


O uCVM 0.0s !
2000
O,I

E
Vo 1500

LI_ 1000

500

% 30 40 50 60 70 80
FLOWRATE (cma/s)
Figure 3: A comparison between the FVM (symbols) and the FEM
results of Dhahir and Walters (solid and dotted lines).

The best fit polynomial to the calculation of K is


K(hp) - 10.5245 - 6.4120hp,
showing an approximately linear convergence with decreasing mesh size.
An extrapolation of this formula gives K - 10.5245 at zero mesh size,
which is comparable to the FEM result of 10.5313 [45], and asymptotic
result of 11.0199 [46].
For viscoelastic flows, mesh refinement is the only check of the con-
sistency and accuracy of the numerical results. A monotonic decrease in
drag force with the decreasing polar mesh size is found, with the maximum
mesh size fixed at hm - 0.1818. For a given mesh, a decrease in K with
increasing D e is found. The results indicate a good mesh convergence with
0 < D e _< 1. At higher values of D e , the convergence is slower. More
iteration is needed for higher Deborah numbers. For Newtonian flow 350
iterations are needed for a convergent result, the number of iterations in-
creases to 670 at D e - 0.914 and 800 at D e - 4.0225. Figure 3 shows a
comparison of the FVM results (symbols) and the F E M results of Dhahir
and Walters [40] (lines, taken from Fig. 19 of their paper).
350

6 .... ,..-d:".'-., .... 6 .... ,..~:':..', .... 6, . . . . ," ~:''''', ....

' F
4- - 4[- '- 4- -

2 2

+ . ~

o ' :~.~
a
o,.-~-
~. Iz~.z-a o ~a

-2 - 2 .~--- . q -
9 , b

,I, _ -,JI I. I
. . . . . . . . . I"" -6 '" " " " . . . . . . . . . . .
-6_ -1 o 12 ~ 1 2 6-2 lo 12
Figure 4: A comparison between the computed (FVM: solid lines, FEM:
triangles and squares) and experiment (circle) results by H. Baaijens [42]
along several cross sections for P T T fluid with A - 0.0431 s, p - 0.8
g/cm 3 and e - 0.39. Left: axial velocity, middle: normal stress difference
and right: shear stress. The shear rate at entry is so chosen that
D e - 2.31.

We have chosen the same geometry, with L = 5/3R, and 2113 control
volumes were used over the full flow domain. Using the UCM model, the
force per unit length on the cylinder is computed at varying flow rates
and two values of A (0.0 and 0.06 s). Inertia effects are neglected by
assuming creeping flow. Figure 3 displays the variation of normalised force
components F* - F = / % (cm2/s) with the flow rate. The Deborah number
at the highest flow rate of 80 cm2/s is 0.64. It is seen that F= increases
with the flow rate in both Newtonian (A - 0) and UCM (A - 0.06 s) cases.
Increasing elasticity reduces the force on the cylinder. The direction of
the force is along the flow direction, as it must. The drag force in both
the Newtonian and UCM cases are in excellent agreement with the F E M
results of Dhahir and Walters [40].
The flow of a 5% P I B / C 1 4 solution past a symmetrically confined cylin-
der between two plates of a distance 2L apart is simulated. Following [41],
the P T T model is chosen. The density of the liquid is taken into account
and fixed at 0.8 g / c m 3, the relaxation time is fixed at 0.04313 s and P T T
351

model parameter is fixed at e = 0.39 and ~ = 0 (these values were given in


[41]). The computation is done on half of the flow domain and with mesh
M4 (with a total nodes of 3434, closed to the total nodes of 3981 in [41]).
Figure 4 shows a comparison between the measured and computed (fi-
nite volume) results at cross sections x = - 5 R , x = - 2 R , x = - 1 . 5 R ,
x = 1.5R and x = 2R at the highest measured Deborah number of 2.31,
where the velocity has been normalised with the mean velocity U, and
stresses with 7o - 3~oU/R. To provide a visual comparison with the mea-
sured data, the computed results in the half domain are either symmetri-
cally or asymmetrically mapped to the other half of the flow domain. Due
to experimental errors, the measured results are not symmetric about the
centreline. Figure 4 shows a good agreement for the velocities at all cross
sections. The measured and computed first normal stresses agree well at
all cross sections except at x = - 2 R , especially near the walls, although a
reasonable agreement is still found inside the flow domain. The computed
and measured shear stresses also agree reasonably well in all cross sections
except at x = - 5 R . At x = - 5 R and near the entry, the flow is still fully
developed, the shear stress is linear with respect to y / R , and tends to zero
at centre line. The computed and measured shear stress agree better in
the left domain than in the right domain.
The finite volume solutions in the previous subsection are also checked
against F E M results [41], plotted in Fig. 4 as triangles for a l-mode and
squares for a 4-mode P T T model (the F E M data were taken from Fig. 4.14
and Fig. 4.16 from [41]). In Fig. 4, F E M results are plotted at the cross
sections Ix/RI - 1.5 for the first normal difference (middle figure) and
shear stress (right figure) where significant differences are found between
measured and computed results. The shear stress profiles are displayed on
the right of Fig. 4. An overall agreement with F E M results is noted.
In summary, the unstructured finite volume results are mesh convergent,
comparable to F E M results, and agree well with available experimental
data for the flow past a cylinder in a channel.

5. T H R E E - D I M E N S I O N A L ENTRY FLOW

In the study of the entry flow problem, two types of the cross-sections,
circular and rectangular, are of particular interest. The corresponding flows
are referred to as axi-symmetric contraction and planar contraction flows,
respectively. A large number of experimental works and two-dimensional
numerical simulations on the flow behavior of various types of viscoelastic
fluids in both types of geometry have been performed. Most earlier work
has been reviewed in the literature (see, for example [47-49]). Here, we
352

only briefly list some of the main observed phenomena.


Z

Figure 5: Three-dimensional planar contraction geometry and graded


meshes (one quarter of the domain).

l)JJtll | '0[ ///


JJli
.,

3 4 1 2 3 4 3
(a) w =0.82 (b) W = 1.29 (c) w.-- 2.75

Figure 6: Streamlines for the Boger fluids fitted using the UCM model
with different elasticity under the same flow condition (Re - 2.1), (a)
W~ - 0.82; (b) We - 1.92; (a) We - 2.75. The aspect ratio is one.
In the axi-symmetric contraction flow for Newtonian fluids, it has been
clearly shown experimentally and computationally [50-51] that the pres-
ence (or absence) of the upstream corner vortex is determined by fluid
inertia only.
353

0.05 IIII II t i t
w. -- 3.40E-3 1/
0.00
- - ~ -~~;'[2 / / //
-0.05

- - - .----'m -

2.7 2)I 2.9 1.0

(a)

_ ItfJJ,{tf /I
0.05 Wmi. -8.73E-3 lilltfltt/t
li i/ /
0.00

-0.05 - ' - - ~ - - "~-"~~'~ ~"'- ~".,'1 / / /! / I


--~.____._ -2_~ 7 - - ~ - ~ / ! / /
, ~ . - I ~ - - ~ i i-.~ ~ . ii 9 .,i

2.6 2.7 1.0

(b)

0.05
Wm.~n--- 2 . 1 4 E - 2
0.00 ........

, --..----
-'--'7"--

2.6 2.7 218 ' - 2.9 3.0

(c)

Figure 7: A close-up view of the meshes and flow fields in terms of


streamlines near the re-entrant corner of Fig. 6.

However, for viscoelastic flow, experimental work has demonstrated many


unusual flow phenomena which distinguish it from the corresponding New-
tonian counterpart, such as the vortex enhancement [52], the appearance
of the lip vortex [53], and the flow transition from the steady-state to
three-dimensional, time-periodic and aperiodic flow states as the flow rate
(or the Weissenberg or the Deborah number) increases to some level [54].
The dependence of these flow phenomena on the fluid properties is highly
non-linear, and it seems that fluid elasticity based on steady and dynamic
354

shear properties alone is not adequate to explain the different behavior for
different viscoelastic fluids. It was argued that the extensional behavior
of the fluid plays a more important role on the vortex activities for vis-
coelastic fluids: strong vortex enhancement is expected to occur for fluids
with strain-rate thickening extensional viscosity; on the other hand, strain-
rate thinning extensional viscosity will reduce the strength of the vortex
activity [55-56].
In contrast to the axi-symmetric contraction flow, where a strong vor-
tex enhancement is expected for some Boger fluids, the vortex activity is
weak in planar contraction flows [57-59]; only under some extreme flow
conditions [53] does a lip vortex appear. While for some shear thinning
viscoelastic fluids, such as polyacrylamide aqueous (PAA) solutions, vortex
growth has been observed in both cases. More remarkably, at very high
flow rates, a non-trivial 3D unstable and alternate flow pattern is observed

40 - . , - . , - - , . . . . , - - 140 . . . .
-= we = 0.4 r We = 0.4

35 . - - - o ~ We=0.83 120 ~ we = o.83


+ We = 1.39 *~gzz~tb ~ ; We = 1.39

30 We = 1.88 We = 1.88

X We=2.9 (~=0) 100 v - We=2.9 (~=0)


r~
"~ 2s . ~ We=4.15
We= 2.9(~ =O.1 ~
-----o----- We= 4.15
W e = 2 . 9 ( P, = 0.1
,~, 80
20 Z-- 60 ~ We=2.9(Exp.)

15

4O

20
o ..............

0 ...... -.----~----,-+ ~---; . : . i . . . . . . . . . . . ' " . .._';;;;;;,_.~l:J


-' 2 -9 -6 -3 0 3 6 1 2 -9 . . .-6
.. -3 0 3 6
Z Z
(a) (b)
Figure 8" Computed distributions of (a) axial velocity w; (b) the first
normal stress difference N1 along the centreline.

Some of the selected results of the three-dimensional simulation of the


entry flow [30] are given here (see Fig. 5 for the flow geometry), to illustrate
the performance of the finite volume method. First, the 3D flow patterns
of the Boger fluid fitted using the UCM model are simulated under the ex-
perimental flow conditions used by Walters and Webster [57] and Walters
and Rawlinson [58]. The streamlines calculated for the fluids with different
elasticity are given in Fig. 6, at a Reynolds number of 2.1 and an aspect
ratio of W / H - 1. Here, the Reynolds number R~ defined_ as R~ - pUh/rlo,
and the Weissenberg number is defined as We - A U / h , where h and U are
355

the half height and the mean flow rate in the downstream channel, respec-
tively. As we can seen from the figure, there is virtually no corner vortex,
which is in good agreement with experimental observation. However, upon
close-up inspection of the streamlines near the re-entrant corner, as shown
in Fig. 7, there exists a tiny lip vortex which spreads over several meshes
and increases in both size and strength with increasing fluid elasticity. AI-
though the lip vortex feature for Boger fluids in planar abrupt contraction
flows was observed only at very high elasticity (We at order of 100) in the
flow geometry with higher contraction ratio [53], the trend of our simu-
lation results is consistent with the experimental observation. There may
be two plausible reasons for the discrepancies. First, the lip vortex may
be very hard to discern in the flow visualisation due to its being small in
size. Secondly, the UCM model may not have all the relevant physics in
complex flows of Boger fluids.
Recently, Quinzani et al. [61] measured the detailed flow fields of a well-
characterised shear-thinning polymer solution (polyisobutylene dissolved
in tetradecane (PIB/C14)) flowing in a 3.97:1 planar abrupt contraction
with the Laser-Doppler velocimetry (LDV) and flow-induced birefringence
(FIB). Their experimental results have been partly simulated by some 2D
numerical simulations with the fluid modelled by an integral K-BKZ equa-
tion with a spectrum of four relaxation times [62], and by multi-mode PTT,
Giesekus and UCM models [63]. All of the simulations showed a good gen-
eral qualitative agreement with experimental data. But the agreement is
still semi-quantitative, especially for the extensional stress near the entry
section (up to 30 - 40% discrepancy depending on the level of elasticity).
The PIB/C14 shear-thinning solution used in the experiment work has
been well characterised by Quinzani [64] with several non-linear multi-mode
models, and they found that the best quantitative fit to the transient exten-
sional viscosity could be made using the P T T model with p - 0.8 g/cm 3,
/~ 1, ~0 - 1.424 Pa.s, A - 0.06 s, ~ - 0 (or 0.1), ~ - 0.25 (or 0.05). A
-

finite volume simulation has been performed with these parameters using a
graded mesh. A comparison between the computed values for the centreline
velocity (w), and the first normal stress difference (centreline birefringence)
is shown in Fig. 8. Generally, the numerical results agree reasonably with
experimental data; however, the peak values predicted for N1 are 20-40%
higher than the observed values (depending on (). The discrepancies are
believed to be mainly due to the inadequacy of the constitutive model.
356

6. C O N C L U S I O N S

The results of the finite volume calculations presented here are rather
encouraging for computations of complex flow with arbitrary geometry.
The method is very robust, allowing converged solutions to be obtained
at a high Weissenberg number. The results also compare favourably with
those obtained with traditional finite element methods. In addition, the
use of unstructured mesh allows us to handle more complex computational
domain, much the same way as the finite element method. Coupled with
"both-side" diffusion, which is basically an under-relaxation mechanism,
the finite volume method appears to us a valid way of computing stable
solutions at high Weissenberg number.

ACKNOWLEDGEMENTS
This research is supported by then Australian Research Council (ARC).
The calculations were performed using the facility of Sydney Distributed
Computing Laboratory (SyDCom).

REFERENCES
1. M.J. Crochet, A.R. Davies, and K. Walters, Numerical Solution of
Non-Newtonian Flow, Elsevier, Amsterdam, 1984.
2. J.N. Reddy and D.K. Gartling, The Finite Element Method in Heat
Transfer and Fluid Dynamics, CRC Press, Florida, 1994.
3. N. Phan-Thien and S. Kim, Microstructure in Elastic Media: Princi-
ples and Computational Methods, Oxford University Press, New York,
1994.
4. S.V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere
Publishing Corporation, New York, 1980.
5. M.A. Mendelson, P.-W. Yeh, R.A. Brown and R.C. Armstrong, J.
Non-Newtonian Fluid Mech., 10 (1982) 31.
6. J.M. Marchal and M.J. Crochet, J. Non-Newtonian Fluid Mech., 26
(1987) 77.
7. R.C. King, M.R. Apelian, R.C. Armstrong and R.A. Brown, J. Non-
Newtonian Fluid Mech., 29 (1988) 147.
8. D. Rajagopalan, R.C. Armstrong and R.A. Brown, J. Non-Newtonian
Fluid Mech., 36 (1990) 159.
9. J. Sun, N. Phan-Thien and R.I. Tanner, J. Non-Newtonian Fluid
Mech., 65 (1996) 75-91.
357

10. R.R. Huilgol and N. Phan-Thien, Fluid Mechanics of Viscoelasticity:


General Principles, Constitutive Modelling, Analytical and Numerical
Techniques, Elsevier, Amsterdam, 1997.
11. N. Phan-Thien and R.I. Tanner, J. Non-Newtonian Fluid Mech., 2
(1977) 353.
12. R.B. Bird, R.C. Armstrong and O. Hassager, Dynamics of Polymeric
Liquids: Vol i: Fluid Mechanics, John Wiley and Sons, 2nd Edn, New
York, 1987.
13. A.R. Davies, S.J. Lee and M.F. Webster, J. Non-Newtonian Fluid
Mech., 16 (1984) 117.
14. E.B. Bagley, I.M. Cabot and D.C. West, J. Appl. Phys., 29 (1958)
109-110.
15. A.M. Kraynik and W.R. Schowalter, J. Rheol., 25 (1981) 95-114.
16. A.V. Ramamurthy, J. Rheol., 30 (1986) 337-357.
17. D.S. Kalida and M.M. Denn, J. Rheol., 31 (1987) 815-834.
18. F.J. Lim and W.R. Schowalter, J. Rheol., 33 (1989) 1359-1382.
19. J.R.A. Pearson and C.J.S. Petrie, Proc. 4th Int. Cong. Rheol., Part
3, (1965)265-282.
20. S.G. Hatzikiriakos and N. Kalogerakis, Rheol. Acta, 33 (1994) 38-47.
21. O.D. Kellogg, Foundations of Potential Theory, Dover, New York,
1953.
22. M. Renardy, ZAMM, 65 (1985) 449-451; J. Non-Newt. Fluid Mech.,
36 (1990)419-425.
23. A.J. Chorin, J. Comput. Phys., 2 (1967) 12-26.
24. F.R. Phelan, Jr., M.F. Malone and H.H. Winter, J. Non-Newt. Fluid
Mech., 32 (1989) 197-224.
25. H. Jin, N. Phan-Thien and R.I. Tanner, Comp. Mech., 13 (1994) 443-
457.
26. J.P. van Doormaal and G.D. Raithby, Num. Heat Transfer, 7 (1984)
147-163.
27. R.I. Issa, J. Comput. Phys., 62 (1985) 40-65.
28. J.Y. Yoo and Y. Na, J. Non-Newt. Fluid Mech., 39 (1991) 89-106.
29. Y. Na and J.Y. Yoo, Comp. Mech., 8 (1991) 43-55.
30. S.-C. Xue, N.Phan-Thien and R.I.Tanner, J. Non-Newt. Fluid Mech.,
in press (1997).
31. H.H. Hu and D.D. Joseph, J. Non-Newt. Fluid Mech., 37 (1990) 347-
377.
32. X.-F. Huang, N. Phan-Thien and R.I. Tanner, J. non-Newt. Fluid
Mech., 64 (1996) 71-92.
358

33. B. Gervang and P.S. Larsen, J. Non-Newt. Fluid Mech., 39 (1991)


217-237.
34. S.-C. Xue, N.Phan-Thien and R.I.Tanner, J. Non-Newt. Mech., 59
(1995) 191-213.
35. G.P. Sasmal, J. Non-Newtonian Fluid Mech., 56 (1995) 15.
36. X.-F. Huang, N. Phan-Thien and R.I. Tanner, J. Non-Newt. Fluid
Mech., in press (1997).
37. C. Prakash and S.V. P a t a ~ , Numer. Heat Trans., 8 (1985) 259-280.
38. C. Masson, H.J. Saabas and B.R.Baliga, Inter. J. Numer. Meth. Flu-
ids, 18 (1994) 1-26.
39. L. Davidson, Inter. J. Numer. Meth. Fluids, 22 (1996) 265-281.
40. S.A. Dhahir and K. Walters, J. Rheol., 33 (1989) 781-804.
41. H.P.W. Baaijens, G.W.M. Peters, F.P.T. Baaijens and H.E.H. Meijer,
J. Rheol., 39 (1995) 1243-1277.
42. H.P.W. Baaijens, Evaluation of Constitutive Equations for Polymer
Melts and Solutions in Complex Flows, PhD Thesis, Eindhoven Uni-
versity of Technology, Netherlands (1994).
43. G. Barakos and E. Mitsoulis, J. Rheol., 39 (1995) 1279-1292.
44. P.Y. Huang and J. Feng, J. Non-Newt. Fluid Mech., 60 (1995) 179-
198.
45. A.W. Liu, D.E. Bornside, R.C. Armstrong and R.A. Brown, J. Non-
Newt. Fluid Mech., in press (1997).
46. J. Happel and H. Brenner, Low Reynolds Number Hydrodynamics.
Kluwer, Boston (1964).
47. S.A. White, A.D. Gotsis and D.G. Baird, J. Non-Newtonian Fluid
Mech., 24 (1987) 121.
48. D.V. Boger, Ann. Rev. Fluid Mech., 19 (1987) 157.
49. B. Tremblay, J. Non-Newtonian Fluid Mech., 43 (1992) 1.
50. M. Viriyayuthakorn and B. Caswell, J. Non-Newtonian Fluid Mech.,
6 (1980) 245.
51. M.E. Kim, R.A. Brown and R.C. Armstrong, J. Non-Newtonian Fluid
Mech., 13 (1983) 341.
52. P.J. Cable and D.V. Boger, AIChE J., 24 (1978) 869.
53. R.E. Evans and K. Walters, J. Non-Newtonian Fluid Mech., 20 (1986)
11.
54. G.H. McKinley, W.P. Raiford, R.A. Brown and R.C. Armstrong, J.
Fluid Mech., 223 (1991) 411.
55. F.N. Cogswell, Polym. Eng. Sci., 12 (1972) 64.
359

56. J.L. White and A. Konodo, J. Non-Newtonian Fluid Mech., 3 (1977)


41.
57. K. Walters and M.F. Webster, Philos. Trans. R. Soc. London, A308
(1982) 199.
58. K. Walters and D.M. Rawlinson, Rheol. Acta, 21 (1982) 547.
59. D.M. Binding and K. Walters, J. Non-Newtonian Fluid Mech., 30
(1988) 233.
60. K. Chiba, T. Sakatani and K. Nakamura, J. Non-Newtonian Fluid
Mech., 36 (1990) 193.
61. L.M. Quinzani, R.C. Armstrong and R.A. Brown, J. Non-Newtonian
Fluid Mech., 223 (1994) 1.
62. E. Mitsoulis, J. Rheol., 37 (1993) 1029.
63. F.P.T. Baaijens, J. Non-Newtonian Fluid Mech., 48 (1993) 147.
64. L.M. Quinzani, R.C. Armstrong and R.A. Brown, J. Rheol., 39 (1995)
1201.
361

S E G R E G A T E D F O R M U L A T I O N S IN C O M P U T A T I O N A L
ASPECTS OF C O M P L E X V I S C O E L A S T I C F L O W S

Jung Yul Yoo

Department of Mechanical Engineering


College of Engineering
Seoul National University
Seoul 151-742, Korea

1. INTRODUCTION

The past quarter century has seen a significant advancement in numerical


simulation of complex viscoelastic flows. In the early years, the breakdown of
numerical algorithms even for modest values of elasticity parameter (such as
Weissenberg number) which was observed regardless of the choice of constitu-
tive equation, descretization method and iteration method had been a long-
standing problem in the study of viscoelastic flows. Joseph and co-workers [1-
3] suggested that the problem might be associated with, among others, hyper-
bolicity and change of type in the governing equations, by showing that cou-
pling the constitutive equations of hyperbolic type with the momentum equa-
tions of elliptic type for incompressible flow led to a system of mixed charac-
ter. In this regard, Song and Yoo [4] argued that a type dependent upwinding
scheme like the one used in transonic flow calculations might be useful, and
performed a FDM simulation of the flow of an upper convected Maxwell fluid
through a planar 4:1 contraction using a type dependent upwinding scheme for
the vorticity equation. This resulted in an enhancement of the Weissenberg
number limit to a higher value than ever, but did not quite resolve the high
Weissenberg number problem. More comprehensive discussions on hyperbol-
icity and change of type in steady viscoelastic flows can be found in the book
of Joseph [5].
362

In the last decade, several coupled finite element methods have been devel-
oped for the solution of viscoelastic flow problems. They include the sub-ele-
ment method of Marchal and Crochet [6], the explicitly elliptic momentum
equation (EEME) method of King et al. [7], and the elastic-viscous split-stress
(EVSS) method of Rajagopalan et al. [8]. These techniques enabled the re-
searchers to develop f'mite element methods that were stable enough to obtain
converged solutions at Weissenberg numbers as high as possible. It seems that
the success of all these strategies resulted from applications of streamline up-
winding technique to the constitutive equations of hyperbolic type. However, it
is noted that these finite element methods were based on fully coupled algo-
rithms, in which the velocity, pressure and stress were solved simultaneously.
In addition, they were restricted to very simple geometries requiting modest
number of elements even with supercomputers. For example, in one of the cal-
culations reported in [6], the degrees of freedom for stresses were 14116 with
a mesh containing 210 elements.
In the meantime, another group of researchers have studied viscoelastic
flows in terms of decoupled strategies, in which the velocity, pressure and
stress are solved separately. In decoupled algorithms, the momentum equation
is solved separately from the constitutive equation, which is significantly less
expensive than their coupled counterparts. Decoupled technique is also a natu-
ral choice for integral viscoelastic models since it is not a simple task to devise
coupled algorithms for them [9]. Furthermore, a properly constructed decou-
pled scheme should in general have a larger radius of convergence than a cou-
pled scheme which often relies on Newton type iterations [10]. The numerical
accuracy and computer efficiency of a decoupled algorithm depends on three
aspects: (i) solution of the constitutive equation subject to a given velocity
field; (ii) solution of the Navier-Stokes equation subject to a given viscoelastic
stress field; (iii) the way how the two equations are decoupled.
There has been a traditional difficulty associated with solving the Navier-
Stokes equation for an incompressible flow, that is, the incompressibility con-
straint, or the coupling of the pressure and velocity. For an incompressible
flow, the continuity equation in itself does not have an explicit reference to the
pressure. Therefore, a pressure equation, which enables the segregation of the
solution procedure for the pressure from that for the velocity, must be devised
by some further manipulation of the governing equations. A very well-known
way of devising such an equation is the SIMPLE (semi-implicit method for
363

pressure linked equations) algorithm [11] used in the finite volume methods.
Another approach calls for a 'pressure correction', in which an initially
guessed pressure is corrected successively at each iteration by adding the pres-
sure correction obtained from the associated equation. This pressure correction
is also common to artificial compressibility [12] and augmented Lagrangian
[13, 14] techniques. In the framework of the FEM, this difficulty can be easily
overcome by using integrated formulations in which the velocity and the pres-
sure are obtained simultaneously. However, the global matrices become in-
evitably larger than those of the segregated formulations. Therefore, in segre-
gated formulations, one can save more memory and CPU time since the pres-
sure is obtained separately from the momentum equation.
Some time ago, Boger [15] pointed out that, experimentally, flows in two-
dimensional geometries tend to become time-dependent and three-dimensional
when viscoelasticity is important [16]. As was also discussed in the work of
Evans and Waiters [17], concerning the lip-vortex mechanism of vortex en-
hancement in planar contraction flows, it is clear that any numerical simulation
aimed at predicting the vortex growth in such flows should also be extended to
three space dimensions and time dependent behaviour. These features of the
flow have to be taken into account in the numerical simulation of such flows
[18]. Recent researches pay more and more attention to computationally effi-
cient algorithms, since the flows of recent interest are ultimately three-dimen-
sional and time dependent [16]. Decoupled and segregated methods, which
deemed insufficiently accurate when they first appeared, have been developed
so that they now can produce credible results. Furthermore they remain very
attractive because of significantly reduced core memory and CPU time re-
quirement, which are crucial in transient and three-dimensional problems.
The objectives of the present article are to examine the segregated formula-
tions currently adopted in numerical simulations of viscoelastic flows and to
discuss the possibilities of emerging algorithms. To avoid possible confusions
in the usage of terminologies in this article, the term 'decoupled' is to be used
for the decoupling of the extra stress into the viscoelastic component and the
optional Newtonian component as they appear in the momentum equations
(Keunings [9]), and the term 'segregatea~ is now used for the segregation of the
pressure from the momentum equations. In section 2, governing equations for
viscoelastic flows in conjunction with basic decoupled techniques for the extra-
stress are presented. In section 3, several strategies handling the hyperbolic
364

nature of the constitutive equations are discussed. The segregated formulations


of the Navier-Stokes equation are treated in section 4. The possibilities of
emerging segregated algorithms are considered briefly in the conclusion.

2. G O V E R N I N G EQUATIONS FOR V I S C O E L A S T I C F L O W S

For an incompressible viscoelastic flow, the governing equations are the


continuity, momentum and constitutive equations. Since the main concern of
the present article is to take a general view of segregated algorithms in numer-
ical simulation of viscoelastic flows, we do not intend to elaborate on the vis-
coelastic modelling. In this regard, we will simply adopt the 'upper-convected
Maxwell model', which is one of the most widely studied differential models
for viscoelastic flows. Then we are to consider the following set of equations:

V.u=O, (1)
~gu
(2)

v
r+A,~=2r/d, (3)

where u is the velocity, p is the pressure, d is the rate of strain, lr is the extra
stress, p is the density,/7 and A, are the viscosity and relaxation time, respec-
tively, and v denotes the upper-convected derivative def'med by

v 9
lr = - ~ + u . V l r - V u r . l r - lr. V u . (4)

When equations (3) and (4) are taken in conjunction with the momentum
equation (2) and the continuity equation (1), it is apparent that the extra-stress
components have to be calculated together with the velocity components and
the pressure. The concept of the decoupled method is based on splitting the
extra-stress into the Newtonian part and the viscoelastic part as

r=2rld+ (5)
365

By substituting equation (5) into equations (2) and (3), the momentum and
constitutive equations are now written as

81/
P-~ (6)

v v
I'E + A,l-e = - 2 ~ r/d . (7)

The above alternative formulation was first introduced by Perera and Wai-
ters [19-20], Mendelson et al. [21] and Crochet and Keunings [22]. This was
further modified by Rajagopalan et al. [8] to construct the elastic-viscous split
stress (EVSS) formulation, which is categorized as a coupled method. In this
coupled formulation of the EVSS, the convected derivative of d in the
Galerkin discrete form of equation (7) involves second-order spatial deriva-
tives of the interpolating function ~, which can be eliminated by performing
v
an integration by parts on the term O d, with the resulting boundary integral
containing the flux of d through the boundary. In a decoupled formulation,
however, this complication may be avoided by first calculating r/d at node
points from the known velocity field and then consistently interpolating r/d (as
part of the extra stress) in the same manner as 1:~ [10].
Now, the momentum equation turns out to be the Navier-Stokes equation
with a pseudo-body-force V. ~:~,so that the solution always reduces to that of a
Newtonian flow when A, vanishes. The velocity and the pressure can be ob-
tained by solving equation (6) with the continuity equation in the given stress
field, and the stress can be obtained by solving equation (7) in the given veloc-
ity field, iteratively. The main disadvantage of decoupled techniques lies in the
iterative procedure because Picard schemes usually used in decoupled methods
are often slow in convergence, which is not even guaranteed no matter how
closely the initial estimates are chosen to the solution. However the decoupled
technique generally requires less core memory than a coupled method, which
turns out to be more efficient for larger computations.
Another attractive feature of decoupled methods is the breakup of the
problem into the solution of an elliptic Newtonian-like flow with pseudo-body-
force, and the integration of constitutive equations using fixed flow kinematics.
366

Thus classical methods can be used to discretize the Navier-Stokes equation,


while special techniques for the extra-stress computation may be developed to
take into account the special mathematical nature of the constitutive equation
involved [9].

3. NUMERICAL TREATMENT OF CONSTITUTIVE EQUATIONS


IN DECOUPLED METHODS

3.1 Background
The main difficulty in dealing with the constitutive equation (7) is that its
type is hyperbolic. Even though this advection equation is linear in the stress
with its structure being much simpler than the Navier-Stokes equation, it is still
a challenge to computation because the advection equation has no built-in
mechanism to smooth discontinuities in directions normal to the flow [23]. This
problem is similar to that encountered in high Reynolds number flows of
Newtonian fluids where high velocity along streamlines generate wild oscilla-
tions in the numerical solution. Before the advancement of the upwinding
technique, the only way to eliminate these oscillations seemed to be to refine
the mesh, such that convection was made not to be dominant on an element
level. However, Keunings [24] showed in a study on the high Weissenberg
number problem that the use of a very f'me mesh rather reduced significantly
the range of the Weissenberg number for which the viscoelastic solutions could
be obtained. It is because the local Peclet number, which represents the ratio of
the convection to the diffi~ion, remains to be infinite from equation (7), no
matter how we ref'me the mesh. Furthermore, it is reminded that one must cal-
culate the stress closer to the comer singularity with a f'mer mesh in the case of
the problem involving corner singularity, such as the 4:1 contraction flow
problem.

3.2 SUPG (Streamline Upwind/Petrov-Galerkin) Method


There have been several methods to overcome this difficulty, regardless of
the coupled or decoupled method. In a finite element framework, the most
popular approach for this problem is the Streamline Upwind Petrov-Galerkin
(SUPG) method [25]. Some earlier upwind formulations for multi-dimensional
problems often exhibited numerically false diffusion in the direction perpen-
367

dicular to the flow direction. The basic idea of the streamline upwind method
is to add diffusion which acts only in the flow direction. Although the SUPG
method has been highly successful in solving the Navier-Stokes equations for
Newtonian flows, it was never introduced into solving the constitutive equa-
tions of the viscoelastic flows, until the first attempt was made by Marchal and
Crochet [6] who adapted the SUPG method for their coupled method.
Following Brooks and Hughes [25] and Marchal and Crochet [6], we define
modified weighting function q~ as

. _ luAl+ ,eh
9 =T+ku~-v~P, k = 2 (8)

where q~ is the Galerkin weighting function, u~ and Uy are the components of


the velocity vector u h at the center of an element, h~ and hy are the character-
istic lengths of the element. The normalization of the velocity field is intro-
duced so that the weighting function remains to be O(1) in regions even where
the velocity field vanishes. Note that in the coupled formulation of Marchal
and Crochet [6], the form of k in equation (8) is not used because it is not dif-
ferentiable at u~ = U y - 0 , and one cannot calculate the Jacobian matrix with
Newton type iterations. Instead, they adopted the following form:

%= U hx)
(9)
2

Applying the modified weighting function to all terms in the constitutive


equation, we end up with the consistent streamline upwind method:

q~+]T~-~-V~P ~ + , ~ 2~, = .

An alternative method for the solution of the constitutive equation is the in-
consistent streamline upwind method, in which the modified weighting func-
tion is applied solely to the advection term u. V~g of the constitutive equation:
368

~n 9 ~:e+~:E+2,q,r/d + .V~'E d O = 0 . (11)

Comparing (11) with the conventional Galerkin weighting form, it can be


considered as if an extra term is artificially added to the Galerkin formulation.
This method is equivalent to the artificial diffusion method commonly used for
hyperbolic equations in the FDM or FVM. Note that the artificial diffusion
method is second-order accurate in space, while conventional upwinding
method is first-order accurate.

3.3 DG (Discontinuous Galerkin) Method


It is well known that a cure for the numerical oscillations by introducing a
certain amount of artificial diffusion into the usual upwinding technique only
results in a loss of accuracy in the numerical solution. Discontinuous Galerkin
(DG) method of Fortin and Fortin [26] was developed for a decoupled finite
element approach to avoid smooth and stable but physically unrealistic solu-
tions. This technique first introduced by Lesaint and Raviart [27] for the neu-
tron transport equation was successfully applied to the numerical simulation of
viscoelastic flows at high Weissenberg numbers. However, at first it was con-
sidered to be impossible to solve the steady flow problem by using the Picard
iteration scheme. Therefore a time stepping scheme had to be employed, pos-
sibly because steady-state Picard's scheme does not provide information on the
qualitative behaviour of the numerical solutions as Keunings [9] pointed out.
Later, Fortin and Fortin [28] showed that the steady flow problem could be
solved by applying the Generalized Minimal Residual (GMRES) algorithm of
Saad and Schultz [29]. The value of the Weissenberg number thus obtainable
was limited again and the extra-stress field showed an oscillatory behaviour.
It was subsequently shown by Basombrio et al. [30] that the use of a linear
interpolation for the extra stress tensor, rather than a quadratic as in Fortin
and Fortin [26], significantly reduce the oscillations. However, it is quite un-
fortunate that for the DG method, the stress variable cannot be eliminated on
the element level as, for instance, in the work of Baaijens [31] and that oscilla-
tion free solutions are only obtained at limited values of the Weissenberg nmn-
ber. Baaijens [31] and Baaijens [32] combined the DG method with the Opera-
tor Splitting (OS) method [23], which was originally designed to decouple the
369

advection treatment and the rest of the procedure for a time-dependent prob-
lem. A shared feature of the OS and DG method is the use of discontinuous
interpolations of the extra stress tensor. This allows the elimination of the ex-
tra stress variables on the element level, yielding an efficient algorithm for
multi-mode fluids. More recently, Fortin et al. [33] and Gurnette and Fortin
[34] have combined the DG method with the EVSS method, by which they
could obtain good convergence and oscillation-free solutions for higher Weis-
senberg numbers.

3.4 Method of Characteristics


Another method for handling the hyperbolic type equation in the decoupled
finite element approach is the method of characteristics, which was first intro-
duced in the numerical simulation of viscoelastic flows by Fortin and Esselaoui
[35]. Hadj and Tanguy [36] also employed the weak formulation of the method
of characteristics to compute the full convected derivative of the stress tensor
and obtained solutions at a higher Weissenberg number with an Oldroyd-B
fluid than that attained by Fortin and Esselaoui with a Maxwell fluid. In a re-
cent work of Basombrio [37], the hyperbolic equations for the advective trans-
port of stresses were integrated directly at each node using nonconservative
strong formulation of the method of characteristics [38]. This method was also
combined with the EVSS method by Kabanemi et al. [39] in which the hyper-
bolic nature of the constitutive equation was coped with using a weak formula-
tion.

3.5 Other Methods in FDM and FVM


In the finite difference or finite volume framework, the applicable methods
for the hyperbolic type equations are the upwinding techniques and the artifi-
cial diffusion techniques. Hu and Joseph [40] used a conventional, first order
upwind scheme for the constitutive equations. Choi et al. [41 ] and Yoo and Na
[42] used the deferred correction method, which is an upwind corrected
scheme involving an artificial diffusion term to attain second-order accuracy
and unconditional stability, while Xue et al. [43], and Huang et al. [44] used an
artificial diffusion method only. Sato and Richardson [45] used a finite volume
approach for the constitutive equation, while a finite element approach was
used for the momentum equation. They introduced the FCT (flux-corrected
transport) concept into solving the constitutive equations, since numerical
370

schemes for advection should be TVD (total variation diminishing) in order to


obtain stable results. Recently, Mompean and Deville [18] applied, so-called,
QUICK (Quadratic Upstream Interpolation Scheme for the Convective Kine-
matics) to the constitutive equation, which was proposed by Leonard [46]. For
a uniform grid system this scheme is of third-order accuracy, and for a non-
uniform grid system it is of second-order accuracy [46].
The staggered grid system has been widely used in the FDM or FVM ap-
proaches, where the pressure and the stresses are located in the center of the
control volume and the velocities are located on the faces of the control vol-
ume as shown in Figure 1. The staggered grid is first designed to eliminate the
checkerboard pattern in the pressure field for the Newtonian flows of the FDM
or FVM, which is quite similar to the mixed f'mite element methods satisfying
the Babuska-Brezzi compatibility condition [47, 48] on the spaces for the ve-
locity and pressure. This formulation of the viscoelastic flow calculation offers
a major advantage in avoiding the stress calculations, for example, at the re-en-
trant corner of the 4:1 contraction flow problem, while the method correctly
preserves the flow physics in its neighboring control volumes.

I u

I p,l: I
L . . . . _1

Figure 1 Staggered grid in the computational domain

In recent works of Saramito [49, 50], the original finite volume element
(Figure 2) was proposed for the mixed FEM which was somewhat similar to
the staggered grid in the FVM. It is noticeable that the staggered formulation is
relatively difficult to discretize the governing equations and it nearly cannot be
used for an unstructured grid mostly because one cannot define a control vol-
ume consistently for arbitrarily shaped meshes. Therefore the non-staggered
371

grid system in which all dependent variables are located at the grid points be-
comes more and more popular in the Newtonian flow calculations. In the un-
structured FVM approach of Huang et al. [44], they used a non-staggered for-
mulation, which corresponds to equal-order element in the FEM.

19 0 E
[-] ~2

) 9 () Ou

0 "r "t'22,p
l
/ 0 E

Figure 2 Modified finite element

4. SEGREGATED FORMULATIONS IN THE NAVIER-STOKES


EQUATION

4.1 Pressure Equation


One of the complications encountered in solving the Navier-Stokes equation
is that it includes the pressure gradient term, whereas there is no independent
equation for the pressure. Furthermore, the continuity equation does not have
a dominant variable in it. Mass conservation is a kinematic constraint on the
velocity field rather than a dynamic one. In the FEM approach, this difficulty
is simply avoided by using an integrated formulation in which the momentum
and continuity equations are solved simultaneously. However, this formulation
requires a large memory and computing time. An alternative way to overcome
this difficulty is to construct a pressure equation in order that the pressure
obtained from it intrinsically satisfies the continuity equation. In this seg-
regated velocity-pressure formulation of the Navier-Stokes equation, velocities
and the corresponding pressure field are computed alternately in an iterative
sequence, in contrast to the integrated formulation. The most attractive aspect
of the decoupled approach in the numerical analysis of viscoelastic flows is that
it works very well with the segregated strategy for the Navier-Stokes equation
which requires a much smaller memory and computing time.
372

This segregated formulation has been one of the most commonly used ap-
proaches in the finite difference procedure [51], since it does not solve the mo-
mentum and continuity equations simultaneously as in the integrated formula-
tion of the FEM. For the purpose of solving them simultaneously, Song and
Yoo [4], Choi et al. [41] and Sasmal [52] applied the stream function-vorticity
formulation to viscoelastic flow problems. However this approach cannot be
extended to three dimensional flows, and it needs assumptions on stream func-
tion and vorticity in treating the boundary conditions.
Derivation of a pressure equation is based on combining the momentum and
continuity equations. A pressure equation of Poisson type can be obtained by
taking the divergence of the momentum equation (2):

= v. +v. . (]2)

Sato and Richardson [45] used this formulation for their viscoelastic flow cal-
culation, in which MAC (Marker And Cell) method of Harlow and Welch [53]
was incorporated into a finite element solution. A similar procedure to this
formulation is also used in the three-dimensional FVM calculation of Mompean
and Deville [ 18].

4.2 SIMPLE (Semi-Implicit Method for Pressure Linked Equa-


tions)-like Algorithm
The well-known approach in the FVM for the pressure equation is the SIM-
PLE algorithm of Carretto et al. [54], which can be simply described as fol-
lows:
(i) Predictor stage
In this stage a guessed pressure field is used, usually denoted by p*, and by
solving the momentum equation the velocity field u* is obtained.
(ii) Corrector stage
To obtain the velocity and pressure fields which satisfy both the continuity
and momentum equations, the corrections u' and p' are obtained in this stage,
and then u = + u p = p + p' This stage is repeated by treating the cor-
rected pressure p and velocity u as a newly guessed pressure p* and velocity
u* until a converged solution is obtained.
373

However, the SIMPLE algorithm does not converge rapidly because the ve-
locity correction is obtained improperly. Its performance depends greatly on
the size of the time step, o r - for steady flows - on the value of the under-re-
laxation parameter used in the momentum equations. It was found by trial and
error that convergence characteristics can be improved by controlling the un-
der-relaxation. Thus some modifications to this conventional SIMPLE algo-
rithm were proposed. They are SIMPLER (SIMPLE Revised) of Patankar
[11], PISO (Pressure-Implicit with Splitting Operators) of Issa [55] and SIM-
PLEC (SIMPLE Consistent) [56] which do not need under-relaxation of the
pressure correction [57].
Recently, the FVM approaches based on these algorithms were adopted for
viscoelastic flow computations, because of their attractive features of space and
time savings, as well as their numerical stabilities. For steady flow problems,
Hu and Joseph [40], Yoo and Na [42] and Huang et al. [44] used SIMPLER al-
gorithm, and its modification is used in the work and Xue et al. [43], while
Gervang and Larsen [58] used PISO algorithm and Luo [59] used classical
SIMPLE algorithm. Note that in these algorithms the pressure and velocity
fields obtained at the end of each time step (or iteration step for steady flow
problems) do not satisfy one and the same momentum equation so that for
time-dependent computations, sub-iterations are necessary at each time step, or
very small time steps must be chosen, which accordingly degrades the effi-
ciency of these algorithms.

4.3 Augmented Lagrangian Method


In the frame work of the FEM, there is an alternative approach for the seg-
regation of the pressure, that is augmented Lagrangian method [13, 14] which
was derived from the variational formulation of the Stokes flow problem and
turned out to be a powerful iterative method for solving it. Using the varia-
tional formulation of the Stokes equation, one can obtain the symmetric posi-
tive-definite functional equation, which is equivalent to the minimization
problem. For the solution to this minimization problem, there have been many
variants from the following very simple algorithm known as Uzawa algorithm
[13, 14]-

P St - V ' ~ = - V p ' , (13)


374

p"+~ =p" - e V . u " , (14)

where m denotes the present iteration step and e is a step length for the itera-
tive procedure. Hadj and Tanguy [36] and Kabanemi et al. [39] used this classi-
cal Uzawa algorithm for the momentum and continuity equations. Fortin and
Fortin [26] and Gu6nette and Fortin [34] solved the Stokes flow problem by
this algorithm combined with a condensation method [60] that eliminates the
velocity degrees of freedom associated with the center of each element and the
pressure gradients. This is achieved by using the continuity equation at the el-
ement level, thanks to the discontinuous approximations of the pressure. This
method greatly reduces the number of unknowns, resulting in a very efficient
Stokes solver. However, the convergence properties of this algorithm are not
so good because the determination of e is not easy so that e is usually fixed
with the iterative procedure. The way around this is quite simple but requires a
few developments [13, 14, 23]. Fortin and Fortin [28] showed that an intro-
duction of the preconditioned residual for the dependent variables enhance the
convergence, but the computations still had to be performed in double preci-
sion.
In fact, the algorithm (13)-(14) can be interpreted as a gradient algorithm
applied to the minimization of the functional. With this interpretation in mind
it is natural to seek the more effective iterative methods for the minimization
of quadratic functionals, such as the steepest decent method, the minimum
residual method or the conjugate gradient method. In these modified methods,
the conjugate gradient method is especially attractive for solving quadratic
problems because theoretically it converges in a finite number of iterations and
moreover, it leads to quadratic convergence in the general case. These algo-
rithms require the additional presence in memory, but this increased memory
requirement will be justified if the automatic determination of the step length
e leads to a very clear improvement in the speed of convergence compared
with original algorithm (13)-(14) [ 13, 14, 23 ].

4 . 4 0 S (Operating Splitting) Scheme


Clearly, the augmented Lagrangian method is not applicable in itself to the
Navier-Stokes equation containing the non-linear advection term for the veloc-
ity. As mentioned in section 3, the Operator Splitting (OS) scheme has been
used in order to decouple an advection treatment and the rest of the procedure
375

for a time-dependent problem. This method is also called the alternating direc-
tion implicit technique [13, 14, 23, 61], which separates the main two difficul-
ties in the computation of the Navier-Stokes equation: the non-linearities in the
momentum equation and the incompressibility constraint. In the conventional
numerical simulation of viscoelastic flows, the non-linear term is usually ne-
glected. In an attempt to answer some of the unresolved questions in their ear-
lier work, Evans and Walters [17] investigated aqueous solutions of polyacry-
lamide and concluded that a lip-vortex mechanism may be responsible for
vortex enhancement for all planar contraction ratios. These experimental stud-
ies consist of observations and measurements of the phenomena resulting from
the change of the flow rate of the same viscoelastic fluid. Therefore, in order
to simulate such flows numerically, the inertia must be retained in the momen-
tum equations.
Glowinski and Pironneau [23] proposed the 0-scheme based on the above-
mentioned OS method to treat the nonlinearity and the incompressibility of the
Newtonian flows. It consists of three step time marching technique such that
(i) First step

gl n+O _ II n
P Ok _aV..g,+o +Vpn+O =fn+O ..[_~V. ,[,n (/gn .V)/gn , (15)

V.u "+~= 0 , (16)

(ii) Second step

un+l-O _ 11n+O
P (1 - 2 0 ) k - f l Y . Zn+l-O "b (U n+l-O" V ) U n+l-O : fn+l-O + O~V..[n+O _ V p " + ~ ,

(17)

(iii) Third step

U n+l _ ign+l-O
P Ok _ o ~ V . ,[n+l q_Vpn+l = f n + l + ~ V . , [ n + l - O (lln+l-O.v)lln+l-O , (18)

V-/g n+l = 0 , (19)


376

where a = (1- 20)/(1-0)and/3 = 0/(1-0), and 0 can be selected by a numeri-


cal experiment.
One can see that the first and third steps are the Stokes-like problems in
which the nonlinear terms are eliminated by the second step. Thus the Uzawa
type algorithm can be used in the first and third steps and the nonlinear term is
treated in the second step by using the method mentioned in section 3, such as
the upwinding technique. This scheme is of order two in time, and allows one
to also compute steady solutions efficiently. Glowinski and Pironneau [23] in-
troduced a preconditioning stage to the Uzawa type conjugated gradient
method for an acceleration of the convergence in the Newtonian flow compu-
tations. In the case of the Stokes flow computation, the second step can be
skipped since it needs not the advection treatment. However, when the 0-
scheme is combined with the constitutive equations of viscoelastic flows, the
second step can not be skipped because the advection term in the constitutive
equation is always involved.
Recently, Saramito [49, 50] applied this 0-scheme to the calculation of the
viscoelastic flows, but he neglected the inertia in the momentum equations,
while Luo [62] extended it to the case of including the inertia in the momentum
equations. Until now, no one used the 0-scheme for time-dependent flow
problems. As Szady et al. [63] mentioned in their recent work of coupled ap-
proach, the 0-scheme can be efficiently utilized for the three-dimensional
transient flow where it can save much CPU time and guarantee numerical sta-
bility.

4.5 Fractional Step Method


The alternating direction method is closely related to the fractional step
method [14]. Actually the alternating direction or fractional step method has
been extensively used for quite a long time for solving time-dependent partial
differential equations. Concentrating particularly on the Navier-Stokes equa-
tion for incompressible viscous fluids, the first works were performed by
Chorin [64] and Temam [65]. However, the fractional step method for the
Navier-Stokes equation had not been used extensively until successful result of
its application to the three-dimensional unsteady turbulent flow calculation was
reported by Kim and Moin [66], which was later used for the Direct Numerical
Simulation (DNS) [67, 68]. A typical formulation of the fractional step method
is as follows"
377

(i) First step

P At + 3(un'V)un-(un-l"V)un-1)=-2

(ii) Second step

V 2~bn+l = P V.u (21)


At

(iii) Third step

Un+l --fl _v~n+l (22)


P At = '

where r is a pseudo-pressure and V~b#+~- Vp #+~+O(At2). One can see that the
third step is derived from taking the divergence of equation (21) and requiting
u n+~ to satisfy the continuity equation. There are many variants of the frac-

tional step method such as a fully implicit fractional step method [69], due to a
vast choice of approaches to time and space discretization, but they are all
based on the principles described above.
The fact that the pressure is segregated from the velocity in the fractional
step method was utilized in the finite element analysis of the incompressible
Navier-Stokes equation by several researchers [70-74]. Because the fractional
step method does not include any approximation procedure, such as adopted in
the SIMPLE algorithm based finite element methods [75-78], this approach is
more accurate than the SIMPLE algorithm based approach for the same grid.
Furthermore, in the fractional step method, the coefficients of the pressure
equation are fixed and the inverse matrix of the pressure equation does not
need to be calculated at every time step like the SIMPLE algorithm based ap-
proach [79].
Carew et al. [80] and Baloch et al. [81] used this fractional step method
combined with the Taylor-Galerkin method for the advection treatment. The
details of that scheme is described in the work of Hawken et al. [82]. Recently,
Baloch et al. [83] were able to carry out a numerical simulation of a three-di-
mensional viscoelastic 40:3:3 expansion flow, using the Phan-Thien and Tanner
378

(PTT) constitutive fluid model, where they showed the streamlines in the cen-
ter plane of the 3D expansion and made comments on the differences in the
vortex development from the 2D simulation. Recently, Sureshkumar et al. [84]
applied the fractional step method of Kim and Moin [66] to the DNS of the tur-
bulent channel flow of a polymer solution by spectral method.
The major difference between the 0-scheme with preconditioned conjugate
gradient (PCG) method (15)-(19) and the fractional step method (20)-(22) is
how to advance one time step. The two methods are both of order two in time
and the Poisson equations for the pressure are involved respectively. However,
sub-iterations for satisfying the convergence criteria of the momentum and
continuity equations, that is, equations (15)-(16) and equations (18)-(19), are
needed in the PCG method of the 0-scheme, while the Poisson equation is
solved just one time at the second step (21) in the fractional step method. It is
well-known that the Poisson equation for the pressure is the most time consum-
ing one in the segregated formulation of the Navier-Stokes equations. In view
of this, comparing with the fractional step method, the 0-scheme may be inef-
ficient, particularly for three-dimensional problems, despite the fact that the
Poisson equation is solved in the discrete pressure space whose dimension is
much smaller than the dimension of the discrete velocity space in mixed FEM
formulation. Note that the 0-scheme with Uzawa type PCG method cannot
avoid the Babuska-Brezzi condition [13], and thus it must be discretized via the
mixed formulation where the discrete pressure on a grid is defined twice
coarser than the one used to discretize the velocity. On the other hand, the
fractional step method can avoid the compatibility condition as shown in the
recent work of Choi et al. [79] which combined the SUPG with an equal-order
finite element formulation. This work was extended to the numerical analysis
of a Bingham plastic [85]. However, the boundary condition treatments for the
intermediate velocity ti and the pressure in the fractional step method is some-
what unclear and some further assumption is needed, while such an assumption
is not needed in the 0-scheme.

5. CONCLUDING REMARKS

Recently, the streamline upwind (SU) scheme for the control volume finite
element method (SUCV) [86] has been proposed. When we consider that the
379

FEM in conjunction with the SU technique gives satisfactory results in analyz-


ing viscoelastic flows, adopting SUCV into the existing algorithm based on the
FVM seems to be promising. Furthermore, in the last decade the segregated
finite element formulations [70-79, 82] have been steadily developed, which
combine the merits of both the FEM and segregated FVM using SIMPLE-like
or split methods. Therefore, future research efforts implementing these segre-
gated FEM in the numerical simulation of viscoelastic flows are also highly
recommended.
It can be said that the earlier difficulties in the high Weissenberg number
problem have been now overcome by several coupled formulations such as a
sub-division method [6], EEME [7] and EVSS [8] methods, which can obtain
the converged solutions at Weissenberg numbers as high as possible. In the
meantime, another group of researchers using a decoupled and segregated
formulation showed that they can also handle the high Weissenberg number
problems. In particular, some of those segregated methods utilized special
techniques used in coupled methods for hyperbolic type constitutive equations.
The new trend of recent researches of a viscoelastic flow computation is the
analysis on unsteady and three-dimensional flow including the inertia [16] ,
since those flows are related to the most interesting phenomena of non-
Newtonian fluids such as the lip-vortex growth or the turbulent drag reduction
by polymer additives. Therefore fully decoupled, segregated formulations of
viscoelastic flows in the FDM, FVM and FEM seem to be more and more
attractive nowadays due to their capability of handling those problems as well
as their cost-effective properties.

ACKNOWLEDGEMENT
I would like to express my deepest thanks to Mr. Taegee Min for his in-
valuable efforts in preparing this manuscript.

REFERENCES

1. D. D. Joseph, M. Renardy and J. C. Saut, Arch. Ration. Mech. Anal., 87


(3) (1985), 213.
2. J. Y. Yoo, M. Ahrens and D. D. Joseph, J. Fluid Mech., 153 (1985), 203.
380

3. J.Y. Yoo and D. D. Joseph, J. Non-Newtonian Fluid Meeh, 19 (1985), 15.


4. J.H. Song and J. Y. Yoo, J. Non-Newtonian Fluid Mech., 24 (1987), 221.
5. D. D. Joseph, Fluid Dynamics of Viscoelastic Liquids, Springer-Verlag,
New York, 1990.
6. J. M. Marehal and M. J. Crochet, J. Non-Newtonian Fluid Mech., 26
(1987), 77.
7. R. K. King, M. R. Apelian, R. C. Armstrong and R. A. Brown, J. Non-
Newtonian Fluid Mech., 29 (1988), 147.
8. D. Rajagopalan, R. C. Armstrong and R. A. Brown, J. Non-Newtonian
Fluid Mech., 36 (1990), 159.
9. R. Keunings, Simulation of viscoelastic fluid flow, in: C. L. Tucker III
(Ed.), Computer Modeling for Polymer Processing, Hanser Publishers,
Munich, 1989.
10. X.-L. Luo and R. I. Tanner, J. Non-Newtonian Fluid Mech., 31 (1989),
143.
11. S. V. Patankar, Numerical Heat Transfer and Fluid Flow, Hemisphere,
New York, 1980.
12. A. J. Chorin, J. Comput. Phys., 2 (1967), 12.
13. M. Fortin and R. Glowinski, Augmented Lagrangian Methods: Applica-
tions to the Numerical Solution of Boundary-Value Problems, North-Hol-
land, Amsterdam, 1983.
14. R. Glowinski, Numerical Methods for Nonlinear Variational Problems,
Springer-Verlag, New York, 1984.
15. D.V. Boger, Ann. Rev. Fluid Mech., 19 (1987), 157.
16. B. Caswell, J. Non-Newtonian Fluid Mech., 62 (1996), 99.
17. R. E. Evans and K. Waiters, J. Non-Newtonian Fluid Mech., 32 (1989),
95.
18. G. Mompean and M. Deville, J. Non-Newtonian Fluid. Mech., in press
(1997), private communication.
19. M. G. N. Perera and K. Waiters, J. Non-Newtonian Fluid Mech., 2
(1977), 49.
20. M. G. N. Perera and K. Waiters, J. Non-Newtonian Fluid Mech., 2
(1977), 191.
21. M. A. Mendelson, P.-W. Yeh, R. A. Brown and R. C. Armstrong, J. Non-
Newtonian Fluid Mech., 10 (1982), 31.
381

22. M. J. Crochet and R. Keunings, J. Non-Newtonian Fluid Mech., 10


(1982), 339.
23. R. Glowinski and O. Pironneau, Ann. Rev. Fluid Mech., 24 (1992), 167.
24. R. Keunings, J. Non-Newtonian Fluid Mech., 20 (1986), 209.
25. A. N. Brooks and T. J. R. Hughes, Comput. Meth. Appl. Mech. Eng., 32
(1982), 199.
26. M. Fortin and A. Fortin, J. Non-Newtonian Fluid Mech., 32 (1989), 295.
27. P. Lesaint and P. A. Raviart, On a finite element method for solving the
neutron transport equations, in: C. de Boor (Ed.), Mathematical Aspects
of Finite Elements in Partial Differential Equations, Academic Press, New
York, 1974.
28. A. Fortin and M. Fortin, J. Non-Newtonian Fluid Mech., 36 (1990), 277.
29. Y. Saad and M. H. Schultz, SIAM J. Sci. Stat. Comput., 7 (3) (1986), 856.
30. F. G. Basombrio. G. C. Buscaglia and E. A. Dad, J. Non-Newtonian
Fluid Mech., 39 (1991), 189.
31. F. P. T. Baaijens, J. Non-Newtonian Fluid Mech., 48 (1993), 147.
32. F. P. T. Baaijens, J. Non-Newtonian Fluid Mech., 51 (1994), 141.
33. A. Fortin, A. Zine and J.-F. Agassant, J. Non-Newtonian Fluid Mech., 42
(1992), 1.
34. R. Gu6nette and M. Fortin, J. Non-Newtonian Fluid Mech., 60 (1995), 27.
35. M. Fortin and D. Esselaoui, Int. J. Numer. Methods Fluids, 7 (10) (1987),
1035.
36. M. E1 Hadj and P. A. Tanguy, J. Non-Newtonian Fluid Mech., 36 (1990),
333.
37. F. G. Basombrio, J. Non-Newtonian Fluid Mech., 39 (1991), 17.
38. J. P. Benqu6, G. Labadie and J. Ronat, Int. J. Numer. Meth. Fluids, 8
(1988), 869.
39. K. K. Kabanemi, F. Bertrand, P. A. Tanguy and A. A'it-Kadi, J. Non-
Newtonian Fluid Mech., 55 (1994), 283.
40. H. H. Hu and D. D. Joseph, J. Non-Newtonian Fluid Mech., 37 (1990),
347.
41. H. C. Choi, J. H. Song and J. Y. Yoo, J. Non-Newtonian Fluid Mech., 29
(1988), 347.
42. J. Y. Yoo and Y. Na, J. Non-Newtonian Fluid Mech., 39 (1991), 89.
43. S.-C. Xue, N. Phan-Thien and R. I. Tanner, J. Non-Newtonian Fluid
Mech., 59 (1995), 191.
382

44. X. Huang, N. Phan-Thien, R. I. Tanner, J. Non-Newtonian Fluid Mech.,


64 (1996), 71.
45. T. Sato and S. M. Richardson, J. Non-Newtonian Fluid Mech., 51 (1994),
249.
46. B. P. Leonard, Comput. Meth. Appl. Mech. Eng., 19 (1979), 59.
47. I. Babuska, Numer. Math., 16 (1971), 322.
48. F. Brezzi, Revue Franqaise d'Automatique Informatique Recherche
Op6rationelle, Analyse Num6rique, 8 (R-2) (1974), 129.
49. P. Saramito, Math. Modelling Num. Anal., 35 (1994), 1.
50. P. Saramito, J. Non-Newtonian Fluid Mech., 60 (1995), 199.
51. P. J. Roache, Computational Fluid Dynamics, Hermosa, 1976.
52. G. P. Sasmal, J. Non-Newtonian Fluid Mech., 56 (1995), 15.
53. F. H. Harlow and P. H. Welch, Phys. Fluids, 8 (1965), 2182.
54. L. S. Caretto, A. D. Gosman, S. V. Patankar and D. B. Spalding, Proc.
Third Int. Conf. Numer. Methods Fluid Dyn., Paris, 1972.
55. R. I. Issa, J. Comput. Phys., 62 (1986), 40.
56. J. P. van Doormal and G. D. Raithby, Num. Heat Transfer, 7 (1984), 147.
57. J. H. Ferziger and M. Peric, Computational Methods for Fluid Dynamics,
Springer, Berlin, 1996.
58. B. Gervang and P. S. Larsen, J. Non-Newtonian Fluid Mech., 39 (1991),
217.
59. X.-L. Luo, J. Non-Newtonian Fluid Mech., 64 (1996), 173.
60. M. Fortin and A. Fortin, Int. J. Num. Methods Fluids, 5 (1985), 911.
61. D. W. Peaceman and H. H. Rachford, J. Soc. Ind. Appl. Math., 3 (1955),
28.
62. X.-L. Luo, J. Non-Newtonian Fluid Mech., 63 (1996), 121.
63. M. J. Szady, T. R. Salamon, A. W. Liu, D. E. Bomside, R. C. Armstrong
and R. A. Brown, J. Non-Newtonian Fluid Mech., 59 (1995), 215.
64. A. J. Chorin, Math. Comp., 22 (1968), 745.
65. R. Temam, Theory and Numerical Analysis of the Navier-Stokes Equa-
tions, North-Holland, Armsterdam, 1977.
66. J. Kim and P. Moin, J. Comput. Phys., 59 (1985), 308.
67. J. Kim, P. Moin and R. D. Moser, J. Fluid Mech., 177 (1987), 133.
68. H. Choi, P. Moin and J. Kim, J. Fluid Mech, 262 (1994), 75.
69. H. Choi, P. Moin, J. Comp. Phys., 113 (1994), 1.
383

70. J. Donea, S. Giuliani, H. Laval and L. Quartapelle, Comput. Meth. Appl.


Mech. Eng., 30 (1982), 53.
71. B. Ramaswamy and T. C. Jue, Int. J. Num. Methods Fluids, 35 (1992),
671.
72. B. Ramaswamy, Comput. Fluids, 16 (1988), 349.
73. M. Kawahara and K. Ohmiya, Int. J. Num. Methods Fluids, 5 (1985), 981.
74. O. C. Zienkiewicz and J. Wu, Int. J. Num. Methods Eng., 32 (1991),
1189.
75. J. G. Rice and R. J. Schnipke, Comput. Meth. Appl. Mech. Eng., 58
(1986), 135.
76. A. C. Benim and W. Zinser, Comput. Meth. Appl. Mech. Eng., 57 (1986),
223.
77. H. G. Choi and J. Y. Yoo, Numer. Heat Transfer B, 25 (1994), 145.
78. H. G. Choi and J. Y. Yoo, Numer. Heat Transfer B, 28 (1995), 155.
79. H. G. Choi, H. Choi and J. Y. Yoo, Comput. Meth. Appl. Mech. Eng., 143
(1997), 333.
80. E. O. A. Carew, P. Townsend and M. F. Webster, J. Non-Newtonian
Fluid Mech., 50 (1993), 253.
81. A. Baloch, P. Townsend and M. F. Webster, J. Non-Newtonian Fluid
Mech., 54 (1994), 285.
82. D. M. Hawken, H. R. Tamaddon-Jahromi, P. Townsend and M. F. Web-
ster, Int. J. Num. Methods Fluids, 10 (1990), 327.
83. A. Baloch, P. Townsend and M. F. Webster, J. Non-Newtonian Fluid
Mech., 65 (1996), 133.
84. R. Sureshkumar, A. N. Beris and R. A. Handler, Phys. Fluids 9 (3)
(1997), 743.
85. T. Min, H. G. Choi, J. Y. Yoo and H. Choi, Int. J. Heat and Mass Trans-
fer, 40 (1997), 3689.
86. C. R. Swaminathan, V. R. Voller and S. V. Patankar, Finite Elements in
Analysis and Design, 13 (1993), 169.
385

CONSTITUTIVE EQUATIONS FROM TRANSIENT NETWORK


THEORY

C.F. Chan Man Fong and D. De Kee

Department of Chemical Engineering, Tulane University, New Orleans, LA


70118, USA

1. I N T R O D U C T I O N

An understanding of the flow properties of non-Newtonian fluids is


important in many areas of application. To solve flow problems, it is required
to introduce a constitutive equation that relates the stress tensor to various
kinematic tensors. The determination of a constitutive equation which
describes adequately the rheological properties of a fluid under all flow
conditions is one of the central problems in rheology. So far constitutive
equations have been developed based on either continuum mechanics or on the
molecular structure of the fluids.
In the continuum approach, no explicit consideration is given to the
molecular structure of the material. A relationship between suitable dynamic
and kinematic variables is postulated. The conditions that this relationship has
to satisfy were stated by Oldroyd [ 1].
A suitable dynamic variable is the extra stress tensor v and a suitable
kinematic variable is the relative right Cauchy-Green tensor C t. The tensor C t
relates the square of the relative distance between two arbitrary material points
at two different times, one at the present time t and the other at a past time t'.
The assumption that __zdepends on the history of _Ct_ from time t' - - ~ to the
present time t can be expressed mathematically as
t
1;- ~ (Ct) (1)
t ! "-" ~ oo
386

where _~ is an isotropic tensor-valued functional.


Equation (1) is the constitutive equation of a simple fluid.
The constitutive equation for a Newtonian fluid is
3C
=
(2)
t' =t

where 770 is the constant viscosity.


Various simplified forms of equation (1) have been proposed and are
discussed in [2-5].
In the molecular approach, the molecular structure of the fluid is taken into
consideration. In modeling a polymeric material, we first represent the
polymer molecules by mechanical models. We then introduce a probability
distribution and an evolution equation for the distribution. Finally we calculate
the average of all quantities so that a relationship between the macroscopic
quantities can be obtained. Three models, namely the bead-rod-spring model,
the transient network model, and the reptation model, have been popular
among rheologists. We shall discuss the development of the transient network
model after a brief description of the other two models.

2. B E A D - S P R I N G M O D E L

This model was proposed to describe dilute polymer solutions. The polymer
molecule is idealized as a dumbbell consisting of two beads, each of mass m,
joined by a spring. The beads are labeled 1 and 2; their vector positions are
r l a n d r 2 and R = ( r 2 - r l ) is the vector joining bead 1 to bead 2. The

polymer solution is modeled as dumbbells suspended in a Newtonian fluid of


constant viscosity r/s. We introduce a d i s t r i b u t i o n f u n c t i o n

F ( r 1 , r 2, trl, .7 2, t ) . such . that .F ( r 1,. r 2, 71


. , 7 2, .t ) dr .l d r 2 d r 1 d r 2 is the

number of dumbbells in the position range F__i t o F_.i + d r i and the beads have

velocities in the range r'i t o t~i + dT_i (i = 1, 2). It is customary to write F as a

product of a configuration distribution !/'t(rl,r2 , t) and a velocity


387

distribution ~ (F__'I, /~2) which is often assumed to be Maxwellian. If the

solution is homogeneous, 7~ can be written as

(L1, r2, t) = n ~ (_R, t) (3)

where n is the number density of dumbbells.


The conservation of the number of dumbbells leads to an equation of
continuity of the form (Bird et al. [6])

3t - - ~_R -It ~ 1 1 V (4)

where ~ 11denotes an average in the velocity space.


The forces acting on bead i are:
(a) the hydrodynamic f o r c e Ff h) which is assumed to be proportional to the

difference between the average bead velocity and the velocity of the
solution at that point. Assuming that the velocity of the solution is not
affected by the other beads, _F}h~ can be written as

_F{h) = -~ {I]~i ]]-Vi} (5)

where ~" is a constant and v i is the velocity of the solution at F_.i .

(b) the Brownian force _F~B~which is assumed to be of the form

F{ B)-- - k T ( ~ 2 n W / ~ r i ) (6)

where k is Boltzmann's constant and T is the temperature.

(c) the intramolecular force F~ ~ and, in the present model, it is the force

due to the spring. The forces on the two beads are equal and opposite
and a connector force F (c~ can be defined as
388

F (c)- _F~*)= -F~ *) (7)

In the absence of external forces and neglecting inertia terms, the equations
of motion for the beads are

--~ {[I i l ]] --V1} -- kT (~)2nW/arl) + _F(c)- 0 (8a)

--~ {[I i 2 ]] -- V2 } -- k T ( O 2 n U d / O L 2 ) - _F (c) - 0 (8b)

On subtracting we obtain

{[I ll~]] + (v 1 - v 2 ) } + 2kT (~)2n~/~)_R) + 2F(C) = 0 (9)

By considering the flow to be homogeneous, v 1 - V2 can be written as

V_.I--V 2 -- - _ g ' R (10)

where L is the velocity gradient.


Combining equations (9, 10) yields

~1I ~11 = ~ _L- _R- 2kT (a,env/a_13)- 2F (c) (11)

Substituting equation (11) into equation (4) yields

m 9
(L-R_) ~ - ( 2 k T / ~ ) ( ~ / ~ _ R ) - 2F(C)~/~I } (12)
~t

The total stress/7 for the solution can be written as

=l-=I = 1-=I__s + FI_p_ (13)

where =/-/sand Hp are respectively the contributions from the solvent and the
polymer.
The solvent is a Newtonian fluid and _/-/s is given by

Us - Ps---- ns ~ (14)
389

where Ps is the pressure associated with the solvent, 6 is the unit tensor, and
is the rate-of-strain tensor.
--(b)
The contributors to __/-/pare the connector, ~___p,
(c)/7 and the beads, l=lp . It is
shown in Bird et al. [6] that __/_/(pc)and __/_/(pb)are given by

(c) (c)
lip= = - n < R F ) (15a)

=l-=Ip(b) = 2 nkT 8 (15b)

where < > denotes the average over the configuration space.
Combining equations (13, 14, 15a, b) yields

__H - Ps ~ - rls "~ - n < R F (c)) + 2 n k T ~ (16)

It has been shown (Bird et al., [6]) that at equilibrium

n ( R F~C)>0 = n k T ~ (17)

From equations (16, 17) we deduce that

__H- (Ps + n k T ) ~ +~= (18a)

~_ = -rls T - n <R F(C)> + n k T 8 (18b)

Equation (18b) is the Kramers expression for the stress tensor. To calculate
<R F(C)) we usually need to solve equation (12) for V and then evaluate

f v R F (c) dR_ (= < R F(C))). It is difficult to obtain a closed form solution to


equation (12). However, if the spring is assumed to obey Hooke's law
(_F(C)=HR_, H is a constant), we can derive an expression for =v without having

to solve equation (12).


Starting from equation (12), it can be shown that (Bird et al. [6])
V
D<RR)_L.(RR)_<RR).L~r=(RR)_(4kT/~)~_(4/~)<RF Cc)) (19)
Dt = - -
390

where D/Dt is the substantial derivative, v denotes the upper convected


derivative, and t denotes the transpose.
From equations (18b, 19) we deduce that =rp (= __r+ 7/s _/79is given by

V
= ln~<RR> (20)
~P 4

Equation (20) is the Giesekus form of the stress tensor. For a Hookean
spring, we deduce from equation (18b) that =Vp is given by

~p = - n i l <R R> + nkT=8 (21)

V
Combining equations (19, 20, 21) and noting that =~ = - ~ yields

V
_'lYp + ~1 "l~p= - - n k T ~ , 1 "~ (22)

where A 1 = ~/4H is the relaxation time.


The extra stress ~" is given by

V
V
~-+ ~'I ~ = --nO (~r+ ~2 It) (23)

where 770 - 77s + n k T ~ 1 is the viscosity and ~ 2 = /]" 177s ] (77 s -I- n k T X 1) is the
retardation time.
Equation (23) is the constitutive equation of a dilute polymer solution
modeled as Hookean dumbbells suspended in a Newtonian solvent. Equation
(23) was also proposed by Oldroyd [ 1] from a continuum mechanics approach
and was labeled liquid B.
An integral form of equation (23) is (Bird et al. [6])

[(nkT/k ,) e -(t-t'),~''] C~ 1(t') dt' (24a)

-1
m ( t - t') _C t (t') dt' (24b)

m (t - t') = rlsS' (t - t') + (n k T / ~ 1) e-(t-t')/Z" (24c)


391

where C -l(t')
t is the relative Finger tensor, ~' ( t - t ' ) is the derivative with
respect to t' of the Dirac delta function , and m ( t - t') is the memory
function.
In this model, macromolecules are represented by single dumbbells.
Figure 1 illustrates a more realistic model for a macromolecule consisting of
N beads connected by ( N - 1) springs. In the Rouse model, the connectors are
Hookean springs and a constitutive equation can be deduced as in the case
where the macromolecules are assumed to be single dumbbells. Equation (24a)
is modified to obtain (Bird et al. [6])

~_ = - r l s ' l + nkT e P C_t ( t ' ) d t ' (25)

where/~/p is the relaxation time associated with dumbbell p.

_R2

Figure 1. Six beads connected by five springs.

The Rouse model is widely used to interpret linear viscoelastic


measurements. The dynamic viscosity 77' and the dynamic rigidity 77" are
given by
392

rl'-rls - nkT ~ ~,lp/[1 + (klpf.0) 21 (26a)


P

rl" - n k T ~ ~21pf'0 1[1 + (~ lp(.0) 2] (26b)


P

where co is the frequency of the imposed small amplitude oscillations.


The Aqp can be related to the molecular structure of the polymers (Rouse
[7]). Increasing the number of terms in the series allows for a better fit of 7?'
and 77". However, in steady viscometric flows the Rouse model predicts a
constant viscosity whereas for most polymer solutions the viscosity decreases
with increasing shear rate. Various improvements, such as replacing the
Hookean spring by a non-Hookean spring, including hydrodynamic interaction
and internal viscosity, have been proposed to the simple dumbbell model and
these modifications are thoroughly discussed in Bird et al. [6].
In the models discussed so far, it is assumed that the polymer chains move
freely in space, unhindered by the other chains. In concentrated solutions and
in polymer melts, the motion of a chain is constrained by other chains and this
constraint is taken into account in the next two models.

3. R E P T A T I O N M O D E L

3.1. D o i - E d w a r d s Model
In the reptation theory proposed by de Gennes [8], the constraint that
polymer chains cannot pass through each other is interpreted such that each
chain is confined to move inside an imaginary tube, as shown in Figure 2. The
chain is assumed to be longer than the tube and the slack is considered to be a
defect capable of moving along the tube. As a result of such motion, part of
the chain can move out of the original tube and a new tube is created. This
type of motion is called reptation. Based on this concept, Doi and Edwards [9]
derived a constitutive equation for polymeric liquids.
Doi and Edwards [9] introduced the concept of a primitive chain which is
the center line of the tube, as shown in Figure 2. The motion of the real chain
is represented by the motion of the primitive chain. The real chain wriggles
around the primitive chain. This motion is very rapid and an equilibrium is
established after a time Teq (equilibration time) which is proportional to N 2,
where N O is the degree of polymerization of the real chain. On a time scale
393

greater than Teq , this wriggling motion can be ignored and the motion of the
real chain coincides with the motion of the primitive chain.

4,--,,
/ \

f
/
/
\ /
o"

f
\ ,"
,_/

Figure 2. Polymeric chain ( ~ ) in an imaginary tube. (o 9 o): primitive chain.

The primitive chain is assumed to be composed of freely jointed segments


with step length a and arc length L. The stress "r of the primitive chain is

~= -- - < Z F i _ R i > (27)


i

where F i and g i refer to the force and end-to-end vector of segment i

respectively.
If the primitive chain is assumed to be continuous, any of its points at time t
can be denoted by R_ (s, t), where s is the contour length measured from one

end of the chain. The unit tangent u (s, t) to the primitive chain is given by

u (s, t) = ~ _R(s, t) / Os (28)


394

It is assumed that _F i e_i can be written as

F i R i = Feq tl (s) u (s) ds (29)

where Feq is the force acting on the primitive chain at equilibrium.


Combining equations (27, 28, 29) yields

X = < u (s) u (s) > ds (30)


-- L - -

where Go is a constant.
Suppose that the material is deformed such that a material particle which is
at x at present time t, was at x' at an earlier time t'. The deformation _E is

defined by

E = (Ox/~x') (31)

Doi and Edwards [9] assumed that the unit tangent vector _u' (s) at time t' is

related to u (s) at time t by (independent alignment approximation)

u'(s) = (E_o u ) / I E 9u l (32)

The length of the primitive chain L is assumed to be constant and the extra
stress tensor ~ can be written as

I: - - G o Q (33a)

(E 9 u) ( E - u)
Q - - = - (33b)
iE.ul ~
0

where < >0 is an average over u in equilibrium.

In equation (33a), the random motion of the primitive chain has not been
taken into account. The random motion of the chain is responsible for the
395

relaxation process. This random motion is governed approximately by


diffusion and a closed form solution can be obtained (Doi and Edwards [9]).
The stress _vis given by

~ =-G0f_toog ( t - t ' ) __Q[E_


' (t,_ t')]dt (34a)

(t) = ~ [8/(7~2p2"l;p)] exp (-t/'l;p) (34b)


p odd
"~p = L2/(D/I;2p 2) (34c)

where D is the constant diffusion coefficient of the random motion.


Equation (34a) is the equation proposed by Doi and Edwards [9] and they
have shown that it can be written as

~== -
f_ t M 1 [ t - t ' , I i ( t , t ' ) , I 2 ( t , t ' ) ] C t - l ( t ' ) + M 2 [ t - t ' , I i , I 2 ] ( _ C2t_- l) 1 dt' (35a)

M 1 = ~(t-t')Hl[Ii(t,t'),I2(t,t')] (35b)

M 2 = ~ (t-t')H2[II(t,t'),I2(t,t')] (35c)

where I1 and 12 are the invariants of ~_Ct1, H1 and H2 are functions which are
related to Q.
Equation (35a) is known in continuum mechanics as the K-BKZ equation
(Carreau et al. [4]). Note that the memory functions M 1 and M2 are written as
a product of a function of ( t - t ' ) only [/2 ( t - t ' ) ] and a function of
deformation only [H 1(I1 ,I2), H2(11 ,I2)]" The function p ( t - t') can be
obtained from linear viscoelasticity. Equations (25, 35 a to c) are similar if we
set 7"/~= 0, H 1 = - 1 , and H 2 = 0. Various modifications have been proposed to
the present model and they are discussed in [4, 5, 10].

3.2. Curtiss-Bird Model


Curtiss and Bird [ 11] presented another version of the reptation motion and
derived a constitutive equation based on their phase-space kinetic theory (Bird
et al. [6]. The polymer chain is modeled as a Kramers chain consisting of
396

N identical beads freely linked by ( N - 1) massless, rigid rods each of length


a, as shown in Figure 3.

C)
r

Figure 3. Segments of a Kramers chain.

The method employed to obtain a constitutive equation is similar to that used


in section 2. It is assumed that the differences between the hydrodynamic
forces on adjacent beads v and v + 1 can be written as

_Fv(h+)1 _--(h)
gV -- -- ~
13~V . { l l r- - v + l r l l - [ v- -( r - - +Lv
----V +1
) - v t- -r --
+Lv)]} (36a)

~v = ~ [ ~ - ( 1 - e ) u v u v] (36b)
m

where N is the average number of beads per chain, r v is the vector position of

bead v relative to the center of mass (see Figure 3), fl, ~', and e are constants,
and u v is the unit vector joining bead v to bead v + 1.

If the link tension coefficient e is one and the friction tensor ~'v is isotropic,
397

equation (36a) is similar to equation (5). The parameter e determines the


anisotropy of ~'v and Bird et al. [6] proposed e to be in the range 0.3 to 0.5.
The Doi-Edwards model corresponds to the case e = 0. The constant 13 is also
in the range 0.3 to 0.5. If the flow is assumed to be homogeneous, equation
(36a) simplifies to

F(h)
v+l --
F(h)
2_.V ~ --
Nlaa ~V
9 {iI -fi- V 11-[L- - 9 Uv] } (37)

In the calculation of the Brownian force, it is necessary to evaluate II 1~1i U_'j]]

and in order to do this, Curtiss and Bird [11] introduced the concept of
reptation. They assumed that

lI Lj L v ll = ( e : / 4 ) (Uj_ 1 4"- Uj) (Uv_ 1 + U_v) (38)

where a is a constant independent of j and v for all j and v from 2 to N - 2.


Equation (38) implies that the "average" velocity of a bead relative to the
center of mass is in the direction of the two rods connected to the bead. That is
to say, the motion of the bead is mainly in the direction of the backbone of the
chain, which is the concept of reptation.
In calculating the distribution function 7~, the mild curvature approximation
is introduced. This means that the orientations of adjacent links do not vary
significantly and the Kramers chain can be assumed to be continuous with
continuous derivatives.
Based on the assumptions stated earlier, Curtiss and Bird [11] obtained a
constitutive equation which can be written as

~ = = - N n k T I f _ t o o g ( t - t') =Q (E_)
f _ t dt'+
oo_ e ( t - t') S_=_(E; L__)dt'1 (39a)

( t - t') - (8/)v) ~ exp [-/I;2p2(t- t ' ) / s (39b)


p odd

[5 ( t - t') = (16//1;2~) Z (1/P 2) exp [-/1;2p2(t - t')/)~] (39c)


p odd
398

S= = (1/4rt) ~,L(t) 9 f u u u o du (39d)

where n is the density of polymer molecules, k is Boltzmann's constant, T is


the temperature, and ~ is a time constant.
If e = 0, equation (39a) reduces to the Doi-Edwards model with the time
constant being defined differently. Other forms of the Curtiss-Bird equation
are available and are given in Bird et al. [6].

4. NETWORK MODEL

4.1. Lodge Model


The present model, the transient network model, is an adaptation of the
network theories of rubber elasticity and was developed by Green and
Tobolsky [12], Lodge [13], and Yamamoto [14] for modeling polymer melts
and concentrated polymer solutions. In the network theory for solids, chemical
crosslinks are described as permanent junctions and are forced to move
together at all times. For liquids, which are capable of flowing, the junctions
are temporary and liable to be destroyed. Equally new junctions can be
formed. The polymeric material is represented by a network, as shown in
Figure 4. The molecular segment that joins two successive junctions is an
active segment, the chain that is connected to only one junction and has the
other end free is a dangling segment, and the chain that is not attached to the
network is a stray chain. As a first approximation, we neglect the contributions
of the dangling and stray segments to the stress generated during a
deformation.
Let R_ be the end-to-end vector of an active segment and 7~ (R_, t) be the

probability of finding a segment in the range R_ and R_+ dR. A balance of

segments yields (Bird et al. [6])

= (_R W) + c - 2 (40)
3t 3R

where c and 2? are the rates of creation and loss of segments respectively.
399

Segment "~ )
~' choin

~-Dongling
Loose end Segment

Figure 4. Polymer network.

Note that in equation (40), unlike in equation (4), we have allowed for the
loss and creation of segments. This loss and creation of segments is analogous
to the loss and creation of tubes in the Doi-Edwards model. In their model, the
loss and creation of tubes is due to the reptation of the primitive chain and in
the network model, there is as yet no definite molecular mechanism for the
creation and loss of segments. We need to assume empirically the form of c
and 2 .
If we assume that the junctions move affinely, it follows that

1~ = L o R (41)

For simplicity we assume that all the segments are alike and the tension
acting in each segment is F (c). The total stress/7 can be written as

I-I (t) = - n < F (c) R > (42)

where n is the number of segments.


400

To evaluate < if(c)_R> we need to solve for 7~ from equation (40) and this

implies that we need to postulate c and ,e. In the Lodge model, the following
assumptions are made

c = W0/~ (2, is a constant) (43a)

2 = W/~ (43b)

_F(c) = H_R (H is a constant) (43c)

where ~0 is the value of 7~ at equilibrium.


The choice of the same constant ~ in equations (43a, b) ensures that at
equilibrium there is no net loss or gain of segments.
Substituting equations (41, 43a, b) into equation (40) yields

OWOt = ~R~ (L= 9RW)_ + ~ ( W 0 - W) (44)

Equation (44) seems to be quite different from equation (12), however if


equation (44) is multiplied by R R and integrated over the R_-space, we obtain

(Bird et al. [6]; Chan Man Fong and De Kee [15])

V
<RR> = I((RR> -<RR>) (45a)

< R R > o = J _8 (45b)

where J is a constant.
Combining equations (42, 43c, 45a, b) yields
V
H+LH = -nil J8 (46)

Noting that
FI = P ~5 + "c (47a)

z_ = 0 at equilibrium (47b)
401

V
8 - - ~ (47c)

we deduce from equation (46) that

V
"c + )~'c - - r l "~ (48)

where 77 (= ~, n H J) is a constant.
Equation (48) is the constitutive equation of an upper convected Maxwell
fluid (Carreau et al. [4]) and an integral form of equation (48) is

- - m (t - t') C t (t') dt' (49a)

m (t - t') = (yl/~ 2) exp [-(t - t')/~,] (49b)

We note that in this case, as in the bead-spring model, we have been able to
derive a constitutive equation without having to solve for 7~. The time constant
~, for the network model is empirical whereas the time constant ~,1 in the bead-
spring model is related to the structural parameters of the model (equation 22).
The two models yield identical constitutive equations if the solvent viscosity is
negligible.
For clarity we have so far considered the case of a single time constant ~,.
We can generalize from one time constant to a multitude of constants ~p. We
can associate /~p with the different ages of the junctions (Lodge [16]) and to
each ~p corresponds a __rp. Each __rp satisfies equation (48) and =r (= ~ __Vp) is
p
given by equation (49a) with m ( t - t') defined by

m (t - t') - ~ (T~p/)~ 2p) exp [ - ( t - t' ) / ~ p ] (49c)


P

The Lodge model can describe many of the p h e n o m e n a associated with


linear viscoelasticity (Carreau et al. [4]) but it does not predict shear thinning
as observed in viscometric flows of polymeric fluids. One way of overcoming
this shortcoming is to allow c and ,g to be functions of a macroscopic variable,
such as an invariant of the shear rate, and this is considered next.
402

4.2. Shear Rate Dependent Models


Several authors [15, 17, 22] have assumed that c and ,e are functions of the
second invariant of the shear ~ defined by

= # (1/2) trl~ 2 (50)

We now consider the multimodes model proposed by De Kee and Carreau


[21 ]. Corresponding to the p-mode, the functions cp and ,gp are assumed to be

Cp(t) = Lp [~t(t)] Wpo, ~ p(t) - tlJp/Xp [~(t)] (51a, b)

where ~up0and ~p are the distribution functions associated with the p-mode at
equilibrium and at time t.
Substituting equations (5 l a, b) into the balance equation [equation (20)] for
the p-mode, multiplying the resulting expression by R R and integrating over

the R_-space, we obtain

v
< R R) p = Lp =8- < R R)p/~p (52a)

Lp " - (4rt/3) R4LpdR (52b)

Integrating equation (52a) yields

< R R)p = { Lp exp [ dt"/1;p(t")] C t 1(t') } dt' (53)


f~oo -- ftt'

It is assumed that L p and "t'pcan be written as

Lp = [Tlpfp('~)]/~p, "l;p = ~,pgp('y) (54a, b)


where r/p and Z p are constants and have dimension viscosity and time
respectively.
403

Following the usual procedure, it is deduced that the extra stress tensor T
can be written as

= -ffoo m (t, t', 3;) Ct 1(t') dt' (55a)

'

m ( t , t', 3~) - Z
YI pfp [~ (t')] exp
2
)~p ftt dt"
~,pgp [~(t")]
(55b)

The functions fp and gp cannot be deduced from the model and have to be
prescribed empirically. The following assumptions are made (De Kee [23])

fp = exp [-~tp (3 - 2c)] , p - 1, 2, ..., k (56a)

go = exp [-~tp (c - 1)], p - 1, 2 .... , k (56b)

fp = gp = 1, p = k + l , k+2, ... (56c)

where c and tp are constants.


Equations (56a to c) imply that there are various types of junctions, the rates
of loss and creation of the first k types are shear-rate dependent and the
remaining types are shear-rate independent. Equations (55a, b) have been
found to be adequate to describe the rheological properties of polymeric
systems and the predictions of equations (55a, b) compared to experimental
data will be examined later. Other forms of fp and gp, usually as rational
functions of ~, have been proposed and are discussed in Carreau et al. [4].

4.3. Stress Dependent Models


Kaye [24] assumed the rates of loss and formation of junctions are functions
of the invariants of the stress tensor. He proposed that cp and ,gp are functions
of Q1 and Q2 and they are defined by

Q1 - 12lx-212~ ' Q2 = 213lx - 9 I lx I 2x + 2 7 I 3x (57a, b)


404

Ilz - tr ~ , I2z = (1/2) [(trx=)2 - try=2] , I3x - det ~ (57c, d, e)

Note that I 1T, 12v, and 13 r are the invariants of __vand Q1 is positive for all
'~"~

Equations (5 l a, b, 54a, b) are now replaced by

Cp(t) = Lp [Ql(t), Q2(t)] W0 , 2p(t) = Wp/1:p [Ql(t), Q2(t)] (58a, b)

Lp - [lip gp (Q1, Q2)] [ )~p, "l:p - ~,p gp (Q1, Q2) (59a, b)

The constitutive equation is given by equation (55a) with the memory


function defined by

m (t,t',Q1, Q2) - ~
P I
TlPgP[QI(t')' Qz(t')] exp
~2
P
Iftt dt"
~pgp [QI (t,,), Q2(t,,)]
(6O)
tl
Kaye [24] considered one mode only (p = 1) and assumed g l to be a linear
function of Q1 and independent of Q2. In a viscometric flow, this model
predicts a viscosity which decreases with increasing shear rate and the ratio of
the first normal stress difference to the square of the shear stress is constant.
Phan-Thien and Tanner [25], Phan-Thien [26], and Tanner [27] have
assumed that the rates of loss and creation of junctions are functions of
<lR2]>p and (IRl2)p0 respectively. Since the segments are assumed to be

Hookean (=v is proportional to < R R>),< IRl2)p is proportional to tr v__p. Their

model is a stress dependent model. In addition they have assumed that the
deformation is non-affine and their model is discussed in section 4.6. The
Marrucci model [28] was proposed as a modification of the upper convected
Maxwell model but it has been shown that it is derivable from the network
theory (Jongschaap [29]) and is of the stress dependent type. Acierno et al. [28]
have also introduced a structural parameter related to the segment density in
their model and we shall discuss these models in section 7.
405

4.4. Strain Dependent Models


A strain dependent model was proposed by Wagner [30]. He assumed the
rate of creation Cp to be constant and the rate of loss ,s to depend on 11
-1
(= tr C_t ) and I 2 (= tr C_t ) . Further he proposed the loss of junctions to be due
to the Brownian motion (a constant rate of loss) and to the deformation.
Instead of equations (59a, b) we now have
Lp = l]p/~,p , ~p = ~,p + g (I1, I2) (61a, b)

where g (associated with the loss of junctions due to the deformation) is


independent of p (the type of junction).
The constitutive equation is given by equation (55a) and m can be written as

m (t,t , ,I1,I2) = E [ (lqp/~p)


P
2 exp IItt g
at"
--
[Ii(t")i I2(t")] t (62a)

= h (II,I2) Z 2
('qp/~p) e_(t_t, )/~,p (62b)
P

The damping function h is defined by

h (I 1' 12) -" expIftt' dt" (63)


g (I 1' 12)

and is to be determined empirically.


Based on relaxation data, Wagner [30] proposed the following form for h

h = exp [-[3 4t~ I 1 + (1--t~) 12 - 3 ] (64)

where a and fl are positive constants.


Wagner and Stephenson [31] added a further restriction on the function h.
Equation (64) is valid only when the deformation is a non-decreasing function
in time. If the deformation is a decreasing function of time, h is taken to be the
minimum value of h over the relevant period. Physically this means that
junctions which are lost over a previous increasing period of deformation are
lost forever.
406

Comparing equations (55a, 62a) with equations (35 a to c), we observe the
Wagner model is a special case of the Doi-Edwards model or the K-BKZ
model.

4.5. Y a m a m o t o Model
In all the transient network models considered so far, the functions c and A?
depend on the macroscopic variables and it was possible to deduce a
constitutive equation in a closed form. Yamamoto [14] proposed that the
rates of creation and loss of segments depend on the end-to-end vector
R. It is assumed that c and ,8 can be written as

c - fl (_R) q J 0 , ~ - f2(-R)~I j (65a, b)

Since at equilibrium c = 2 , we deduce from equations (40, 65a, b) that

3q'
- (R W) + f ( R ) ( W o - W) (66)
~)t ~)R - -

where f = fl - f2-
If we proceed as in the previous cases, that is by multiplying equation (66)
by R R and integrating over the configuration space, we are led to evaluate

<f(R_) R R ) and we can only evaluate this integral if !P is known. But so far

we have not been able to obtain an analytic solution to equation (66). The
Yamamoto model has no closed form constitutive equation.
Petruccione and Biller [32] and Herman and Petruccione [33] have solved
equation (66) and its multimode version by a numerical simulation method.
The process governing equation (66) is simulated as a stochastic process and
the amount of computation required to determine the stress in a given flow is
reasonable. The agreement between predictions and experimental data in the
case of an elongational flow is good.

4.6. Non-Affine Deformation


So far we have assumed that the junctions move affinely with the
macroscopic imposed velocity. Lodge et al. [34] and Giesekus [35] have
407

expressed reservations concerning this affine motion assumption. A non-affine


model was proposed by [25, 36]. Equation (41) is modified to

1~ = Lo R (67a)

L = L - (112~) ~ (67b)

where ~ is a constant and is known as the slip parameter.


Substituting equations (43a to c, 67a) into equation (40) and proceeding in
the usual manner yields an analog of equation (40) which can be written as [ 15]

~=+~= = - r 1 7 (68a)

7 = (1 - ~) ~ (68b)

~2
__z"is the Gordon-Schowalter [37] derivative of =vand is given by

Q D'~ .-, ---t


1; -- = Lox-I:oL (69a)
Dt = = = =
V
= 1: + (,~12) (~t,~ + ~ . ~) (69b)

The Gordon-Schowalter derivative reduces to the Oldroyd upper, lower


convected derivative and to the Jaumann derivative if ~ is set to 0, 2, and 1
respectively. Since ~ is a constant, we may write the simplest constitutive
equation based on a transient network model with non-affine deformation as
[combining equations (68a, 69a)]

+ + (,,?. + ,c._ = -n ";' (70)

Equation (70) is a special case of the Oldroyd [1] model derived from
continuum considerations. Equation (70) predicts a variable viscosity and the
ratio of the secondary (N2) to the primary (N]) normal stress differences in
viscometric flows is given by
408

N2 _ ~ (71)
N1 2

By allowing for a non-affine motion, the simplest model [equation (48)] of


the network model has been improved considerably in the sense that a variable
viscosity and a secondary normal stress difference can be predicted.
The constitutive equations obtained on the assumption of affine deformation
can be extended to non-affine deformation by replacing the Oldroyd derivative
by the Gordon-Schowalter derivative if the equations are written in the
differential form. If the equations are written in the integral form, the Finger
-1 --~-1
tensor _Ct-1 is replaced by a modified Finger tensor C
t= . The tensor C
t= is
calculated from the reduced velocity _v [= (1-~/2) v] instead of from v.

The Phan-Thien-Tanner [25] model, which is a stress dependent and non-


affine deformation model, is written as
= ~ ~ (72a)
_~P
P
o
Y (tr ~ p) ~=p + )~ p ~= = -- TI p ~t_ (72b)

Two empirical forms of Y (tr r=p) have been proposed and they are

1 - ~ (tr ~ p) ~ p/lip
y (73a, b)
exp [-~ (tr ~p) ~ p [ ] q p ]

where e is a non-negative constant.


If e - 0 , equation (72b) simplifies to equation (70). A non-zero value of e
allows for a finite extensional viscosity (r/E) for all finite extensional rates (~).
If equation (73a) is chosen, r/e increases monotonicly to a finite limit as ~ tends
to infinity; equation (73b) predicts that r/e goes through a maximum as ~ goes
to infinity (Bird et al. [6]).
The extra stress z is arbitrary to the extent of an arbitrary isotropic
pressure and C -t 1 can be replaced by (_Ct 1_ _S) o r ( ~ - - C -1
t ) in the constitutive

equation. The same applies to C t .


409

An alternative form of equation (49a) is obtained by integrating it by parts.


The result is

~= = f _ t o o G ( t - t ' ) _ E ( t , t ' ) . ~ (=t ' ) . _ E t (_t , t ' ) d t ' _ (74)

where G is the modulus and is related to m ( t - t') by

G (t - t') = m (t - t') (75)


~t'

If the junctions move non-affinely, equation (74) is written as

= G (t - t') E_ (t, t ' ) o "~ 9 E_t (t, t') dt' (76)

where E is calculated from v instead of from v.

In this section we have considered two major modifications of the Lodge


model, namely variable rates of loss and creation of junctions and the non-
affine motion of the junctions. In the next section we shall examine the role of
the dangling and free segments.

5. RECENT DEVELOPMENTS IN N E T W O R K THEORY

5.1. Stress Jump


The stress jump is defined as the instantaneous finite change in the stress due
to a sudden finite change in the rate of deformation. Liang and Mackay [38]
have reported measurements of stress jump at the cessation of shear flow for
solutions of xanthan gum.
Not all constitutive equations can predict a stress jump. The equations as
given in section 4 do not predict a stress jump. These equations are valid on a
time scale greater that the equilibration time (Teq) and the rapid change in the
stress due to the sudden finite change in the shear rate is ignored. The model
proposed by De Kee and Carreau [21 ], discussed in section 4.2, can predict the
stress jump if the two conditions they imposed are included. These two
conditions are
410

M
Z Tip exp ( - t / ~ p )
p=k+ 1
= I]oo [1 - H ( t ) ] (77a)

M
2Z Tlp(t + ~p) exp ( - t / ~ p ) = ~tl~ [ 1 - H ( t ) ] (77b)
p=k+ 1

where 7/00and ~1oo are the limiting infinite viscosity and primary normal stress
coefficient respectively and H (t) is the Heaviside step function.
Note that equations (77a, b) imply a singularity at t = 0 and are valid for
~ p - - > 0. Further, it is not possible to distinguish the various modes
(p = k + 1, k + 2, ...) and there is no loss of generality in confining to one
mode only (p = M).
De Kee and Chan Man Fong [39] have shown that equations (77a, b) are
equivalent to introducing a singularity to the memory function m in equation
(55b). This equation is now written as
m = mL + m s (78a)

mL= ~
p=l l rlpfp
~p
2 exp
Iftt ~dt"
g
p p 1[ (78b)

ms - 11oo8'(t - t') - (~1oo/2) 8"(t - t") (78c)

where ~' and 6" are the first and second derivatives of the Dirac delta
function.
Within the network theory we can interpret m L as being the response of the
active segments associated with a time scale greater than Teq. The function ms
represents the response of the solvent molecules [see equation (24c)], the stray,
and the dangling segments and is associated with a time scale less than Teq. To
predict the stress jump phenomenon it is necessary to include the short term
response which is the Newtonian and possibly also slightly viscoelastic
responses [10]. This in turn implies that the memory function must have a
singularity [40].
Other models that can predict the stress jump are discussed in [38, 41, 42].
411

5.2. Shear Thickening


In recent years there has been a growing interest in complex fluids that
exhibits shear thickening. Ahn and Osaki [43] have tabulated various systems
that are shear thickening. They have classified the rheological flow curves into
four types: thinning-thickening, thinning-thickening-thinning, thickening only,
and thickening-thinning. The cause of the shear thickening is the flow induced
complex formation of the material and the origin of this complex formation
can be electrostatic interaction, hydrogen bonding, etc.
Chan Man Fong and De Kee [15, 22] have included a Coulombic force
between the junctions of the network model. They assumed that c and ,e in
equation (40) are given by

c = kcW 0 + k 1f2 (~/) q~ (79a)

= kc tlJ + k,e f l (~t) ~ (79b)

where k c, k l, and k~ are constants and the functions fl and f2 are prescribed
later.
To include the electrostatic force, they wrote F_(c) as

F (c) = H R_ - (2q2/e R 3) R_ (80)

where H, q, and e are constants.


Proceeding in the usual manner, they deduced a constitutive equation valid
for steady flows and for a time scale greater than Teq. Their constitutive
equation for a single mode can be written as

1;
~
= + )l, m I;
= = (1 - ~ ) )~mkT
/n o ~
= - b--E----
~
i n [2 (k 2 f 1 - k 1 fa) 7= + 7]
= (81)

where ~ m -" (kc + k ~ . f l - klf2) -1, E = ( k c + k l f 2 - k,e f l), n o and n are the
number density junctions at equilibrium and at present time respectively, ~ is
the slip parameter, k is Boltzmann's constant, T is the temperature, b and I
are constants.
412

Equation (81) can be used to predict shear thinning followed by shear


thickening.
Tanaka and Edwards [44], Wang [45], and van der Brule and Hoogerbrugge
[46] have attributed shear thickening to the dangling segments being trapped by
the active network. This leads to an increase in the number of junctions and
hence an increase in viscosity. The stray segments may also join the active
network. By postulating approximately the form of c and ,8 and the
probabilities of the stray and dangling segments joining the network it is
possible to describe the four types of rheograms mentioned earlier. In the van
der Brule and Hoogerbrugge [46] model the onset of shear thickening occurs at
a constant value of shear stress independent of the model parameters.
A non-Gaussian network can also contribute to shear thickening [16, 47, 48].
By considering both non-affine deformation which leads to shear thinning and
a non-Gaussian network which contributes to shear thickening, Vrahopoulou
and Mc Hugh [47] have been able to predict both shear thinning and shear
thickening.

5.3. Slip-Link N e t w o r k
In section 3.1 the reptation process is modeled as the motion of the primitive
chain inside a tube. Doi and Edwards [9] have also proposed two other
alternative models, namely the cage model and the slip-link model. Slip-links
are small rings through which the chain can pass freely, as shown in Figure 5.
Wagner and Schaeffer [49] have implanted the slip-link model in the transient
network theory. The transient slip-link network is a network of active
Gaussian segments between slip-links. Suppose that at equilibrium the distance
between two slip-links is 2' 0 with n o monomer units. The end-to-end vector
can be written as

L = 2 ou (82)

where u is a unit vector.

After deformation, r is transformed to r ' which is expressed as

L' - 2oU' (83)


413

Figure 5. Slip-link network at equilibrium and after deformation.

Due to the possibility of slip the number of monomers is no longer no and it


is assumed that the number of monomers n is a function of u' ( - I _u' I). A slip

function S (u') is defined as

S (u') = n (u') / n o (84)

The rates of creation and loss of junctions are assumed to be independent of


deformation as in the case of the Lodge model, but due to slip there is
possibility of disentanglement. A disentanglement function D (u') is introduced
and is also assumed to be function of u'. The extra stress v is now given by

-l_~m(t-t')g<(D/S) u'u'>dt' (85)


q l r

where m ( t - t') is given by equation (49b) and g is unspecified.


A mass balance of network segments with regard to slip and disentanglement
leads to

<D/S> = 1 (86)
414

If S = D = 1 and g = 3, equation (85) reduces to equation (49a). By suitable


choices of S, D, and g, Wagner and Schaeffer [49] have shown that equation
(85) reduces to the Wagner model and the reptation model.

6. C O M P A R I S O N WITH E X P E R I M E N T A L DATA

6.1. Viscometric Flows


To test the usefulness of constitutive equations it is necessary to compare
their predictions with experimental observations in well defined flows. One
such flow is the viscometric flow. Referred to the Cartesian coordinate system
(x, y, z) the velocity field is given by

v x - y ~ (y), Vy = Vz = 0 (87a, b, c)

where ~ (y) is the shear rate.


The viscometric flow is characterized by three material functions, the
viscosity function (r/), the primary ('Cxx-'ryy) and the secondary ('ryy- Z'zz)
normal stress differences. The viscosity and the normal stress coefficients gti
(i = 1, 2) are defined by

~yx -- -- Yl (~t) ~ (88a)

%xx -- Xyy -- --/[/l(~t ) ~r (88b)

Xyy -- ~zz = --/If 2 (~t) ~/2 (88C)

In equations (88a to c), the functions are defined when the equilibrium state
has been reached. For viscoelastic systems it takes a finite time to attain the
steady state. If a constant shear rate )'oo is applied at time t = 0, we define the
growth functions 77+, I/t~, and lit~ by

11+ (t, ~'oo) = -'l~yx (t, ~/oo) / Too (89a)

V~ (t, Z/oo) - - [a:xx (t, ~(oo)- %yy (t, ~(oo)1 / ~2 (89b)

~g~ (t, "]'oo) - - [~yy (t, Too)- "c= (t, ~'oo)] / ~/2 (89c)
415

Similarly if a constant shear rate ~'0 is kept for a sufficiently long time until
the steady state is established and the applied shear rate is stopped, which can
be taken to be at time t = 0, the stress relaxation functions r/-, gt l, and ~2 are
defined by

rl-(t, - -Tyx (t, ;/0) / Y0 (90a)

~ (t, Yo) - - ['Cxx (t, ~'o)- '[yy (t, ~'o)] / ~/2 (90b)

~g2(t, Z/o) = - ['l:yy (t, ~/o)- 1:~z(t, ;/o)] / ~o2 (90c)

For the De Kee-Carreau model (equations 55a, 78a to c) the material


functions are
k
r I (~t) = Z TIp exp (-tp~t) + 1]oo (91a)
p=l
k
tJl 1 (~t) _-- 2 ~ Tlp)~ p e x p (-Ctp~/) + ~1oo (91b)
p=l
k
rl-(t, ~/0) - ~ Tip [exp (-tp~/0)] [exp (-t/~,p)] + 1]oo [1 - H (t)] (91c)
p=l
k
~ ( t , ~/0) = 2 ~ TIp~ p [exp (-Ctp~t0) ] [exp (-t/~,p)] + Igloo [1 - H ( t ) ] (91d)
p=l

11+ (t, ~/oo) = n (~'oo) + ~ k Enp


p=l
t [1 - exp (-tp~/oo(2 - c))] - exp (-tp~/oo) /
x exp {-t/)~p exp (-tp~/oo(C- 1))/1 + rloo[H(t)- 1] (91e)

+
I1/1 (t, "Yoo) -- /[/1 (~oo) 4- Z np k L {t [1-exp(-tpyoo(2 - c))] - 2t exp (-tp~/oo)
p=l

-2)~peXp(-Ctp~/oo)l x exp l-t/)Lp exp (-tp~/oo(C- l))}l +/1/loo [ H ( t ) - l ] (91f)

For this model the secondary normal stress difference is zero.


416

Figure 6 compares the predicted and measured values of 7"/and Iffl for a 6.0
mass % p o l y i s o b u t y l e n e solution in Primol 355, a pharmaceutical grade white
oil with a v i s c o s i t y of about 0.15 Pa~ at 298 K. Figures 7 to 9 s h o w the
theoretical and experimental values of r/-, gt 1 , a n d gt I for various values o f
shear rates for the PIB solution.

I I I I ! I

3 _ iO 4
I0

O
n

3
i0 2 - I0

D-

It"

I 10 2
I0 L ~ I ........ 1 I I 1 I
0 I 2 3 4 5 6 7
~' is-' ~

Figure 6. V i s c o s i t y and primary normal stress c o e f f i c i e n t for a 6 %


p o l y i s o b u t y l e n e solution in Primol 355. 770 = 6 , 5 1 3 Pa-s, r/1 = 4 , 9 6 3 Pa~
7/2 = 1 , 2 2 8 P a . s , 7"/3 = 2 6 3 P a . s , 7"/0o = 58 Pa-s, tl - 6 6 . 2 4 s, t2 - 6 . 4 5 s,
t 3 - 0 . 6 3 2 s, 2,1 = 7 9 . 4 1 s, ~2 = 19.65 s, 23 = 4.21 s, c - 1.21, gtloo =
140 Pa.s 2. 9 m o d e l with tp - 10 tp+l.
417

08 \\\~\
0.7

o ,o 2'0 3b
t(s)

Figure 7. Shear stress relaxation for the 6 % PIB solution of Figure 6.


~" data; 9 model predictions [equation (91c) with k - 3].

,o1
09
Q8
~'ox~/~,,,,, ~ =4.34 S-t . . _
.-. 0.7-
=

._~_ 06.

.~ 0.5-

.,., 0.4-

~"-- 0.3- \\\


0.2-
\ \\\ \ ----
0.1-

0.0
0 20 4'0 6'0
t(s)

Figure 8. Normal stress relaxation for the 6 % PIB solution of Figure 6.


~" data; 9 model predictions [equation (91d) with k - 3].
418

~=o = 4 . 3 4 S -I

~,00= 1.37S - I

*-0.6- I
~,0 = 0.137 s - !

"T- T T T
0 2O 40 6O
t(s)

Figure 9. Normal stress growth for the 6 % PIB solution of Figure 6.


~" data; 9 model predictions [equation (91f) with k = 3].

It can be seen that the steady values of 7/and ~l are accurately predicted by
the model. Accurate transient data are difficult to obtain and the discrepancy
between the experimental data and theoretical predictions in Figures 7 to 9 can
be attributed to experimental error associated with the lack of stiffness of the
Weissenberg rheogoniometer model R-18 used for these experiments. High
rigidity is required to reduce coupling effects between the sample and the
instrument which should be capable of very rapid acceleration.
It was mentioned in section 5.1 that Liang and Mackay [38] have reported
measurements of shear stress jumps. They defined a stress jump ratio R as
follows

R-(t, T0) = rl-(t, T0) / [11 (~/0) - TIs] (92)


419

Substituting equations (9 l a, 9 l c) into equation (92) yields

R (0+, ~0) = 1 [1 -I- (Tloo--Yls)/Z lip exp(-tp~/0)] (93)


p=l
Figure 10 compares equation (93) with the experimental measurements of
Liang and Mackay [38]. We observe that R - ( 0 + , ~)0) is a decreasing function

of ~'o in agreement with the experimental data.

1.0

0.7
_ ~"~'~|~-.,,~, .,.-'..,
0.5

0.3 -- II

0.2 -

0. I I I I I I I I '.l
i I

0. I 0.2 0.5 I 2 5 I0 20 50

Figure 10. Predicted and measured stress jump ratio for solutions in xanthan
gum in a mixture of 75 mass % fructose and 25 mass % water.
0.05 mass % r/1 = 13.48 Pa.s, r/2 = 2.28 Pa.s, 77oo= 0.83 Pa-s,
t 1 = 1 0 . 1 3 s , t 2 = 0.28 s.
9 99 0.025 mass % r/~ = 3.13 Pa.s, q2 = 0.91 Pa.s, r/= = 0.64 Paos,
tl = 6.02 s, t2 = 0 . 1 6 s.
0.01 mass % r/~ = 0.61 Pa.s, 772 = 0.33 Pa.s, r/= = 0.53 Paos,
tl=4.69s, t2 = 0 . 2 1 s.
e , II, A : experimental data for 0.05, 0.025, and 0.01 mass % respectively.

Furthermore R - ( 0 + , 70) is less than 1 and this implies that r/= > r/s. This is
in accord with our interpretation that q= represents the viscosity of the solvent
and the stray and dangling segments.
420

From equation (91a) it is seen that q is a decreasing function of ~ and to


predict shear thickening we can use equation (81). Chan Man Fong and De Kee
[ 15] have assumed that
k~ f l - kl f2 - kc d~/m (94)

where d and m are constants and ~ is the dimensionless form of ~.


Combining equations (81, 87, 94) yields
1"1 = 1"10 (1 + 2~kcd~/m) ] (1 + d~/m) (95a)

~ = 2r10 [~ + ~ (1 + 2 ~, k~ d ~/m)] ] (1 + d ~ m) (95b)

-- 1 / k c (1 + d ~/m) (95c)

where 7"/0and ~ are constants.


Figure 11 compares the values of q and ~l given by equations (95a, b) with
experimental values of a partially hydrolyzed polyacrylamide solution
containing 2 g/2 of NaC1 reported by Ait-Kadi [50].

.-,..

tO

g.
0.5
i ,, ,, \A\ /
--2

- 0.5
g.
--3-
- 0.2

-0.1

0.1 I , , I, ,,,I I , , I ,,,,I I , , 0.05


I0 ~ I01 I0 z

Figure 11. Plot of viscosity (A) and primary normal stress coefficient (A)
versus shear rate (~) for a partially hydrolyzed polyacrylamide solution
containing 2 g/2, of NaC1. Model parameters are: 7/0- 0.50 Pa-s, kc =
0.040 s -1, ~ = 0.016 s, d = 0.13, m - 0.95.
421

6.2. M u l t i s t e p Flows
In section 6.1 we have considered the case of a constant shear rate being
applied or removed at time t = 0. It is also possible to apply finite increments
of shear rates and measure the corresponding stresses. Xu et al. [51, 52] have
tested the De Kee-Carreau model for various multi-rate-step flows. Figure 12
shows the calculated and the measured shear stress for a PDMS sample (a
viscoelastic standard material provided by Rheometrics) in a concave step
flow. In this flow a constant shear rate ~ is applied at t = 0 and is maintained
until the steady state is reached. The shear rate is then halved and is kept at this
value (~/2) for a finite time after which it is increased to its original value (~5).
When the steady state is reached, the shearing is stopped.

~6
'0
4 _

2
?

ol
o o o o o o

1 I I I
0 20 40 60 80 I00 120 140
t(s)

Figure 12. Comparison of model predictions ~ : equations (55a to 56c) with


shear stress data (o) for concave steps with different levels.
Upper Curve" ~ - 0.5, 0.25, 0.5, 0.0 s-1. Middle Curve: ~ - 0.25, 0.125, 0.15,
0.0 s-1. Lower Curve: ~ - 0.125, 0.06, 0.125, 0.0 s-1. The calculation is based
on the measured spectrum from dynamic data for the PDMS sample with c =
1.3 and f0 = 1.45 for all steps.
422

Figure 13 compares the theoretical and experimental values of "Cyx for the
same material in a reverse shear rate flow. In this flow the fluid is subjected to
a constant shear rate ~ at time t = 0 and after the steady state is attained the
shear rate is reversed in direction but equal in magnitude (-~). This cycle is
repeated.

10"

5 x 10 3

A
o
O.

-5xlO 3 _

I I I I I
0 20 40 60 80 I00 120 140

t (s)

Figure 13. Shear stress for reversed shear steps. Model predictions
equations (55a to 56c). The model parameters are those of Figure 12.

It can be seen in Figures 12 and 13 that the De Kee-Carreau model


adequately predicts the response of the material. Details of calculations and
experimental procedures are given in Xu et al. [52].

7. D I S C U S S I O N

In this chapter, constitutive equations have been derived in a fixed


coordinate system. The constitutive equations can also be deduced in a
convected (body) coordinate system and via an energy method. Lodge [ 16, 53]
has discussed the merits of adopting the convected coordinate system in the
formulation of constitutive equations.
423

Several constitutive equations based on structural kinetics have been shown


to be related to the transient network theory. The Marrucci model mentioned
in section 4.3 is one of them. The other models are discussed in De Kee [54].
These structural equations can be quite useful at describing the rheological
properties of complex materials. De Kee and Chan Man Fong [55] have
explored the capability of a structural equation in various flow situations. It is
shown in sections 6.1 and 6.2 that the De Kee-Carreau model, a shear rate
dependent model, can predict satisfactorily the rheological properties of
polymeric systems. However, the shear rate dependent models have been
objected to on the ground that they do not recover the linear viscoelastic
behavior in small amplitude oscillatory flows. This objection can be refuted by
noting that when the strain is small, the creation and loss of junctions are due
to Brownian motion [6, 22, 31] and are independent of the imposed flow. In
equations (56a, b), fp and gp are constants and we recover the Lodge model
which reduces to linear viscoelasticity in small amplitude oscillatory flows.
Ahn and Osaki [43] have examined the cases where fp and gp a r e in turn
functions of strain rate, strain, chain length, and effective strain which is
defined as the ratio of the primary normal stress difference to twice the shear
stress. They found that the predictions obtained by assuming fp and gp t o be
functions of the effective strain agreed best with their experimental data. Since
both the primary normal stress difference and the shear stress are functions of
the shear rate, we may conclude that fp and gp are functions of the shear rate.
Hinch [56] in a computer simulation of the uncoiling of a polymer molecule in
an elongational flow has found that the stress generated is strain rate dependent
and not strain dependent. Further support for the strain rate dependent
equation is given in Carreau et al. [4] and Macdonald and Carreau [57].
At present the model parameters of the network theory are not related to the
molecular structure of the material and it is desirable to seek such a
connection. Lodge et al. [34] have proposed to relate the rates of creation and
loss to the molecular weight. In Figure 14 we have plotted tl and t2 of the De
Kee-Carreau model against 7/0 M w for various polymer solutions. The curves
are parallel straight lines implying that for all p, tp is proportional to
(770 Mw)~, where n is a constant depending on the polymer solutions. We also
deduce that t2, t3 .... are multiples of tl and De Kee [23] found that the factor
0.1 fits the data (see Figure 6).
424

9 9

I 1 1800
I 0"s i i , , ,,,l
K)* iO? _ sOe iO t
%Mw

Figure 14. Correlation between the parameters tp and the product 77o M w .

The reptation model (section 3.1) has been improved by the des Cloizeaux
[58, 59] double reptation model and he has been able to deduce that the
viscosity 77 is proportional to M 3.4, where M is the molecular weight. In the
double reptation model, the reptation of a stress point is considered in addition
to the reptation inside the tube. A stress point is a point of entanglement of two
polymers and if either of the two polymers reptates out of the stress point, the
stress disappears. The stress point is the junction in the network theory and this
prompts Mead [60] to state that "the evolution of molecular constitutive
equations has gone full circle in the short span of 35 years." In the next circle,
we need to examine the process of loss and creation of junctions and the
deformation of the segments. The works of Wagner and Schaeffer [49] and des
Cloizeaux [58, 59] which combine both the reptation and network models need
further exploration.
425

REFERENCES

~ Oldroyd, J.G., Proc. Roy. Soc., A200 (1950) 523; A245 (1958) 278.
2. Barnes, H.A., J.F. Hutton, and K. Walters, An Introduction to Rheology,
Elsevier, New York, NY (1989).
. Bird, R.B., R.C. Armstrong, and O. Hassager, Dynamics of Polymeric
Liquids, Vol. 1 - Fluid Mech., Second Ed., Wiley, New York, NY
(1987).
, Carreau, P.J., D. De Kee, and R.J. Chhabra, Polymer Rheology:
Principles and Applications, Hanser, New York, NY (1997).
. Larson, R.G., Constitutive Equations for Polymer Melts and Solutions,
Butterworths, Boston, MA (1988).
. Bird, R.B., C.F. Curtiss, R.C. Armstrong, and O. Hassager, Dynamics
of Polymeric Liquids, Vol. 2 - Kinetic Theory, Second Ed., Wiley, New
York, NY (1987).
~ Rouse, P.E., J. Chem. Phys., 21 (1953) 1272.
8. de Gennes, P.G., J. Chem. Phys., 55 (1971) 572.
9. Doi, M. and S.F. Edwards, J. Chem. Soc. Faraday Trans. II, 74 (1978)
1789, 1802, 1818; 75 (1979) 38.
10. Doi, M. and S.F. Edwards, The Theory of Polymer Dynamics, Oxford
University Press, New York, NY (1986).
11. Curtiss, C.F. and R.B. Bird, J. Chem. Phys., 74 (1981) 2016, 2026.
12. Green, M.S. and A.V. Tobolsky, J. Chem. Phys., 14 (1946) 80.
13. Lodge, A.S., Trans. Faraday Soc., 52 (1956) 120.
14. Yamamoto, M., J. Phys. Soc. Japan, 11 (1956) 413; 12 (1957) 1148; 13
(1958) 1200.
15. Chan Man Fong, C.F. and D. De Kee, Physica, A218 (1995) 56.
16. Lodge, A.S., Rheol. Acta, 7 (1968) 379.
17. Bogue, D.C., Ind. Eng. Chem. Fund., 5 (1966) 253.
18. Bird, R.B. and P.J. Carreau, Chem. Eng. Sci., 23 (1968) 427.
19. Meister, B.J., Trans. Soc. Rheol., 15 (1971) 63.
20. Carreau, P.J., Trans. Soc. Rheol., 16 (1972) 99.
21. De Kee, D. and P.J. Carreau, J. Non-Newtonian Fluid Mech., 6 (1979)
127.
426

22. Chan Man Fong, C.F. and D. De Kee, J. Non-Newtonian Fluid Mech.,
57 (1995) 39.
23. De Kee, D., Ph.D. Thesis, University of Montreal, Quebec (1977).
24. Kaye, A., Brit. J. Appl. Phys., 17 (1966) 803.
25. Phan-Thien, N. and R.I. Tanner, J. Non-Newtonian Fluid Mech., 2
(1977) 353.
26. Phan-Thien, N., J. Rheol., 22 (1978) 259.
27. Tanner, R.I., J. Non-Newtonian Fluid Mech., 5 (1979) 103.
28. Acierno, D., F.P. La Mantia, G. Marrucci, and G. Titomanlio, J. Non-
Newtonian Fluid Mech., 1 (1976) 125.
29. Jongschaap, R.J.J., J. Non-Newtonian Fluid Mech., 8 (1981) 183.
30. Wagner, M.H., Rheol. Acta, 18 (1979) 33.
31. Wagner, M.H. and S.E. Stephenson, J. Rheol., 23 (1979) 491.
32. Petruccione, F. and P. Biller, J. Chem. Phys., 89 (1988) 577.
33. Herman, W. and F. Petruccione, J. Rheol., 36 (1992) 1461.
34. Lodge, A.S., R.C. Armstrong, M.H. Wagner, and H.H. Winter, Pure &
Appl. Chem., 54 (1982) 1349.
35. Giesekus, H., Viscoelasticity and Rheology, edited by A.S. Lodge, M.
Renardy, and J.A. Nohel, Academic Press, New York, NY (1985) 157.
36. Johnson, C.W. and D. Segalman, J. Non-Newtonian Fluid Mech., 2
(1977) 255.
37. Gordon, R.J. and W.R. Schowalter, Trans. Soc. Rheol., 16 (1972) 79.
38. Liang, C.H. and M.E. Mackay, J. Rheol., 37 (1993) 149.
39. De Kee, D. and C.F. Chan Man Fong, J. Non-Newtonian Fluid Mech.,
62 (1996) 307.
40. Gerhardt, L.J. and C.W. Manke, J. Rheol., 38 (1994) 1227.
41. Manke, C.W. and M.C. Williams, J. Rheol., 36 (1992) 1261.
42. Hua, C.C. and J.D. Schieber, J. Non-Newtonian Fluid Mech., 56 (1995)
307.
43. Ahn, K.H. and K. Osaki, J. Non-Newtonian Fluid Mech., 55 (1994) 215.
44. Tanaka, F. and S.F. Edwards, Macromolecules, 25 (1992) 1516.
45. Wang, S.Q., Macromolecules, 25 (1992) 7003.
46. van der Brule, B.H.A.A. and Hoogerbrugge, P.J., J. Non-Newtonian
Fluid Mech., 60 (1995) 303.
427

47. Vrahopoulou, E.P. and A.J. Mc Hugh, J. Rheol., 31 (1987) 371.


48. Marrucci, G., S. Bhargava, and S.L. Cooper, Macromolecules, 26
(1993) 6483.
49. Wagner, M.H. and J. Schaeffer, Rheol. Acta, 31 (1992) 22.
50. Ait-Kadi, A., Ph.D. Thesis, University of Montreal, Quebec (1985).
51. Xu, Y.Z., D. De Kee, and C.F. Chan Man Fong, J. Appl. Polym. Sci.,
55 (1995) 779.
52. Xu, Y.Z., C.F. Chan Man Fong, and D. De Kee, J. Appl. Polym. Sci.,
59 (1996) 1099.
53. Lodge, A.S., Body Tensor Fields in Continuum Mechanics, Academic
Press, New York, NY (1974).
54. De Kee, D., Recent Developments in Structured Continua, edited by D.
De Kee and P.N. Kaloni, Longman Scientific and Technical, Harlow,
U.K. (1986) 150.
55. De Kee, D. and C.F. Chan Man Fong, Polym. Eng. & Sci., 34 (1994)
438; 35 (1995) 1031.
56. Hinch, E.J., J. Non-Newtonian Fluid Mech., 54 (1994) 209.
57. Macdonald, I.F. and P.J. Carreau, J. Rheol., 33 (1989) 367.
58. des Cloizeaux, J., Europhys. Lett., 5 (1988) 437; 6 (1988) 475.
59. des Cloizeaux, J., Macromolecules, 23 (1990) 4678.
60. Mead, D.W., J. Rheol., 40 (1996) 633.
429

CONSTITUTIVE BEHAVIOR MODELING AND FRACTIONAL


DERIVATIVES

Chr. Friedrich ~, H. Schiessel b'c and A. Blumen b

"Freiburg Materials' Research Center, Freiburg University, Stefan-Meier-Str.


21, 79104 Freiburg, Germany
bTheoretical Polymer Physics, Freiburg University, Rheinstr. 12, 79104
Freiburg, Germany

~Materials Research Laboratory, University of California, Santa Barbara, CA


93106, USA

1. I N T R O D U C T I O N

The simplest decay behaviors are exponential, such as the dielectric relaxation
associated with Debye and the mechanical relaxation named after Maxwell.
Exponential decays depend on a single mode (or, equivalently, a single
characteristic time). But most relaxation processes are governed by a large variety
of characteristic times, see references [1-4] for reviews, and vast types of decay
patterns follow, most popular being stretched exponentials (Kohlrausch-Williams-
Watts [2,5,6]) and power law behaviors. In this review we focus on the cases in
which the decay function follows a power law for reasonably extended time or
frequency intervals. Note that too short time or frequency windows do not allow
to distinguish between different decay patterns [7].
The transition from the glassy relaxation zone to the transitional zone in the
case of stress relaxation of a glassy polymer is an example for power-law
relaxation. Consider polyisobutylene at the reference temperature To = 25~ In
Figure l(a) we reproduce, following reference [8] its shear storage and loss
moduli G' and G" as a ftmction of the frequency o~; in Figure 1(b) we display the
corresponding shear relaxation modulus G and the shear creep compliance J as a
function of time. As is evident by inspection, the high modulus plateau is
followed by a power law that covers about four decades in time (Figure 1(b)) and
frequency (Figure l(a)). Then a second plateau zone (called the entanglement
430

10 10 -4

G' -5

.~ 8~ ~. 8 -6 "7t~

O ~- 7 -7

o 6 -8 e~
@

O
o 5 -9 o

i i i i l i i
-10
-4 0 4 8 -12 -8 -4 0 4

log co/s-1 l o g t/s

Figure l(a). Storage modulus Figure l(b). Relaxation function G(t)


G'(co) and loss modulus G"(co) and creep function J(t) for polyiso-
for polyisobutylene as a function of butylene (as in figure 1(a)) as a func-
frequency. The data are from tion of time.
Tobolsky and Catsiff [8].

~.~ 5

~o
3

2
* .ce, G~- PB302
O
,-- 1 ~ G'- PB304

l 1 1 I I I I I I

-5 -4 -3 -2 -1 0 1 2 3 4 5

log (aTCO) / rad s "l

Figure 2. Storage modulus G' and loss modulus G" of unmodified (PB300) and
urazole modified polybutadiene (PB302 and PB304) vs. the reduced frequency
aTa~. The molecular weight of all samples is Mw - 31 kg/mol; the samples 302
and 304 correspond to the 2 mol % and to the 4 mol % modification, respectively.
431

plateau) shows up. This transition from one plateau to another, generally via a
power law, is characteristic for amorphous polymers.
Furthermore, power law behaviors also appear in the terminal relaxation zone
of polymers. In the case of a single relaxation time one has a sharp transition from
the entanglemem plateau to the flow zone, which obeys a typical liquid-like
behavior, namely G'oc co2 and G"oc co. Physically or chemically cross-linked
polymers [9], polymers with star-, H-, or comb-like topologies (see e.g. [10-19])
show a more general pattern, namely an intermediate power-law domain with
G' oc co~ and G" oc coa , where cz and 13 lie between zero and one. In Figure 2 we
contrast the behavior of neat, monodisperse polybutadiene to that obtained by
attaching to its backbone active groups, which are able to form H-bonds. While
for neat polybutadiene the shear storage and loss moduli obey G'oc o92 and
G" oc co, the moduli of the modified polymers follow more general power laws.
The reason for this behavior is that the active groups form temporary, random
links between the polymers, fact which renders the relaxation process multimodal
and cooperative; this leads here to power laws. Such laws are not the hallmark of
polymers only; vast classes of substances, which range from inorganic glasses to
proteins show such behaviors [1], and we like to recall the early works of
Meinardi et al. [20-22] on power law relaxation in metals, in rocks and in glasses.
In this chapter we focus on the possibility to portray such complex viscoelastic
features of polymers by means of fractional calculus, a formalism which turns out
to be exceedingly well-suited for this purpose. To this effect we start here by
illustrating, using a simple example, how fractional calculus comes into play as a
result of the superposition principle. We start with G(t), the shear relaxation
modulus of a linear system. Now G(t) is defined as the response of the shear
stress r(t) to a jump in the shear strata y(t)-yoO(t) where O(t) is the unit step
function. We assume that G(t) obeys a power law, i.e.

G(t) - r ( 1 - p) (1)

with 0 _ 13< 1. In equation (1) E and ~, are constants and I-'(x) denotes the
Gamma function [23]; for convenience we chose in equation (1) the prefactors in
such a way as to conform to the mare body of the Chapter. Due to the lmearity of
the system, the response of the stress to a previous history of deformations y(t)
is given by the superposition integral [24,25]"
432

r(t) - i dr' G(t- t') dy(t'___)) (2)


dr'
Inserting G(t) given by equation (1) into equation (2) we have

t
Eft ~dt' (t- t') -p dy(t') (3)
r(t)- r ( 1 - p ) at'

Now equation (3) can be rewritten in the following compact form

r ( t ) - E 2 p dP?'(t) (4)
dtp '

in which ~/dt ~ denotes the fractional derivative of order 13 [26,27] (see section 2
for details). Equation (4) with 0 <13 < 1 interpolates between Hooke's law
describing solid behavior (13 - 0), i.e.

r(t)- E?'(t) (5)

and Newton's law describing fluid behavior ( / 3 - 1)

dr(t)
r(t)- r/---~ (6)

Equation (4) is an example of a rheological constitutive equation (RCE) with


fractional derivatives. As we proceed to show below, a whole series of complex
behavior patterns (including crossover situations as discussed above) can be
described by relatively simple fractional RCEs.
This Chapter is organized as follows In the next section we give a brief
introduction to fractional calculus. After surveying the historical development of
fractional RCEs in section 3, we discuss in section 4 the representation of such
RCEs by mechanical analogues. In section 5 we highlight the usefidness of the
formalism by applying the fractional Maxwell and Kelvin-Voigt models to a
variety of polymeric systems. In section 6 we discuss more complex fractional
models. Finally, we conclude with a summary in section 7 and relegate some
important but more mathematical expressions to the Appendix.
433

2. FRACTIONAL D I F F E R E N T I A T I O N AND INTEGRATION

The straightforward extension of classical calculus to its fractional counterpart


is most readily visualized by using a notation which unifies ordinary integration
and differentiation:

f(~)(t) , for a = 1,2,3,..., (7)


d~f(t)
- f(t), for a =O,
at a
dt_~f(t_~) , for a =-1,-2,-3,...

together with the Weyl integral [27]:

d~f(t) _ 1 ~ f(t')
dt ~ F ( L a ) dt'(t-t,)~+l
(8)

One has only to realize that for a = - 1 , - 2 , - 3 , . . , equation (8) is nothing but
Cauchy's formula for repeated integration [26,27]; hence for these values of
equations (7) and (8) are equivalent. Now, the basic idea is that equation (8) can
be readily extended to all tx < 0; this defines fractional integration. The
extrapolation to the positive a-range, a > 0, is obtained by first picking an
integer n, n > ct, then performing a fractional integration of order ct - n, followed
by an ordinary differentiation of order n, i.e. [26,27]

d~ f (t) = d__"_( d~-" f (t))


(9)
dt ~ dt" k, ~-" J

Equation (9) defines fractional differentiation.


Let us note that also another version of fractional calculus, the so-called
Riemann-Liouville (RL)formalism [26,27], is of widespread use in rheology. In
RL the lower limit of the integrals in equations (7) and (8) is set to 0. The RL
formalism is particularly suitable for studying the transient response of materials
after switching an external perturbation on, say at t = 0, so that v(t)= y ( t ) - 0
for t < 0. In this spirit the RL version can be understood as being the restriction
of the Weyl formalism to a special class of initial value problems. In the following
we will use the Weyl formalism; the translation into the RL version is
straightforward.
434

Weyl's fractional calculus rams out to be algebraically very conveniem: the


composition role for differentiation and integration obeys the simple form

d ~ d~f dV+~f
= (10)
dt " dt ~ dt "+~

for arbitrary IXand v [27]. Furthermore, the Fourier transform


oo

{f(t)} - f(o)) = I dt f(t)exp(-icot) (11)


.-oo

turns the operation d~/d# into a simple multiplication [26,27]

{ d~f(t) } : (ico) ~f(co) (12)


dt ~

Let us illustrate the usefulness of these properties using the above-mentioned


rheological example. A quick comparision of equation (3) with the Weyl integral,
equation (8), leads immediately to the fractional stress-strata relation

r(t)=EAp d ~-' dy(t) (13)


dt a-~ dt

Furthermore using the composition rule, equation (10) with r = r - 1 and v = 1,


equation (13) turns into equation (4), as stated in the previous section.
The behavior of fractional derivatives under Fourier transformation is
especially useful in determining the dynamical response functions. Let us consider
the complex shear modulus

(14)
G* ( co) -icol;dzG(z)exp(-icoz )

which describes the response of the stress to a harmonic strain excitation


y(t) =Yo exp(icot). From equation (2) it follows by the change of variables
z - t - t' that 7(co) = G*(co) y(co). Using the multiplication rule, equation (12),
one finds, say from equation (4)
435

G*(co) - 7(0)) / ~(co) - E(ico,;t)p . (15)

From the complex modulus G*(co) follow the storage and the loss moduli,
G'(co) = Re(G*(co) ) and G"(co)= Im(G*(co) ), the complex shear compliance
J*(co)=l/G*(co) as well as storage and the loss compliances, J'(co)
= Re(J*(co)) and J"(co) = -Im(J*(co)). Furthermore, we also consider the
shear creep compliance J(t) (the response of the strain to a stress jump
r(t) = r 0 tg(t)), which is given here by

J(t)-/-'(1+ fl)
As a direct consequence of the multiplication relation, equation (12), we can
easily derive the harmonic response functions G*(co) and J*(co) of a given
fractional RCE. On the other hand, the analytical evaluation of the step response
functions, namely of G(t) and J(t), turns out to be a hard task in many cases.
Nevertheless these responses can be derived explicitly for a whole series of
fractional RCEs of practical importance, cf. sections 5 and 6.

3. HISTORICAL SURVEY OF RCEs WITH FRACTIONAL DERIVA-


TIVES

To our knowledge, the mathematically sound use of fractional differentiation to


describe rheological properties of materials starts with Gemant [28,29]. He
modified the Maxwell model by introducing the semiderivative of the stress (i.e.
a = 1 and fl = 1/2 in equation (30), vide infra) in order to portray the properties
of an 'elasto-viscous' fluid under oscillatory excitations. Nutting, on the other
hand, pioneered power laws such as equation (1) to depict experimental results
[30,31], although at that time the relation between power laws and fractional
derivatives was not clear to the materials' science community. Other examples for
the use of power laws are the works by P. Kobeko, E. Kuvshinskij and G.
Gurevitch [32] (an expression used by them is equivalent to the Cole-Davidson
function of dielectric relaxation) and by Alexandrov and Lazurkin [33]. There are
even claims that stretched exponentials turn into power laws for exponents
smaller than 0.4 [34], see, however also reference [7].
436

In rheology Scott-Blair et al. [35-38] made an extensive use of fractional


integrodifferentiation to depict through power-laws the creep and relaxation in
wide classes of materials. Their works rendered clear the intimate relationship
between power laws and fractional calculus, and also introduced fractional
generalizations of Newton's and Hooke's models, in which the fractional elements
(FE) of this chapter were viewed as arising from 'quasi-properties', representing
non-equilibrium states. Thus, in their notation, property X obeys generally an
expression of the following type:

d'~r
X_ (17)

Viewing z as stress and ~/as strain, the property X is an extension of the usual
definition of viscosity, for which a = f l - 1 . Bosworth [39] made first
considerations concerning the use of equation (17).
After these basic pioneering works, rheology experienced a renewed surge of
activity on fractional calculus starting at the end of the '60ies. Thus Slonimsky
[40] applied the calculus to study rheological phenomena of polymers, a
materials' class of growing importance. He described by a fractional relation the
force acting on a polymer segment and the displacement experienced by it; in fact
the operator used by him can be represented as an infinite series of simple
fractional derivatives. Then Smit and deVries [41] investigated several material
functions, such as the complex shear modulus of the fractional Kelvin-Voigt
model (having a fractional derivative of the strain) and compared the results with
experimental data. Next, Memardi and Caputo [20-22] developed the ideas of
fractional calculus ft~her; they provided expressions for several material
functions for the fractional standard solid model (see below), for which the order
of fractional differentiation was the same for the stress and the strain. The
evaluation of the relaxation in this case was an important step towards the
understanding of the basic properties of the whole class of fractional models.
These results were not widely noticed, and later rederived in rheology by
Friedrich [11-13] and Nonnenmacher [42-44], who also solved the general case,
in which the two fractional derivatives are of different order. Memardi and
Caputo used their expressions [20-22] to describe relaxation measurements of
rock materials, metals and glasses.
A study of the fractional versions of the Maxwell and Kelvin-Voigt models,
where only the derivative of the stress was replaced by a fractional counterpart
was performed by Koeller [45]. He also considered the generalization to parallel
437

and serial arrangements, and also pointed out the close connection between the
Rabotnov calculus [46] and fractional integrodifferentiation. The Rabotnov
calculus, developed in the USSR in the '60ies, is based on the Rabotnov-operator
9t5~(f) (also called fractional-exponential operator), and was used widely for
portraying relaxation phenomena in solid mechanics [46]. Friedrich and Hazanov
poined out the relation between 9t-~(f) and fractional integrodifferentiation
[47]. One has namely:

d k(~ d~-lf (18)


~ - ~ ( f ) = ~k=l( - f l ) * - ' -')f
dtk(a-1)
and ~ft~- a ( f ) - dtg- 1

w h e r e 0 _ a < l andfl>_0.
More recently, one of the goals of research in our field was the detailed
analysis of the fractional Maxwell and Kelvin-Voigt models. It was soon clear
that the analytical determination of the response functions of these models is a
mathematically difficult task (see below). Bagley and Torvik [48] connected the
molecular theory of viscoelasticity to the fractional calculus, and showed that
some aspects of the theory are mirrored by the fractional Kelvin-Voigt model.
Friedrich and Heymann [15,16] pointed out the intimate relation between the
order of differentiation in a fractional Kelvin-Voigt model and the degree of
conversion for the sol-gel transition of a crosslinking polydimethylsiloxane.
Later, in 1986, Bagley and Torvik [49] considered the fractional standard solid
model in more detail. They established that this model, which contains two
fractional derivatives- one of the stress and one of the strata- is compatible with
thermodynamics if the order of both derivatives is identical. Thermodynamical
admissibility of fractional order models was also analyzed by Friedrich for
several models [12].
In 1983, Rogers [50] formulated general RCEs containing a large number of
sums or products of fractional derivatives, both of the stress and also of the strain.
The analysis of these models was restricted to those containing mainly two
fractional derivatives, and dealt with (the more accessible) frequency domain.
Another use of fractional derivatives was put forward by VanArsdale who
generalized the Rivlin-Ericksen and White-Metzner tensors by including
fractional orders of differentiation [51 ].
Only in 1991 did Friedrich [11] as well as G1Ockle and Nonnenmacher [42]
succeed in obtaining G(t) and J(t) for the fractional Maxwell model containing
two fractional derivatives of different order. They showed that this solution can
be expressed through a special class of mathematical fimctions which will be
438

discussed later in this Chapter. The derivation in references [44,52] is based on


an integral fractional representation of the standard solid model, whereas
Friedrich's derivation uses the differential picture. Friedrich pointed out that the
differential fractional version is thermodynamically admissible for wide ranges of
parameters, whereas the parameter range of the fractional integral version is very
restricted [17]. In subsequent works GlOckle and Nonnenmaeher clarified the
mathematical basis of fractional calculus by pointing out its relation to the so-
called Fox-functions [42]. Moreover, they explained the interrelation between
fractional relaxation functions and the time-temperature superposition principle
[52]. We note that these models were successfully applied to the description of
the viscoelastic properties of filled polymers [53].
In general, much of the very recent work is characterized by the search for the
physical background underlying fractional calculus. The relation of this calculus
to the classical sprmg-dashpot representation of complex viscoelastic materials is
a central aspect of the works by Schiessel and Blumen [54-58], as well as by
Heymans and Bauwens [59-61]. These works showed how power law relaxation
follows from exemplary arrangements of springs and dashpots. Schiessel and
Blumen succeeded in explaining the process of polymer cross-linking on the basis
of such mechanical networks; they also investigated terminated ladder
arrangements which mimic pre- and postgel behavior [56]. Such mechanical
analogs also showed the way how physically reasonable fractional RCEs can be
constructed, aspects discussed by Schiessel, Metzler, Blumen and Nonnenmacher
[58] and by Heymans [61]. By employing a formal analogy between linear
viscoelasticity and diffusion in a disordered structure Giona, Cerbelli and Roman
derived a fractional equation describing relaxation phenomena in complex
viscoelastic materials. This analogy leads to a power law expression, which is in
agreement with the experimental results [62]. At the moment we are still far from
being able to relate in a reductionistic way the empirical models to a microscopic
background. Nevertheless, as will become evident in section 4 of this Chapter,
the representation of fractional derivatives through sprmg-dashpot analogs is
helpful in understanding a series of phenomena which underlie the fractional
calculus formalism.
Another line of research, widely pursued nowadays, is the phenomenologically
oriented, pragmatic modeling. An example for this is the work by Stastna,
Zanzotto and Ho [63] who provided a relation between the Kobeko ftmction [32]
and fractional calculus and used it to model rheological data of asphalts. Stastna
et al. found a general inversion scheme which generalizes Rogers' results [50]; it
allows the representation of the complex shear modulus in the form:
439

.oo I + o 1,tI (19)

In equation (19) ~lk and ~2k are two sets of characteristic times. The authors of
reference [63] succeeded in deriving the corresponding fractional RCE. Note that
for m = 0 and n = 1 one recovers the Kobeko function [24].
We close this section by noting that nowadays fractional calculus is of
widespread use in describing different rheological phenomena for wide classes of
materials. As examples from the polymer literature we refer to [19,53,64-67]. In
the following section we analyze under which circumstances spring and dashpot
arrangements lead to power law relaxations.

4. THE REPRESENTATION OF RCE WITH FRACTIONAL DERI-


VATIVES BY MECHANICAL MODELS

Fractional RCEs may be formally derived from ordinary RCEs by replacing the
first-order derivatives (d/dt) by fractional derivatives (~/dt ~) of non-integer
order (0 < 13< 1). This formal procedure can, however, not assure a priori that the
resulting expressions are always physically reasonable. This aspect was pointed
out, for instance, in references [11,42,54]. Thus it is useful to have a procedure at
hand that automatically guarantees mechanical and thermodynamical stability.
As a first step Schiessel and Blumen [54-57] and Heymans and Bauwens [60]
have demonstrated that the fractional relation, equation (4), can be realized
physically through hierarchical arrangements of springs and dashpots, such as
ladders, trees or fractal structures (which we will discuss in this section). The
idea is that (disregarding for the moment the specific structure of the hierarchical
constructions, these will be detailed later) equation (4) is obeyed by a fractional
element (FE) which is specified by the triple (fl, E, 3,). We symbolize the FE by a
triangle, as shown it Figure 3(c), where also its classical counterparts are
depicted: the spring (cf. Figure 3(a)) obeying equation (5) and the dashpot (cf.
Figure 3(b)) with the stress-strain relationship equation (6). Then, as a second
step, more complicated RCEs can be constructed by combining two or more FEs
in serial, parallel or more complex arrangements; this was proposed by Schiessel
et al. [58] and by Heymans [61]. We will make use of this method in sections 5
and 6.
440

E __Lq
(a) (b) (c)
Figure 3. Single elements: (a) elastic, (b) viscous and (c) fractional.

Eo
E1
rio
E2
Eg/
r12

1//

Figure 4. A sequential spring-dashpot realization of the fractional element.

Let us now consider different realizations of FEs. In reference [54] Schiessel


and B lumen proposed a ladder-like structure with springs (having spring
constants E0, El, E2,...) along one of the struts and dashpots (with viscosities
r/0, ql, q2,-.-) on the rtmgs of the ladder (cf. Figure 4). As shown in reference
[54] the complex modulus G*(co) admits a continued fraction expression, namely

G * ( c o ) - Eo (it~ -1~~ (it~ -~'~ ~(it~ -~


~..., (20)
1+ 1+ 1+ 1+
441

where we use a standard notation for continued fractions, [a/(b + ) ] f - a/(b + f ) ,


cf. reference (23). Choosing in equation (20) E o = E 1 - . . = E and
r/0 - rh - . . . - r/it can be shown (by comparing terminating approximations of the
continued fraction with the binomial series) that the complex modulus of the
infinite arrangement is given by

G*(co) - E (4(ic~ + 1)1/2 - 1 (21)


2(io92) -1

where we set )~-rl/E. For a)2 <<1 equation (21) reduces to the form
G*(co) "~ E(ico2) ~/2 . Therefore having the same spring constants and viscosities
for the whole arrangement one gets a complex modulus with 13= 1/2, and thus the
ladder model with equal springs and dashpots leads to a FE with the parameters
(1/2, E, A). To extend the ladder model to arbitrary values of 13 with 0 < 13< 1, a
suitable distribution for the spring constants and viscosities has to be chosen; thus
an algebraic k-dependence of the material constants of the form

E k oc Tlk oCk 1-213 (22)

leads to a FE specified by (fl, E,A) [541.

r
(5 2n G :n+lI
....

' (
6 r
n

Figure 5. Recursive construction of the iterated mechanical networks of Heymans


and Bauwens [60].
442

Heyrnans and Bauwens [60] and Schiessel and Blumen [56] demonstrated that
values of 13 with 13r 1/2 can also be attained by arranging equal springs and
dashpots into a network with a more complex structure. Thus reference [60]
focuses on a class of iterated networks which are generalizations of a fractal tree
model; each FE (/3,E,2) can be represented by such an arrangement to any
desired accuracy. We show in Figure 5 the recursive construction of the model:
One starts from an element having the complex modulus G~; this element gets
then represented by four elements, two of which have G~, the two others G2 and
G 3 as moduli (of. Figure 5 with n = 1 ) so that the ensuing structure still has G~
as complex modulus. From Figure 5 it is easily seen that the relation between the
complex moduli of the elements has to be G,, = ~G~,, Gz,+~. In principle the
iteration proceeds for each element indefinitely. By this the complex modulus G~
of the whole arrangement is

(23)
c, - - (c4G, G6c ) ''4 - ..

At each level n of decomposition one replaces some of the elements G,


(k = 2",...,2 "+~- 1 ) by either purely elastic elements, i.e. springs with spring
constant E, or purely viscous elements, i.e. dashpots with viscosity rl. For these
elements one has to replace Gk in equation (23) by E or icor/ and stop their
further decomposition. Let iV, denote the number of springs at the nth level of
decomposition and M, the corresponding number of dashpots. From equation
(23) it follows:

c, - E --
(24)

with /3- ~ " M,,/2" and 2 = ~7/E. Thus one can reach any preset, arbitrary 13,
0 < 13 < 1 to any desired accuracy by a sufficiently fine decomposition. As a
special case, by taking G 2 = E and G3 = i corl we obtain a binary tree. This
network, which we depict in Figure 6, is again a FE with the parameters
(1/2,E,2).
The advantage of the "binary decomposition" of elastic and viscous elements is
its simple analytical tractability; in many cases, however, it leads to complicated
networks whose properties are difficult to grasp intuitively. Thus it is also
443

desirable to work with networks whose structural properties (such as the


connectivity) and the exponent 13are related in a more obvious fashion.

1 ', 1 1 1
I ', ', I I I I I

Figure 6. A special case of the iterated networks shown in Figure 5" binary
spring-dashpot tree.

In reference [56] Schiessel and Blumen showed that for given fractal networks
this relation is straightforward when the springs and dashpots are suitably chosen.
The construction starts by connecting each site r~ of the given network to
neighboring nodes rj by equal springs (with spring constant E) and linking each
r~ to the ground via a site-dependent dashpot (with viscosity r/~ - z(r~)r/, where
z(r i) is the coordination number of node r~ ). Furthermore, the nodes' motion is
perpendicular to the ground. To give an example we show in Figure 7 a section of
the infinite mechanical network constructed from the Sierpinski gasket. The
analogy to random walks can now be seen by comparing the stresses acting on
node r i (whose displacement is ~,~)

r/~ ~'~ (t) - E ~-"~ [7 j(t) - 7 ~(t)] (25)

with the master equation

dt'(r ,t) (26)


444

which governs the probability P(r~,t) of having a random walker at site r~. In
Eqs. (25) and (26) the sums run over all nearest neighbors rj of r~. The transition
probabilities w o. in Eq. (26) obey z(rj)w o - w - constant. Now one can identify
formally r/, y,(t) with P(r~,t). Furthermore, the probability for a random walker
to return to the origin at time t follows an algebraic form:

P(t) oc t -d' '2 (27)

where d~ is the spectral dimension of the network [68]. A power law behavior of
the complex modulus, Eq. (15), follows, with ~ 3 - 1 - d ~ / 2 (cf. Ref. [56] for
details). The ladder model, Figure 2, is the special case of a one-dimensional
lattice with d s - 1 and thus 13- 1/2; the Sierpinski gasket in 2d (cf. Figure 7) has
d,. - 2 ln(3)/ln(5) and hence fl - 1 -ln(3)/ln(5) "- 0.317.

Figure 7. Section of the infinite spring-dashpot arrangement based on the


Sierpinki gasket.

An interpretation of the sol-gel transition in terms of such mechanical networks


can be fotmd in Ref. [56}.
445

5. FRACTIONAL M A X W E L L AND KELVIN-VOIGT MODELS

Figure 8(a) shows the standard Maxwell model, in which a spring and a
dashpot are arranged in series [24,25]. Now this model can be generalized by
replacing these elements by the FEs (a, E 1, 3,1) and (/3, E 2, 22), as shown in
Figure 8(b). Due to the sequential construction the stress v(t) is the same for both
elements. Their respective stress-strain relations are

7 , ( t ) - E~'2-~ ~ d - ~ z-(t) (28)


tit_---------~ , Y2(t)-E21~ d-#z-(
t)dt_~
where both expressions follow from the composition rule, Eq. (10). Due to the
sequential construction of the fractional Maxwell model (FMM), we have
7(t) - 71(t)+ 72 (t), from which it follows that

r(t) + E13;~ d ~-p v(t) d~


E2/~f122 dt a-fl __ E1/~al dt~r(t) (29)

Let us assume without loss of generality that ~z > ]3. Eq. (29) can be further
simplified by setting 2-(E~A~/Ez2~) v('~-~) and E - E~(A1/A) ~ . This leads to

z'(t) + A ~ d~-#z-(t) = EA a d~y(t) (30)


dt.-P dt ~

which is the RCE of the FMM [11,43,58].

q ,E1,21)

(a) (b)
Figure 8. The (a) ordinary and (b) the fractional Maxwell model
446

When one arranges the spring and the dashpot in parallel one obtains the
standard Kelvin-Voigt model [24,25], depicted in Figure 9(a). Its generalization,
which has two FEs, is shown in Figure 9(b). Because of the parallel arrangement,
the stresses acting on the two elements are. additive. Following a procedure
similar to that for the generalized Maxwell model (vide supra), one finds for the
RCE of the generalized Kelvm-Voigt model (FKVM) [58]

r(t) - E,t ~ d~y(t) +mr p dPy(t) (31)


dt ~ dtp

where we have used the same parameters X and E as for the FMM.

I I

(a) I (b) I
Figure 9. (a) The Kelvin-Voigt model and (b) its fractional generalization.

The basic material functions of the FMM and the FKVM were derived
elsewere [58]; we will restrict ourselves to their succinct display and to the
discussion of their basic properties. Table 1 presents for both models their
fundamental response functions.
In Table 1 E p(z) denotes the generalized Mittag-Leffier function of argument
z, whose definition and basic properties are given in the Appendix. This function
is of great importance for fractional RCEs, since it occurs in many material
functions in the time domain. The harmonic response functions, equations (36) to
(39), follow simply from the RCEs, equations (30) and (31), by means of the
multiplication rule, equation (12). Due to the sequential (parallel) construction of
the FEs in the FMM (FKVM), the response J(t) (G(t)) is simply the sum of the
corresponding responses of its structural parts, which obey equation (16)
(equation (1)).
447

Table 1
Material functions of the fractional Maxwell and Kelvin-Voigt models

FMM FKVM

c(t) E(t/'~')-flEa-fl,l-fl(-(t//],)a-fl) (32) F(1 a) F(1 - fl)

E-1 (_~_)a .~-1 (~)e (34) (35)


J(t) V(1 + a) + V(1 + fl~--~

G*(co) E (i c~ E(icoZ) '~ + E(ico2) p (37)


1 + (ia)2) ~-~ (36)

J*(co) E-' (ico2) -'~ + E < (ico2) -p (38) E-1 (39)


1+

Using the asymptotic expansions of the Mittag-Leffier function and focussing


on the behavior of the dynamic moduli and of the dynamic compliances it
becomes clear that both the FMM and the FKVM describe two power law
regions which intersect each other at the characteristic time t = 2 . In the case of
the relaxation ftmction G(t) the FMM describes first a flat decrease (of slope
- f l ) followed by a steeper decrease (of slope - a ), whereas for the FKVM the
flat decrease follows the steeper decrease. The compliance of the FMM shows a
flat increase with slope 13 followed by a steeper increase with slope a; for the
FKVM the situation is reversed.
Thus it follows that in the time domain the FMM can be used to describe
relaxation functions which start from a plateau region and decrease according to
power laws, whereas the FKVM is able to describe the relaxation via a power
law to a plateau. The situation is correspondingly reversed in the frequency
domain. Note, however, that a so-called S-shaped transition between two
plateaus which are connected by a power law cannot be described by means of
FMM or of FKVM.
An example for the use of FMM is presented in Figure 10, in which the
dynamic moduli of polyisobutylene are redrawn from Figure l(a). These are
compared to the results of FMM with the parameters given in Table 2. Note the
448

10r /

[]

~-~ FMM
~ 8
~ ~.

41 ,
-3 0 3 6 9

log co / s-1
Figure 10. Description of the data of Figure 1(a) by the fractional Maxwell model
(solid lines) and the fractional Kelvin-Voigt model (dashed lines).

Table 2
Material parameters used in Figure 10

log E log L c~ 13
FMM 9.09 -8.36 0.583 0.593
FKVM 5.75 -2.96 0.057 0.885

good agreement between the experimental data and the analytical expressions for
FMM. On the other hand FKVM is unable to describe the observed features even
qualitatively. This is especially true for the G"(o~) data where FKVM can not
follow the decline found experimentally in the range of very low frequencies.
In Figure 11 (which is redrawn from Figure 2) we provide another example
which stresses the possibilities offered by FMM to depict relaxation data.
Whereas Figure 1 is concerned with polymers in the glassy and transitional zones,
here we deal with relaxation from the plateau towards the terminal relaxation
zone. Figure 11 shows the storage moduli, both of two modified polybutadienes
(PB302 and PB304) and also of the neat polymer, on which the modifications
were performed (PB300) [64]. Note that the FMM with the parameters given in
Table 3 describes in a quantitative way the observed relaxation- the focus lies
449

here on the power law behavior of the modified PBs at very low frequencies. We
note furthermore that both the storage and the loss module of the neat polymer
melt are quantitatively depicted by this model.

r FMM

Q~ ~b" o G'- PB300


J" * G" - PB300
O o G' - PB302
G'- PB304
0 I I .... I I I I I I l

-5 -4 -3 -2 -1 0 1 2 3 4

log (aTO) / rad S"1

Figure 11. Description of the data of Figure 2 by the fractional Maxwell model
(solid lines).
Table 3
Material parameters used in Figure 11
. . . . . .

logE log)~ ~ 13
PB300 5.52 -1.87 0.882 0.994
PB302 5.60 -0.344 0.553 0.632
PB304 5.48 0.720 0.478 0.590

The same holds true for polystyrene, as shown in the following example, taken
from reference [65]. The situation is depicted in Figure 12; as is obvious, the
storage moduli of polystyrene filled with different amounts of silicagel (N20)
follow a power law in an intermediate region. Here the slope of the low frequency
450

3
k3 Model
e~o 0.5 and 5 vol. %
o 2
1 and 4 vol. %
2vol. %
PS 100 matrix ~ 3 vol. %
I I I I

-2 -1 0 1 2 3
1og (aTo3) / rad s- 1

Figure 12. Storage modulus G'(a~co)


for aerosil-filled polystyrene. The solid
lines give the description through the fractional Maxwell model. The aerosil
concentration increases from below.

Table 4
Material parameters used in Figure 12
log E log ;~ c~ 13 log G~
PS100 4.97 -0.877 0.812 1.000 -oo
0.5 vol. % 5.07 -1.06 0.808 0.988 1.54
1 vol. % 5.17 -1.08 0.749 0.907 2.62
2vol. % 5.26 -1.12 0.718 0.867 3.09
3 vol. % 5.40 -1.03 0.649 0.784 3.99
4 vol. % 5.47 -1.15 0.597 0.725 4.15
5vol. % 5.49 -0.814 0.613 0.768 4.85
451

plateau is zero; consequently the system can be modeled by adding a Hookean


spring (of strength Ge) in parallel to the two FEs. The agreement between this
model and the experimental data is very good. The Ulm group [44,53] used a
similar model (the integral version of the FMM) to describe the complex moduli
of filled polymers.
On the other hand, the example of Figure 12 demonstrates the need to go to
models having more than two FEs in order to describe more complex behavior
patterns, such as S-shaped transitions. Such extensions are the focus of the next
section.

6. MORE C O M P L E X MODELS

6.1 Models based on combinations of 3 fractional elements


Here we present more complex models, which are combinations of three FEs.
We dispense with considering the fully parallel or the fully linear combination of
FEs. We view as much more interesting the parallel arrangement of the FMM
with a third FE or the sequential combination of the fractional Kelvin-Voigt
model with a third FE. We call these models the fractional Zener model (FZM,
also fractional standard solid model) and the fractional Poynting-Thomson model

E3
q
I
(a) I " (b) - I

Figure 13. The (a) ordinary and (b) fractional Zener model.
452

(FPTM), respectively. We stop to note that Heymans [61] has also analyzed more
complex models, which are composed of FEs and additional conventional
elements.
The constitutive equations of FZM and of FPTM (these being our basic,
simplest models containing 3 FEs) can be derived along the lines of the previous
section.
We start with the analysis of the FZM. The mechanical representation of the
FZM is displayed in Figure 13; it consists of an FMM (with the same parameters
as in Figure 8) in parallel to an FE specified by (z, E3,A3). The stresses on the
left and right branches of the arrangement, x L and XR, obey the following stress-
strain-relationships:

v L(t) + 2 ~-p d~-/~r L(t) = E2 ~ d ~y(t) (40)


dt~-P dt ~

of. equation (30) and, equation (4),

rn(t)- E32~ d'y(t) (41)


dt r

Both stresses add, which leads to the RCE of the FZM:

z(t) + 2,~-~ d~-Pr(t) = E2 ~ d~7"(t) + Eo 2r d r y ( t ) + Eo 2y+~-~ dr+"-PY(t) (42)


dt~-P dt ~ dt r dtr+"-P

Here we set E o - E 3(23/2) r


The FPTM consists of a serial arrangement of the FKVM (with the same
parameters as in Figure 9) and an FE (y, E3,//'3) (cf. Figure 14). The RCE of this
arrangement can be derived by noting that the deformations of the Kelvin-Voigt
element and of the FE simply add, which leads to:

E d ~-r v(t) E Ap_r d ~-r r(t) d ~y(t) EA p d~y(t)


v(t) +,, ,,~a-y + = E2,~ ~ + (43)
Eo dt ~-r Eo dt p-y dt ~ dt p

where we again set E o - E3(23/2,) y


453

( y , E3,23)

q ( a , El,21) E2,~)

(a) (b)
Figure 14. The (a) Poynting-Thomson model and (b) its fractional generalization.

Table 5
Material functions of the fractional Zener and the Pointing-Thomson models

FZM FPTM

(v)
go 2C;-2~am~vl(Gat)
G(t) G F ~ ( t ) + G~(t) (44) (45)
E
+~GvE(t), for 3' - [3
E+E o
2
"-'l ,1FKVM
J(t) (46) JFKVM(t) + Jm(t) (47)
+ ~Ej o~ ( t ) , for r -
E+E o
(ic~162 + (ic~ (49)
G*(co) G*Mu(co) + E o (ico2) r (48) E- 1-[- E;l(i(.0,~,)~z_.y --[-E;l(i(_.O/],)fl_y

J*(co) (~)-~'-~+(~'~)-~
E + Eo(i~:k) + Eo(i~X) ~-~ (50) J ~9v M (co) + Eo ~(ico,r -~ (51 )
454

We again summarize (now in Table 5) the relaxation material functions of the


two models. For the detailed derivation of the expressions given in the Table we
refer the reader to [58]. Here we note only that the dynamic moduli as well as the
dynamic compliances follow directly from the corresponding RCEs by means of
the multiplication rule, equation (12). Moreover, the relaxation modulus (creep
compliance) of the FZM (FPTM) is simply the sum of the moduli (compliances)
of the corresponding substructures. On the other hand, the creep compliance
(relaxation modulus) of the FZM (FPTM) is only known analytically for the
special cases y -cx or ~/-13, cf. [58]; in Table 5 the case ~/-13 is presented. The
constants C, and C2 of Table 5 are C,-(Eo/(E+Eo)) v(~-~) and C z
=(E/(E-J-Eo)) 1/(cr-fl) "
The FZM can describe S-shaped transitions from one plateau to another via a
power law. For instance, it can be shown [58] that G(t) displays three regimes
for0_<7_<fl < a and E>>E0:

t -~ for t<<)~
(52)
G(t) oc t -~ for )~ << t << )~,
t -Y for )~ <<t.

where )~, ~(E/Eo)V(~-~)~. In section 5 we have already presented an example


where the FZM fits experimental data very well for the special case ~/- 0 (cf.
Figure 12).

6.2 Other models


Another approach which extends the use of FE-based models is pragmatic. The
previous sections dealt with the exact solutions (relaxation functions) of models
created by combining an increasing number of FEs in parallel and/or in series.
The approach which we discuss here is the direct empirical modification (based
on experimental findings) of the basic relaxation functions of the FE models. We
now provide examples taken from the literature for such modifications.

6.2.1 Modification of the relaxation function


As already mentioned (see Figure 2)the G'(o3) and G"(o3) of monodisperse
linear polymers show a sharp transition from a G' oc o) 2, G" oc co behavior to a
455

G'oc co~, G"oc co-~ behavior. In the example of Figure 2 13 "~ 1/4 and a is
positive and close to zero, hence G' displays a quasi-plateau. In the
corresponding relaxation experiment the stress decays then first according to a
power law with slope - a , which is followed by an exponential relaxation at
longer times. Figure 15 shows the relaxation function G(t) of a modified
polybutadiene [64] where the appearance of an intermediate power-law domain is
clearly evident. In Figure 15, due to experimental limitations, the plateau can not
be seen directly but it can be infered from its dynamic modulus (especially from
the range aTco > 10 ), displayed in Figure 16. Now, fluid behavior is associated
with the termination of all relaxation processes, say at the time )~m, and thus )~m is
a natural cut-off. This can be achieved phenomenologically by multiplying the
corresponding decay function with an exponential. As an example, starting from
GF~(t ) (cf. equation (32)) in reference [18] the following 6-parameter relaxation
function was put forward

c(o- Co ~ exp(-t/~m) . (53)

In equation (53) Go is a constant, 2 and 2~ are time constants (with 2 m >> 2)


and E~,~_~+~ denotes the generalized Mittag-Leffier function (c~ and 13 are related
to the parameters aft ~ and f l ~ of equation (32)via a = a ~ - f l F ~ and
fl = a f t ) . From the asymptotic expansion of the Mittag-Leffier function for
fl < 1 it follows that for t << 9~ (plateau regime) G(t) oc t ~-~ and for )~ << t << )~m
(intermediate regime) G ( t ) o c t -p . Moreover, thermodynamic stability requires
that the power law exponents obey 0 < a _<fl _<1. The cut-off function,
exp(-t/2 m), guarantees the fast release of stresses for t > A m and hence induces
a fluid-type behavior. A detailed discussion of this model is given in references
[10] and [18]. We only provide here the main formulae.
Thus the dynamic moduli associated with equation (53) are

G*( o)-Go + Z/ m) (54)


I nt- (lO)/], "Jr"/],//],m) Ot
and

J*(co) -
(55)
G O io)/~ (i0)/9~ "Jr"2/~l~m) fl-1 "
456

Especially the dynamic moduli turn out to be particularly useful, since they
lead to equations which are easy to use for fits to the data. Figure 16 shows
exemplarily both the data and the obtained fits. For the PB202 sample we obtain
f l - 0.578 and log/~ m = 2.48. This is in good agreement with the parameters
/3'- 0.556 and log2~ - 2 . 2 4 which are determined by fitting equation (53) to
the relaxation function of the sample displayed in Figure 15.
In Figure 15 the glassy short-time behavior (with the plateau modulus Go ) is
not in the experimentally accessible range. In such a case the data may be fitted
by a simpler relaxation function, which one obtains from equation (53) by letting
2 ~ 0. This leads to

(56)
G(t) - Go1. (t/2~) -p exp(-t/2m)

r ~ _ . .

with Go~- rlo/2m, where q0 is the zero-shear viscosity, rio-JoG(t) dt 9


Equation (56) is an empirical relaxation function of widespread use; thus its
dynamic moduli were often used in fitting data (cf. reference [24]).

,~ I PB 202
10s . ~ . .~ Tref= 30 ~

04 _

10 3

102
10l 10 ~ 101 10 2 10 3 10 4
t/s
Figure 15. Relaxation function G(t) of a modified polybutadiene (PB 202, Mw =
21 kg/mol). The solid line is equation (56) with log G0l = 4.68 Pa, log km = 2.24 s
and 13= 0.556.
457

10 6

10 5 Tref= 30 ~

r 10 4

~ 10 3 9 PB200
PB202
o PB204
10 2 Model

101 , , ,,,,,,1 , , ,,i,,,I


d , , ,lll,ll ,L I III I'll I i iiilul ! i lllllll , I I IIIll

1@3 10-2 10-1 10 0 101 10 2 10 3 10 4


a T O / r a d s -1

Figure 16. Storage modulus Gt(aT(O) for unmodified (PB200)and modified


(PB202 and PB204) polybutadiene. The solid lines represent the fit of the
generalized fractional Maxwell model according to equation (54) to the data.

As another example we present in Figure 17 the dynamic moduli of a partly


cross-linked polybutadiene, PB 18, in the pregel stadium [9,69]. PB 18 is a
polymer with a branched structure; it is gratifying to note that also in this case
fractional relaxation provides a very good description of the experimental
observations.
Let us now turn to the RCE associated with the relaxation function, equation
(53). Starting from equation (54) it can be shown by means of the multiplication
rule, equation (12), that the RCE has the following form [10]:

x(t) + Ue -'/x~ d~ [et/X~'c(t)] -G~ e-'/x" dt ~-~d~-[e'/x~


l dy(t) " (57)
dt ~

Equation (57) contains an explicit time dependence that can be removed using the
product rule for fractional differential operators (cf. equation (5.5.2) of [26]). In
this way, the RCE, equation (57) takes the form [10]"
458

oc oo

r(t) + w ~ Z ( ~ ) I I , " d~'rCt----~) - G O w s~ Z(Pf) )t b' dbJT(t----~) (58)


,=o " dt~' - ;=o m dt b,

w i t h a, - c t - i, bj - f l - j and w - 2/2 m .
A RCE with a structure similar to equation (57) was given in reference [63]. A
mechanical model composed of spring, dashpots and fractional elements is not
k n o w n f o r this R C E .

10 6 _

10 5 -

10 4 -

o G'
10 3 -
t~ G t~

10 2 - Model

I I I I I till , i i ~ ,Ittl j , , , I,l,l j ~ J LIHl


10 ]
101 10 2 10 3 10 4 10 5

aTO) / r a d s -1

Figure 17. Storage G'(arco ) and loss modulus G"(arCO) for polybutadiene
(PB18), for details see reference [69]. The solid lines represent the fit of the
generalized fractional Maxwell model according to equation (54) to the data.
459

I I I I I I

0000(

QFIDI3D [

9
/
9
_

PB-M2 o~ o G'
O
2- o
o n G II
I
O

O
oo 1
l I I O I I I I,

-4 -3 -2 -! 0 ] 2 3

log (aT0)) /rad s -1

Figure 18. Storage G'(arCO) and loss modulus G"(arco ) for an unmodified (PB)
and a sidechain modified polybutadiene (PB-M2), for details see reference [70].
The solid lines represent the fit according to equation (59) to the data.

6.2.2 Modification of G*(m) and of associated materialfunctions


Another modification of polybutadienes leads to yet another behavior in the
vicinity of the terminal region; instead of approaching the flow region after
leaving the power law zone, a second power law is observed. We present in
Figure 18 this behavior, which was the research object of reference [70]. Figure
18 shows the dynamic moduli of a polybutadiene which carries sidechains in a
comb-like manner; the sidechains are mesogens, and are attached to the backbone
via a flexible spacer. The polymer is in the isotropic state, but its branched
molecular architecture and the interactions between the mesogens lead to a
dynamic behavior quite different from that of linear chains, as can be seen from
Figure 18.
Here the necessary modification to the material functions starts from the
complex modulus G*(m). The two power law regions at the lowest frequencies
460

are depicted by the FMM; multiplying G'z,~ (co) with an additional relaxation
term leads to:

G*(co)- E (ic~ 1 (59)


1 + (/(_O/~,1)a-fl 1 + (i(__O,~,2)~"

Now the other relaxation functions of the model can be calculated following
the procedures outlined above. The relaxation modulus G(t) and the creep
compliance J(t) turn out to be complex expressions, which can be found in
reference [70]. Figure 18 shows that equation (58) reproduces successfully the
data, as indicated by the fitting curve. We close by giving the RCE belonging to
equation (58) [70]:

d ~r(t) d ~r(t) d a" r(t) dPr(t)


+ Xr2 ~ + 2~2r2 = E2~ (60 /
dt ~ at

Hence the mechanical representation of this model is a combination of 4 FEs in


series; these are of the form (x, E, Z,), with x - 13, 13- ~, 13- 7 and 13- a - 7.

7. CONCLUSIONS AND O U T L O O K

In this chapter we have shown how fractional calculus allows a physically


sound generalization of classical models from the linear theory of viscoelasticity
(see also the review by A. I. Leonov in this book). From a mathematical point of
view, the operation of fractional mtegrodifferentiation is well defmed and can be
easily handled in Fourier or Laplace space. Viewed technically,
integrodifferentiation allows, for instance, to interpolate smoothly between
Hooke's and Newton's laws. Moreover, we have introduced the simplest form of
a fractional rheological model, which we call fractional element (FE), and we
have provided for it several mechanical analogs, namely arrangements made out
of springs and dashpots. In these (infinite) networks the order of fractional
integration or differentiation can be adjusted in several ways; say, by varying the
material constants of the springs and dashpots involved, or by changing the
structure of the arrangement.
Furthermore, we have shown how parallel or serial combinations of FEs lead to
more complex models; in particular we have studied extensions of the Maxwell,
461

the Kelvin-Voigt, the Zener and the Poynting-Thomson models. The relaxation
patterns of these models can be used to fit the experimental results for large
classes of materials. Particularly noticeable candidates are polymeric materials
which display ramified structures (such as cross-linked polymers) or whose
dynamics is characterized by cooperativity (i.e. glasses). Using fractional
elements one can tailor viscoelastic models with given properties, while keeping
the number of parameters involved relatively low.
The representation of generalized viscoelastic models by fractional analogues
also allows a deeper insight into the physics behind fractional stress-strain
relations. Nevertheless, we are still far from a reductionistic understanding of the
fractional relaxation laws. Here, as well as in related areas, we expect much
additional work in the coming years.

ACKNOWLEDGEMENT

The support of the DFG through SFB 428 and of the Fonds der Chemischen
Industrie is gratefully acknowledged.

APPENDIX

Here we display some mathematical relations which are helpful in


understanding the physical models of the Chapter. We start with the generalized
Mittag-Leffier function which occurs, for instance, in the relaxation function of
the fractional Maxwell model (equation(32)) or in the creep compliance of the
fractional Kelvin-Voigt model (equation (35)). The generalized Mittag-Leffier
function E,,~(z) is defined by [71]

z k
, with la > 0 and v > 0. (61)

The special case v - 1 yields the usual Mittag-Leffler function E~,(z). On the
basis of this definition, the generalized Mittag-Leffier functions for some special
cases, in which la and v equal 0.5, 1 or 2, follow readily"
462

Eo.5,1(z ~ - e ~ erfc(_z ~

- - e ~ erfc(-z ~ 1- 2 z
z

El,l ( Z 1 ) - e z
(62)
E1,2(zl) - l [e: - l]
z

E2,1 (z 2 ) - c h ( z )

E2,2 (z2) _ l__sh(z) .


z

All generalized Mittag-Leffier functions increase monotonically for z > 0. To


obtain monotonically decreasing functions one goes to the domain of negative z.
In the parameter range /~, v ~(0,1] the following asymptotic expansions for
z >> 1 are of interest [26]

la z -2 for l a ~ l (63)
V(1 -

1 -1
E,.~(-z) ~: F ( v - l a ) z for v > la.

Consider now the relaxation function G(t) of the FMM, equation (32). Here the
generalized Mittag-Leffier function has the parameters l a - a - 13 and v - 1-13.
From definition (61) and the asymptotic expansion (63) one obtains for the two
power-law regimes:

G ( t ) ~: t -~ for t <<
(64)

G ( t ) ~: {
t -~
t
(a<l)
- 1)
for t >> ~,.

In Figure 19 we show the relaxation function of the FMM, equation (32), for
r - 1 and 13- 0.5. For the sake of comparison two other relaxation functions are
added, which are different from the double power-law relaxation in the long time
range.
463

i i i i i i i i

o- " ~176176

~,-2

~-4
/\\
i
O -6

-8 - (a) x "0"5 e -x / / -
(b) x "0"5 e"xO5
-10 i i i ~ , ~ , i

-4 -3 -2 -1 0 1 2 3 4 5
log x
Figure 19. Dimensionless relaxation function G ( t ) / G o vs dimensionless time
t / 2 for three models corresponding to different asymptotic behaviors (a)
exponential, (b) stretched exponential and (c) power-law.

,~, 0
0

"~ -2
tzr
0 -4

-6 -

F(0.5) x -05 E0.5,0.5(-x 05)


-8 I I I I I I I I I

-5 -4 -3 -2 -1 0 1 2 3 4 5
log x
Figure 20. Comparison of a Mittag-Leffier function, lhs. of equation (65), with
one of its Pad6 approximants, rhs. of equation (65).
464

The use of the generalized Mittag-Leffier function in numerical calculations is


hampered due to the sometimes slow convergence of the series in equation (61).
This can be taken care of by using a functional approximation; it turns out that
Pad6 approximants allow to evaluate the generalized Mittag-Leffier functions
almost exactly and without much effort. The procedure is described in [13]; in the
lowest order approximation it yields exemplarily for cz = 1 and 13= 0.5:

r(0.s) Eo.,,o.,(-x (65)


l+2x

The result is presented in Figure 20; note that deviations from the exact result are
small and are restricted to the transition region between the power-law domains.

REFERENCES

1. R. Richert and A. Blumen (eds.), Disorder Effects on Relaxational Processes:


Glasses, Polymers, Proteins, Springer, Berlin, 1994.
2. A. Blumen, in: Th. Dorfm011er and G. Williams (eds.), Molecular Dynamics
and Relaxation Phenomena in Glasses, Springer, Berlin, 1987, p. 1.
3. J. Klafter, R. J. Rubm and M. F. Shlesmger (eds.), Transport and Relaxation
in Random Materials, World Scientific, Singapore, 1986.
4. K. L. Ngai and G. B. Wright, Relaxation in Complex Systems, Naval
Research Lab., Springfield, VA, 1984.
5. G. Williams and D. C. Watts, Trans. Faraday Soc., 66 (1970), 80; G.
Williams, Adv. Polym. Sci., 33 (1979), 59.
6. F. Kohlrausch, Pogg Ann. Physik, 29 (1863), 337.
7. J. Friedrich and A. Blumen, Phys. Rev. B, 32 (1985), 1434.
8. A.V. Tobolsky and E. Catsiff, Journal of Polymer Science, 19 (1956), 111.
9. M. Mours and H.H. Wimer, Chapter in this book.
10. Chr. Friedrich and H. Braun H., Colloid Polym. Sci., 272 (1994), 1536.
11. Chr. Friedrich, Rheol. Acta, 30 (1991), 151.
12. Chr. Friedrich, in: J. Casas-Vazquez and D. Jou (eds.), Rheological
modelling: Thermodynamical and Statistical Approaches, Lecture Notes in
Physics No. 381, Springer, Berlin, 1991, p. 321.
13. Chr. Friedrich and H. Braun, Rheol. Acta, 31 (1992), 309.
14. Chr. Friedrich and S. Hazanov, in: Advances in Structured and
Heterogeneous Continua, Allerton Press Inc., New York, 1994, p. 173.
465

15. Chr. Friedrich and L. Heymann, J. Rheol., 32 (1988), 235.


16. Chr. Friedrich, L. Heymann and H.-R. Berger, Rheol. Acta, 28 (1989), 535.
17. Chr. Friedrich, J. Non-Newt. Fluid Mech., 46 (1993), 307.
18. Chr. Friedrich, Phil. Mag. Letters, 66 (1992), 287.
19. Chr. Friedrich, H. Braun and J. Weese, Polym. Eng. Sci., 35 (1995), 1661.
20. M. Caputo and F. Meinardi, Riv. Nuovo Cimento (Ser. 2), 1 (1971), 161.
21. F. Meinardi and E. Bonetti, Progress and Trends in Rheology II, Suppl.
Rheol. Acta, 26 (1988), 64.
22. F. Meinardi, Fractional relaxation in anelastic solids, J. All. & Comp.,
211/212 (1994), 534.
23. M. Abramowitz and I. A. Stegun (eds.), Handbook of Mathematical
Functions, Dover, New York, 1972.
24. N.W. Tschoegl, The Phenomenological Theory of Linear Viscoelastic
Behavior, Springer, Berlin, 1989.
25. I.M. Ward, Mechanical Properties of Solid Polymers, Wiley, Chichester,
1983.
26. K.B. Oldham and J. Spaniel The Fractional Calculus, Academic, New York,
1974.
27. K.S. Miller and B. Ross, An Introduction to the Fractional Calculus and
Fractional Differential Equations, Wiley, New York, 1993.
28. A. Gemant, Physics, 7 (1936), 311.
29. A. Gemant, Phil. Mag. 25 (1938), 540.
30. P.G. Nutting, J. of the Franklin Institute, 191 (1921), 679.
31. P. G. Nutting, Proc. Amer. Soc. Test. Mater., 21 (1921), 1162.
32. P. Kobeko, E. Kuvshinskij and G. Gurevitch, Techn. Phys USSR, 4 (1937),
622.
33. A. P. Alexandrov and Yu. S. Lazurkin, J. Tech. Fiz., 9 (1939), 1250; 1261.
34. A. LeMehaute, L. Picard and L. Fruchter, Phil. Mag. B, 52 (1985), 1071.
35. G.W. Scott-Blair and F.M.V. Coppen, Amer. J. Psychol, 56 (1943), 234.
36. J.E. Caffyn and G.W. Scott-Blair, Nature, 155 (1945), 171.
37. G.W. Scott-Blair, B.C. Veinoglou and J.E. Caffyn, Proc. Roy. Soc. Ser. A,
189 (1947), 69.
38. G.W. Scott-Blair and J.E. Caffyn, Phil. Mag., 40 (1949), 80.
39 R.C.L. Bosworth, Nature, 157 (1946), 447.
40 G.L. Slonimsky, Journal of Polymer Science: Part C, 16 (1967), 1667.
41 W. Smit and H. de Vries, Rheol. Acta, 9 (1970), 525.
42 W.G. G10ckle and T.F. Nonnenmacher, Macromolecules, 24 (1991), 6426.
43 T.F. Nonnenmacher, m: J. Casas-Vfizquez and D. Jou (eds.), Lecture Notes in
Physics No. 381, Springer, Berlin, 1991, p. 309.
466

44. T.F. Nonnenmacher, W.G. G1ockle, Phil. Mag. Letters, 64 (1991), 89.
45. R.C. Koeller, J. Appl. Mech., 51 (1984), 299.
46. Y.N. Rabotnov, Elements of Hereditary Solid Mechanics, Mir Publishers,
Moscow, 1980.
47. Chr. Friedrich and S. Hazanov, in: D.A. SiNner and Y.G. Yanovsky (eds.),
Advances in Structured and Heterogeneous Continua, Allerton Press, New
York, 1994.
48 R.L. Bagley, J. Rheol., 27 (1983), 201.
49. R.L. Bagley and P.J. Torvik, J. Rheol., 30 (1986), 133.
50 L. Rogers, J. Rheol., 27 (1983), 351.
51 W.E. VanArsdale, J. Rheol., 29 (1985), 851.
52 W.G. G1Ockle, T.F. Nonnenmacher, Rheol. Acta, 33 (1994), 337.
53 R. Metzler, W. Schick, H.-G. Kilian and T.F. Nonnenmacher, J. Chem. Phys.,
103 (1995), 7180.
54. H. Schiessel and A. Blumen, J. Phys. A, 26 (1993), 5057.
55. H. Schiessel, P. Alemany and A. Blumen, Progr. Colloid Polym. Sci., 96
(1994), 16.
56. H. Schiessel and A. Blumen, Macromolecules, 28 (1995), 4013.
57. H. Schiessel and A. Blumen, Fractals, 3 (1995), 483.
58. H. Schiessel, R. Metzler, A. Blumen and T.F. Nonnenmacher, J. Phys. A, 28
(1995), 6567.
59 J.-C. Bauwens, Colloid Polyrn. Sci. 270 (1992), 537.
60 N. Heymans and J.-C. Bauwens, Rheol. Acta, 33 (1994), 210.
61 N. Heymans, Rheol. Acta, 35 (1996), 508.
62 M. Giona, S. Cerbelli and H.E. Roman, Physica A, 191 (1992), 449.
63 J. Stastna, L. Zanzotto and K. Ho, Rheol. Acta 33 (1994), 344.
64 M. Odenwald, H.-F. Eicke and Chr. Friedrich, Colloid Polym. Sci., 274
(1996), 568; M. Odenwald, Diploma-Thesis, Freiburg 1993.
65. Chr. Friedrich and H. Dehno, m: Progress and Trends in Rheology IV,
Steinkopffverlag, Darmstadt, 1994, p. 45.
66. L.I. Palade, V. Vemey and P. Attane, Rheol. Acta, 35 (1996), 265.
67. S. Hellinckx, Colloid Polym. Sci., 275 (1997), 116.
68. S. Havlin and A. Btmde, in: A. Bunde and S. Havlm (eds.), Fractals and
Disordered Systems, Springer, Berlin, 1991, p. 97.
69. M. Mours and H.H. Winter, Macromolecules, 29 (1996), 7221.
70. Chr. Friedrich, Acta Polymerica, 46 (1995), 385.
71. A. Erd61yi (ed.), Bateman Manuscript Project, Higher Transcendental
Functions, Vol. III, Mc.Graw-Hill, New York, 1955.
467

T H E K I N E T I C T H E O R Y OF D I L U T E S O L U T I O N S OF
FLEXIBLE POLYMERS: HYDRODYNAMIC INTERACTION

J. R a v i P r a k a s h a

~Department of Chemical Engineering, Indian Institute of Technology,


Madras, India, 600 036

1. I N T R O D U C T I O N

The rheological properties of dilute polymer solutions are commonly used


in industry for characterising the dissolved polymer in terms of its molec-
ular weight, its mean molecular size, its chain architecture, the relaxation
time spectrum, the translational diffusion coefficient and so on. There is
therefore considerable effort world wide on developing molecular theories
that relate the microscopic structure of the polymer and its interactions
with the solvent to the observed macroscopic behavior. In this chapter,
recent theoretical progress that has been made in the development of a
coherent conceptual framework for modelling the rheological properties of
dilute polymer solutions is reviewed.
A polymer solute molecule dissolved in a dilute Newtonian solvent is
typically represented in molecular theories by a coarse-grained mechanical
model, while the relatively rapidly varying motions of solvent molecules
surrounding the polymer molecule are replaced by a random force field
acting on the mechanical model. The replacement of the complex poly-
mer molecule with a coarse-grained mechanical model is justified by the
belief that such models capture those large scale properties of the polymer
molecule, such as its stretching and orientation by the solvent flow field,
that are considered to be responsible for the solution's macroscopic behav-
ior. An example of a coarse-grained model frequently used to represent a
flexible polymer molecule is the bead-spring chain, which is a linear chain
of identical beads connected by elastic springs.
Progress in the development of molecular theories for dilute polymer so-
lutions has essentially involved the succesive introduction, at the molecular
level, of various physical phenomena that are considered to be responsible
468

for the macroscopic properties of the polymer solution. For instance, the
simplest theory based on a bead-spring model assumes that the solvent
influences the motion of the beads by exerting a drag force and a Brown-
ian force. Since this theory fails to predict a large number of the observed
features of polymer solutions, more advanced theories have been developed
which incorporate additional microscopic phenomena. Thus, theories have
been developed which (i) include the phenomenon of 'hydrodynamic inter-
action' between the beads, (ii) try to account for the finite extensibility of
the polymer molecule, (iii) attempt to ensure that two parts of the poly-
mer chain do not occupy the same place at the same time, (iv) consider
the internal friction experienced when two parts of a polymer chain close
to each other in space move apart, and so on. The aim of this chapter is
to present the unified framework within which these microscopic phenom-
ena may be treated, and to focus in particular on recent advances in the
treatment of the effect of hydrodynamic interaction. To a large extent, the
notation that is used here is the same as that in the treatise Dynamics of
Polymeric Liquids by Bird and co-authors [2].

2. T R A N S P O R T PROPERTIES OF D I L U T E S O L U T I O N S
2.1. D i l u t e s o l u t i o n s
A solution is considered dilute if the polymer chains are isolated from
each other and have negligible interactions with each other. In this regime
of concentration the polymer solution's properties are determined by the
nature of the interaction between the segments of a single polymer chain
with each other, and by the nature of the interaction between the segments
and the surrounding solvent molecules. As the concentration of polymers
is increased, a new threshold is reached where the polymer molecules begin
to interpenetrate and interact with each other. This threshold is reached
at a surprisingly low concentration, and heralds the inception of the semi-
dilute regime, where the polymer solution's properties have been found to
be significantly different. Beyond the semi-dilute regime lie concentrated
solutions and melts. In this chapter we are concerned exclusively with the
behavior of dilute solutions.
A discussion of the threshold concentration at which the semi-dilute
regime in initiated is helpful in introducing several concepts that are used
frequently in the description of polymer solutions.
469

A polymer molecule surrounded by solvent molecules undergoes thermal


motion. A measure of the average size of the polymer molecule is the
root mean square distance between the two ends of the polymer chain,
typically denoted by R. This size is routinely measured with the help
of scattering experiments, and is found to increase with the molecular
weight of the polymer chain with a scaling law, R ~ M ~, where M is
the molecular weight, and u is the scaling exponent which depends on
the nature of the polymer-solvent interaction. In good solvents, solute-
solvent interactions are favoured relative to solute-solute interactions. As
a consequence the polymer chain swells and its size is found to scale with
an exponent u = 3/5. On the other hand, in poor solvents, the situation
is one in which solute-solute interactions are preferred. There exists a
particular temperature, called the theta temperature, at which the scaling
exponent u changes dramatically from 3/5 to 1/2. At this temperature,
the urge to expand caused by two parts of the chain being unable to occupy
the same location (leading to the presence of an excluded volume), is just
balanced by the repulsion of the solute molecules by the solvent molecules.
Polymer chains in a solution can be imagined to begin to interact with
each other when the solution volume is filled with closely packed spheres
representing the average size of the molecule. This occurs when np R 3 ~ 1,
where np in the number of chains per unit volume. Since np - pp N A / M ,
where pp is the polymer mass density and NA is Avagadro's number, it fol-
lows that polymer density at overlap, p~, scales with molecular weight as,
pp ,-~ M 1-3~. Polymer molecules typically have molecular weights between
104 and 106 gm/mol. As a result, it is clear that the polymer solution can
be considered dilute only at very low polymer densities. Since experimental
measurements are difficult at such low concentrations, the usual practice is
to extrapolate results of experiments carried out at decreasing concentra-
tions to the limit of zero concentration. For instance, in the case of dilute
polymer solutions it is conventional to report the intrinsic viscosity, which
is defined by,

[r/] - lim ~p (1)


pp~O pp rls

where Up is the polymer contribution to the solution viscosity, and r/s is the
solvent viscosity.
470

2.2. H o m o g e n e o u s flows
Complex flow situations typically encountered in polymer processing fre-
quently involve a combination of shearing and extensional deformations.
The response of the polymer solution to these two modes of deformation
is very different. Consequently, predicting the rheological properties of the
solution under both shear and extensional deformation is considered to be
very important in order to properly characterise the solutions behavior.
Rather than considering flows where both these modes of deformation are
simultaneously present, it is common in polymer kinetic theory to anal-
yse simpler flow situations called homogeneous flows, where they may be
treated separately.
A flow is called homogeneous, if the rate of strain tensor, "~ - ( V v ) ( t ) +
(~7v)t(t), where v is the solution velocity field, is independent of position.
In other words, the solution velocity field v in homogeneous flows, can
always be represented as v = v0 + ~(t) -r, where v0 is a constant vector,
~(t) - Vv(t) is a traceless tensor for incompressible fluids, and r is the
position vector with respect to a laboratory fixed frame of reference. While
there is no spatial variation in the rate of strain tensor in homogeneous
flows, there is no restriction with regard to its variation in time. Therefore,
the response of dilute solutions to transient shear and extensional flows is
also used to probe its character as an alternative means of characterisation
independent of the steady state material functions.
Two homogeneous flows, steady simple shear flow and small amplitude
oscillatory shear flow, that are frequently used to validate the predictions
of molecular theories which incorporate hydrodynamic interaction, are de-
scribed briefly below. A comprehensive discussion of material functions in
various flow situations can be found in the book by Bird et al. [1].

2.3. Simple shear flows


The rheological properties of a dilute polymer solution can be obtained
once the stress tensor, 7", is known. The stress tensor is considered to be
given by the sum of two contributions, 7" - 7"s + ~-P, where 7-8 is the con-
tribution from the solvent, and T"p is the polymer contribution. Since the
solvent is assumed to be Newtonian, the solvent stress (using a compressive
definition for the stress tensor [1]) is given by, ~.s _ _ r/8 ~/. The nature of
the polymer contribution "/'P in simple shear flows is discussed below.
471

Simple shear flows are described by a velocity field,

vz-O (2)

where the velocity gradient ~yx can be a function of time. From consid-
erations of symmetry, one can show that the most general form that the
polymer contribution to the stress tensor can have in simple shear flows
is [1],

% % 0 (3)
o o rPzz
where the matrix of components in a Cartesian coordinate system is dis-
played. The form of the stress tensor implies that only three independent
combinations can be measured for an incompressible fluid. All simple shear
flows are consequently characterised by three material functions.

2.3.1. Steady simple shear flows


Steady simple shear flows are described by a constant shear rate, ~ -
I')yxl- The tensor ~ is consequently given by the following matrix represen-

(~176/
tation in the laboratory-fixed coordinate system,

~-x/ 0 0 0 (4)
0 0 0
The three independent material functions used to characterize such flows
are the viscosity, rip, and the first and second normal stress difference co-
efficients, ~1 and ~2, respectively. These functions are defined by the fol-
lowing relations,
< _ _ - (5)

where TPy,~'Px,7~Pyare the components of the polymer contribution to the


stress tensor r p.
At low shear rates, the viscosity and the first normal stress coefficient are
observed to have constant values, rip,0 and ~1,0, termed the zero shear rate
viscosity and the zero shear rate first normal stress coefficient, respectively.
At these shear rates the fluid is consequently Newtonian in its behavior.
At higher shear rates, most dilute polymer solutions show shear thinning
behavior. The viscosity and the first normal stress coefficient decrease
472

with increasing shear rate, and exhibit a pronounced power law region.
At very high shear rates, the viscosity has been observed to level off and
approach a constant value, r/p,~, called the infinite shear rate visosity. A
high shear rate limiting value has not been observed for the first normal
stress coefficient. The second normal stress coefficient is much smaller in
magnitude than the first normal stress coefficient, however its sign has not
been conclusively established experimentally. Note that the normal stress
differences are zero for a Newtonian fluid. The existence of non-zero normal
stress differences is an indication that the fluid is viscoelastic.
Experiments with very high molecular weight systems seem to suggest
that polymer solutions can also shear thicken. It has been observed that
the viscosity passes through a minimum with increasing shear rate, and
then increases until a plateau region before shear thinning again [17].
It is appropriate here to note that shear flow material functions are
usually displayed in terms of the reduced variables, ~?p/~?p,O,~1/~1,o and
~2/~1, versus a non-dimensional shear rate fl, which is defined by fl = )~ph/,
where, Ap = [r/J0M rl~/NA kB T, is a characteristic relaxation time. The
subscript 0 on the square bracket indicates that this quantity is evalu-
ated in the limit of vanishing shear rate, kB is Boltzmann's constant and
T is the absolute temperature. For dilute solutions one can show that,
[?7]/[?']]0 - - ?~p/?~p,O and ~ - ?~p,O ;y/np kB T.

2.3.2. Small amplitude oscillatory shear flow


A transient experiment that is used very often to characterise polymer
solutions is small amplitude oscillatory shear flow. The upper plate in a
simple shear experiment is made to undergo sinusoidal oscillations in the
plane of flow with frequency w. For oscillatory flow between narrow slits,
the shear rate at any position in the fluid is given by [1], ~x(t) - % coswt,
where % is the amplitude. The tensor ~(t) is consequently given by,

0 1 0)
~(t)-%coswt 0 0 0 (6)
0 0 0
Since the polymer contribution to the shear stress in oscillatory shear
flow, ~-~,
p undergoes a phase shift with respect to the shear strain and the
strain rate, it is customary to represent its dependence on time through
473

the relation [1],


7:.yvx -- - r / ' ( w ) a / o cos wt- ~7"(w)"9o s i n w t (7)
where r / a n d ~" are the material functions characterising oscillatory shear
flow. It is common to represent them in a combined form as the complex
viscosity, y* = rf - i r/".
Two material functions which are entirely equivalent to r]/ and r]" and
which are often used to display experimental data, are the storage modulus
G ' - w~"/(nkBT) and the loss modulus G " - wrf/(nkBT). Note that the
term involving G' in equation(7) is in phase with the strain while that
involving G" is in phase with the strain rate. For an elastic material,
G" - 0, while for a Newtonian fluid, G' - 0. Thus, G' and G" are measures
of the extent of the fluid's viscoelasticity.
In flow situations which have a small displacement gradient, termed the
linear viscoelastic flow regime, the stress tensor in polymeric fluids is de-
scribed by the linear constitutive relation,

- - c(t - #(t, (8)

where G(t) is the relaxation modulus.


When the amplitude "~0 is very small, oscillatory shear flow is a linear
viscoelastic flow and consequently can also be described in terms of a relax-
ation modulus G(t). Indeed, expressions for the real and imaginary parts
of the complex viscosity can be found from the expression,

- f0 (9)
Experimental plots of log G' and log G" versus nondimensional frequency
show three distinct power law regimes. The regime of interest is the in-
termediate regime [17], where for dilute solutions of high molecular weight
polymers in good or theta solvents, both G' and G" have been observed to
scale with frequency as w 2/3.
It is appropriate to note here that the zero shear rate viscosity ~p,0 and
the zero shear rate first normal stress difference ~1,0, which are linear
viscoelastic properties, can be obtained from the complex viscosity in the
limit of vanishing frequency,
2 rf(w) (10)
~p,0 - ~a-+0
lira r/',,(w) ; 91,0 - lira
~a-+0 03
474

2.4. Scaling w i t h m o l e c u l a r w e i g h t
We have already discussed the scaling of the root mean square end-to-end
distance of a polymer molecule with its molecular weight. In this section
we discuss the scaling of the zero shear rate intrinsic viscosity It/]0, and the
translational diffusion coefficient D, with the molecular weight, M. As we
shall see later, these have proven to be vitally important as experimental
benchmarks in attempts to improve predictions of molecular theories.
It has been found that the relationship between [y]0 and M can be ex-
pressed by the formula,

[~]o-KM a (11)

where, a is called the M a r k - H o u w i n k exponent, and the prefactor K de-


pends on the polymer-solvent system. The value of the parameter a lies
between 0.5 and 0.8, with the lower limit corresponding to theta condi-
tions, and the upper limit to a good solvent with a very high molecular
weight polymer solute. Measured intrinsic viscosities are routinely used to
determine the molecular weight of samples once the constants K and a are
known for a particular polymer-solvent pair.
The translational diffusion coefficient D for a flexible polymer in a dilute
solution can be measured by dynamic light scattering methods, and is
found to scale with molecular weight as [2],
D-~ M -" (12)

where the exponent # lies in the range 0.49 to 0.6. Most theta solutions
have values of # close to the lower limit. On the other hand, there is wide
variety in the value of # reported for good solvents. It appears that the
upper limit is attained only for very large molecular weight polymers and
the intermediate values, corresponding to a cross over region, are more
typical of real polymers with moderate molecular weights.

2.5. U n i v e r s a l b e h a v i o r
It is appropriate at this point to discuss the most important aspect of
the behavior of polymer solutions (as far as the theoretical modelling of
these solutions is concerned) that is revealed by the various experimental
observations. When the experimental data for high molecular weight sys-
tems is plotted in terms of appropriately normalized coordinates, the most
noticeable feature is the exhibition of u n i v e r s a l behavior. By this it is
475

meant that curves for different values of a parameter, such as the molec-
ular weight, the temperature, or even for different type8 of monomer8 can
be 8uperposed onto a 8ingle curve. For example, when the reduced intrin-
sic viscosity, [u]/[n]0 is plotted as a function of the reduced shear rate/ ,
the curves for polystyrene in different type8 of good solvent8 at various
temperatures collapse onto a 8ingle curve [1].
There is, however, an important point that must be noted. While poly-
mer8 dissolved in both theta solvents and good solvents show universal
behavior, the universal behavior is different in the two cases. An example
of this i8 the observed sca]ing behavior of various quantities with molecular
weight. The scaling is universal within the context of a particular type of
801vent. The term universality class is used to describe the set of systems
that exhibit common universal behavior [40]. Thus theta and good 80lvents
be]ong to different universality classes.
The existence of universality classes is very significant for the theoret-
ical description of polymer solutions. Any attempt made at modelling a
polymer so]ution's properties might expect that a proper description must
incorporate the chemical structure of the polymer into the model, since
this determines its microscopic behavior. Thus a detailed consideration of
bonds, sidegroups, etc. may be envisaged. However, the universal behav-
ior that i8 revealed by experiments suggests that macroscopic properties of
the polymer solution are determined by a few large scale properties of the
polymer molecule. Structural details may be ignored since at length scales
in the order of nanometer8, different polymer molecule8 become equivalent
to each other, and behave in the same manner. A8 a result, polymer 80-
lutions that differ from each other with regard to the chemical structure
or molecular weight of the polymer mo]ecu]es that are dissolved in it, the
temperature, and so on, still behave similarly as ]ong as a few parameter8
that describe molecular features are the same.
This universal behavior justifies the introduction of crude mechanical
models, such as the bead-spring chain, to represent real polymer molecules.
On the other hand, it is interesting to note that in many cases, the pre-
dictions of these models are not universal. It turns out that apart from a
basic length and time scale, there occur other parameters that need to be
prescribed, for example, the number of beads N in the chain, the strength
of hydrodynamic interaction h*, the finite spring extensibility parameter
476

b, and so on. It is perhaps not incorrect to state that any molecular the-
ory that is developed must ultimately verify that universal predictions of
transport properties are indeed obtained. The universal predictions of ki-
netic theory models with hydrodynamic interaction are discussed later on
in this chapter.

3. B E A D - S P R I N G CHAIN MODELS

The development of a kinetic theory for dilute solutions has been ap-
proached in two different ways. One of them is an intuitive approach in
the configuration space of a single molecule, with a particular mechanical
model chosen to represent the macromolecule, such as a freely rotating
bead-rod chain or a freely jointed bead-spring chain [13,36,45]. The other
approach is to develop a formal theory in the phase space of the entire
solution, with the polymer molecule represented by a general mechanical
model that may have internal constraints, such as constant bond lengths
and angles [16,5,2]. The results of the former method are completely con-
tained within the latter method, and several ad hoc assumptions made in
the intuitive treatment are clarified and placed in proper context by the
development of the rigorous phase space theory. Kinetic theories devel-
oped for flexible macromolecules in dilute solutions have generally pursued
the intuitive approach, with the bead-spring model proving to be the most
popular. This is because the lack of internal constaints in the model makes
the formulation of the theory simpler. Recently, Curtiss and Bird [4], ac-
knowledging the 'notational and mathematical' complexity of the rigorous
phase space theory for general mechanical models, have summarised the
results of phase space theory for the special case of bead-spring models
with arbitrary connectivity, ie. for linear chains, rings, stars, combs and
branched chains.
In this section, since we are primarily concerned with reviewing recent
developments in theories for flexible macromolecules, we describe the devel-
opment of kinetic theories in the configuration space of a single molecule.
However, readers who wish the understand the origin of the ad hoc expres-
sions used for the Brownian forces and the hydrodynamic force, and the
formal development of expressions for the momentum and mass flux, are
urged to read the article by Curtiss and Bird [4].
The general diffusion equation that governs the time evolution of the
477

distribution of configurations of a bead-spring chain subject to various


nonlinear effects, and the microscopic origin of the polymer contribution
to the stress tensor are discussed in this section. The simplest bead-spring
chain model, the Rouse model is also discussed. We begin, however, by
describing the equilibrium statistical mechanical arguments that justify
the representation of a polymer molecule with a bead-spring chain model,
and we discuss the equilibrium configurations of such a model.

3.1. Equilibrium configurations


When a flexible polymer chain in a quiescent dilute solution is considered
at a lowered resolution, ie. at a coarse-grained level, it would appear like
a strand of highly coiled spaghetti, and the extent of its coiling would
depend on its degree of flexibility. A quantity used to characterise a chain's
flexibility is the orientational correlation function, whose value Kor (A~),
is a measure of the correlation in the direction of the chain at two different
points on the chain which are separated by a distance Ag along the length
of the chain. At sufficiently large distances Ag, it is expected that the
correlations vanish. However, it is possible to define a persistence length
gps, such that for Ag > gps, orientational correlations are negligible [40].
The existence of a persistence length suggests that as far as the global
properties of a flexible polymer chain are concerned, such as the distribu-
tion function for the end-to-end distance of the chain, the continuous chain
could be replaced by a freely jointed chain made up of rigid links connected
together at joints that are completely flexible, whose linear segments are
each longer than the persistence length ~ps, and whose contour length is
the same as that of the continuous chain.
The freely jointed chain undergoing thermal motion is clearly analogous
to a random-walk in space, with each random step in the walk representing
a link in the chain assuming a random orientation. Thus all the statistical
properties of a random-walk are, by analogy, also the statistical proper-
ties of the freely jointed chain. The equivalence of a polymer chain with
a random-walk lies at the heart of a number of fundamental results in
polymer physics.

3.1.1. Distribution ]unctions and averages


In polymer kinetic theory, the freely jointed chain is assumed to have
beads at the junction points betwen the links, and is referred to as the freely
478

jointed bead-rod chain [2]. The introduction of the beads is to account for
the mass of the polymer molecule a n d the viscous drag experienced by
the polymer molecule. While in reality the mass and drag are distributed
continuously along the length of the chain, the model assumes that the
total mass and drag may be distributed over a finite number of discrete
beads.
For a general chain model consisting of N beads, which have position
vectors r~, u = 1, 2 , . . . , N, in a laboratory fixed coordinate system, the
Hamiltonian is given by,

7{ --/C + r (rl, r 2 , . . . , r N ) (13)

where K: is the kinetic energy of the system and r is the potential energy.
r depends on the location of all the particles.
The center of mass r~ of the chain, and its velocity i% are given by
1 N 1 N
rc = N ~=1
Er~ ; /'~= N E=/ ' ~ (14)
~-1
where/-~ = dr~/dt. The location of a bead with respect to the center of
mass is specified by the vector R~ = r~ - re.
If Q1, Q2, ... Qd denote the generalised internal coordinates required to
specify the configuration of the chain, then the kinetic energy of the chain
in terms of the velocity of the center of mass and the generalised velocities
Q ~ - dQ~/dt, is given by [2],
I c _ m N .2 1
2 rc + -2 V E g~t Qs Qt (15)
t

where the indices s and t vary from 1 to d, m is the mass of a bead, and
g~t is the metric matrix, defined by, g~t - m E~ (OR~/OQ~). (OR~/OQt).
In terms of the momentum of the center of mass, Pc = m N/%, and the
generalised momenta P~, defined by, P~ - (OIC]OQ~), the kinetic energy
has the form [2],

1 2 1
K: - 2m N p~ + 2 V y] G~t P~ Pt (16)
t

where, G st are the components of the matrix inverse to the metric matrix,
F~t G~t gtu - 5s~, and 5~u is the Kronecker delta.
479

The probability, ~eqdrcdQ dp~dP, that an N-bead chain model has a


configuration in the range dr~ dQ about r~, Q and momenta in the range
dp~ dP about p~, P is given by,
Peq (r~, Q, pc, P ) - Z -1 e -u/k"T (17)

where Z is the partition function, defined by,


Z I/f/ e -u/ksT drc dQ dp~ dP (18)

The abbreviations, Q and dQ have been used to denote Q1, Q 2 , . . . , Qd


and dQ1 dQ2 ... dQd, respectively, and a similar notation has been used
for the momenta.
The configurational distribution function for a general N-bead chain,
Ceq ( Q ) d Q , which gives the probability that the internal configuration
is in the range dQ about Q, is obtained by integrating Peq over all the
momenta and over the coordinates of the center of mass,
%, ( Q ) - Z -~ / / / e -n/kBT drc dp~ dR (19)

For an N-bead chain whose potential energy does not depend on the loca-
tion of the center of mass, the following result is obtained by carrying out
the integrations over Pc and P [2],

~)eq ( Q ) - ~/g(Q) e-r


$ ~/-g(Q)e-r dQ (20)

where, g(Q) - det(gst) - 1/det(G,t).


An expression that is slightly different from the random-walk distribution
is obtained on evaluating the right hand side of equation (20) for a freely
jointed bead-rod chain. Note that the random-walk distribution is obtained
by assuming that each link in the chain is oriented independently of all the
other links, and that all orientations of the link are equally likely. On
the other hand, equation (20) suggests that the probability for the links
in a freely jointed chain being perpendicular to each other, for a given
solid angle, is slightly larger than the probability of being in the same
direction. Inspite of this result, the configurational distribution function
for a freely jointed bead-rod chain is almost always assumed to be given by
the random-walk distribution [2]. Here afterwards in this chapter, we shall
refer to a freely jointed bead-rod chain whose configurational distribution
480

function is assumed to be given by the random-walk distribution, as an


ideal chain. For future reference, note that the random-walk distribution
is given by,
( 1 ) N-'I N-1
Ceq (01,.--, 0N-l, (~1,--., (~N-1) -- ~ H sin Oi (21)
i=l

where Oi and r are the polar angles for the ith link in the chain [2].
Since the polymer chain explores many states in the duration of an ob-
servation quantities observed on macroscopic length and time scales are
averages of functions of the configurations and momenta of the polymer
chain. A few definitions of averages are now introduced that are used
frequently subsequently in the chapter.
The average value of a function X (r~, Q, pc, P ), defined in the phase
space of a polymer molecule is given by,

(X)e. -- f f f f x ~:)eqdrc dQ dp~ dR (22)


We often encounter quantities X that depend only on the internal configu-
rations of the polymer chain and not on the center of mass coordinates or
momenta. In addition, if the potential energy of the chain does not depend
on the location of the center of mass, then it is straight forward to see that
the equilibrium average of X is given by,
- f x r (23)
3.1.2. The end-to-end vector
The end-to-end vector r of a general bead-rod chain can be found by
summing the vectors that represent each link in the chain,
N-1
r - y~ a ui (24)
i=1
where a is the length of a rod, and ui is a unit vector in the direction of the
ith link of the chain. Note that the components of the unit vectors ui, i =
1, 2 , . . . , N - 1, can be expressed in terms of the generalised coordinates
Q [2]
The probability Peq(r)dr, that the end-to-end vector of a general bead-
rod chain is in the range dr about r can be found by suitably contracting
the configurational distribution function Ceq ( Q ) [2],

Peq(r)- f 6 ( r - Y~ a u i )r (Q)dQ (25)


i
481

where 5(.) represents a Dirac delta function.


With Ceq ( Q ) given by the random-walk distribution (21), it can be
shown that for large values of N and r - Irl < 0.5Na, the probability
distribution for the end-to-end vector is a G aussian distribution,
3 3/2 _3r 2
1)a') ox. (. 1)a')
The distribution function for the end-to-end vector of an ideal chain with
a large number of beads N is therefore given by the Gaussian distribu-
tion (26/.
The mean square end-to-end distance, / r2 )eq, for an ideal chain can then
be shown to be, (r2)eq - ( N - 1) a 2. This is the well known result that the
root mean square of the end-to-end distance of a random-walk increases
as the square root of the number of steps. In the context of the polymer
chain, since the number of beads in the chain is directly proportional to
the molecular weight, this result implies that R ~ M ~ We have seen
earlier that this is exactly the scaling observed in theta solvents. Thus one
can conclude that a polymer chain in a theta solvent behaves like an ideal
chain.

3.1.3. The bead-springchain


Consider an isothermal system consisting of a bead-rod chain with a
constant end-to-end vector r, suspended in a bath of solvent molecules at
temperature T. The partition function of such a constrained system can
be found by contracting the partition function in the constraint-free case,

Z (r) - f f f f 5(r- ~ aui) e-~/kBTdrcdQdpcdP (27)


i

For an N-bead chain whose potential energy does not depend on the loca-
tion of the center of mass, the integrations over rc, Pc and P can be carried
out to give,

Z(r)- C/5(r- ~
i
aui)r (28)

Comparing this equation with the equation for the end-to-end vector (25),
one can conclude that,
Z (r) -- C Peq (r) (29)
482

In other words, the partition function of a general bead-rod chain (except


for a multiplicative factor independent of r) is given by Peq (r). This result
is essential to establish the motivation for the introduction of the bead-
spring chain model.
At constant temperature, the change in free energy accompanying a
change in the end-to-end vector r of a bead-rod chain, by an infinitesi-
mal amount dr, is equal to the work done in the process, ie., dA- F-dr,
where F is the force required for the extension. The Helmholtz free energy
of a general bead-rod chain with fixed end-to-end vector r can be found
from equation (29),

A(r) - - k B T In Z ( r ) - A o - kBT In Peq (r) (30)

where A0 is a constant independent of r. For an ideal chain, it follows


from equations (26) and (30), that a change in the end-to-end vector by
dr, leads to a change in the free energy dA, given by,
3kBT
dA(r) - (N - 1)a 2 r - d r (31)

Equation (31) implies that there is a tension F in the ideal chain, F -


( 3 k B T / ( N - 1)a 2) r, which resists any attempt at chain extension. Fur-
thermore, this tension is proportional to the end-to-end vector r. This
implies that the ideal chain acts like a Hookean spring, with a spring con-
stant H given by,

H- 3kBT (32)
(N- 1)a2
The equivalence of the behavior of an ideal chain to that of a Hookean
spring is responsible for the introduction of the bead-spring chain model.
Since long enough sub-chains within the ideal chain also have normally
distributed end-to-end vectors, the entire ideal chain may be replaced by
beads connected to each other by springs. Note that each bead in a bead-
spring chain represents the mass of a sub-chain of the ideal chain, while
the spring imitates the behavior of the end-to-end vector of the sub-chain.
The change in the Helmholtz free energy of an ideal chain due to a
change in the end-to-end vector is purely due to entropic considerations.
The internal energy, which has only the kinetic energy contribution, does
not depend on the end-to-end vector. Increasing the end-to-end vector of
483

the chain decreases the number of allowed configurations, and this change
is resisted by the chain. The entropic origin of the resistance is responsible
for the use of the phrase entropic spring to describe the springs of the
bead-spring chain model.
The potential energy S, of a bead-spring chain due to the presence of
Hookean springs is the sum of the potential energies of all the springs in
the chain. For a bead-spring chain with N beads, this is given by,
N-1
1H E qi" qi (33)

where Qi = r i + l - ri is the bead connector vector between the beads i and


i + 1. The configurational distribution function for a Hookean bead-spring
chain may be found from equation (20) by substituting r - S, with the
Cartesian components of the connector vectors chosen as the generalised
coordinates Qs. The number of generalised coordinates is consequently,
d - 3 N - 3, reflecting the lack of any constraints in the model. Since g(Q)
is a constant independent of Q for the bead-spring chain model [2], one
can show that,
3/2
Ceq ( Q 1 , . . . , Q N - 1 ) - rI.
)
H
27~kBT
exp ( - H Qj) (34)
It is clear from equation (34) that the equilibrium distribution function for
each connector vector in the bead-spring chain is a Gaussian distribution,
and these distributions are independent of each other. From the property of
Gaussian distributions, it follows that the vector connecting any two beads
in a bead-spring chain at equilibrium also obeys a Gaussian distribution.
The Hookean bead-spring chain model has the unrealistic feature that
the magnitude of the end-to-end vector has no upper bound and can infact
extend to infinity. On the other hand, the real polymer molecule has a finite
fully extended length. This deficiency of the bead-spring chain model is
not serious at equilibrium, but becomes important in strong flows where
the polymer molecule is highly extended. Improved models seek to correct
this deficiency by modifying the force law between the beads of the chain
such that the chain stiffens as its extension increases. An example of such a
nonlinear spring force law that is very commonly used in polymer literature
is the finitely extensible nonlinear elastic (FENE) force law [2].
484

3.1.4. Excluded volume


The universal behavior of polymers dissolved in theta solvents can be
explained by recognising that all high molecular weight polymers dissolved
in theta solvents behave like ideal chains. However, a polymer chain can-
not be identical to an ideal chain since unlike the ideal chain, two parts
of a polymer chain cannot occupy the same location at the same time. In
the very special case of a theta solvent, the excluded volume force is just
balanced by the repulsion of the solvent molcules. In the more commonly
occuring case of good solvents, the excluded volume interaction acts be-
tween any two parts of the chain that are close to each other in space,
irrespective of their distance from each other along the chain length, and
leads to a swelling of the chain. This is a long range interaction, and as a
result, it seriously alters the macroscopic properties of the chain. Indeed
there is a qualitative difference, and this difference cannot be treated as
a small perturbation from the behavior of an ideal chain [40]. Curiously
enough however, all swollen chains behave similarly to each other, and
modelling this universal behavior was historically one of the challenges of
polymer physics [40,44,7,8,6]. Here, we very briefly mention the manner in
which the problem is formulated in the case of bead-spring chains.
The presence of excluded volume causes the polymer chain to swell.
However, the swelling ceases when the entropic retractive force balances
the excluded volume force. The retractive force arises due to the de-
creasing number of conformational states available to the polymer chain
due to chain expansion. This picture of the microscopic phenomenon
is captured by writing the potential energy of the bead-spring chain as
a sum of the spring potential energy and the potential energy due to
excluded volume interactions. The excluded volume potential energy is
found by summing the interaction energy over all pairs of beads # and u,
E - (1/2) E.,~=~
N E (r~ - r~), where E ( r ~ - r~) is a short-range function
u u lly t kr E - - k.T - r.); b ing t h e exclua a vol-
u m e parameter with dimensions of volume. The total potential energy of
a Hookean bead-spring chain with s excluded volume interactions
is consequently,

1 N-1 1 N
r - -~ H ~ Qi . Qi + -~v kBT Z 5(r~-ru) (35)
i--1 ~,~=1
tt#u
485

The equilibrium configurational distribution function of a polymer chain


in the presence of Hookean springs and excluded volume can be found by
substituting equation (35) into equation (20), and all average properties of
the chain can be found by using equation (23). Solutions to these equa-
tions in the limit of long chains have been found by using a number of
approximate schemes since an exact treatment is impossible. The most ac-
curate scheme involves the use of field theoretic and renormalisation group
methods [6]. The universal scaling of a number of equilibrium properties
of dilute polymer solutions with good solvents are correctly predicted by
this theory. For instance, the end-to-end distance is predicted to scale with
molecular weight as, R ~ M ~
The spring potential in equation (35) has been derived by considering
the Helmholtz free energy of an ideal chain, ie. under theta conditions.
It seems reasonable to expect that a more accurate derivation of the re-
tractive force in the chain due to entropic considerations would require the
treatment of a polymer chain in a good solvent. This would lead to a non-
Hookean force law between the beads [7,29]. Such non-Hookean force laws
have so far not been treated in non-equilibrium theories for dilute polymer
solutions with good solvents.

3.2. N o n - e q u i l i b r i u m c o n f i g u r a t i o n s
Unlike in the case of equilibrium solutions it is not possible to derive
the phase space distribution function for non-equilibrium solutions from
very general arguments. As we shall see here it is only possible to derive
a partial differential equation that governs the evolution of the configura-
tional distribution function by considering the conservation of probability
in phase space, and the equation of motion for the particular model chosen.
The arguments relevent to a bead-spring chain are developed below.

3.2.1. D i s t r i b u t i o n f u n c t i o n s and averages


The phase space of a bead-spring chain with N beads can be chosen to
be given by the 6N - 6 components of the bead position coordinates, and
the bead velocities such that,
~O ( r l , . . . , rN, r l , . . . , rg, t ) d r 1 . . , drN d~l . . . drN
is the probability that the bead-spring chain has an instantaneous config-
uration in the range d r l , . . . , d r g about r l , . . . , r g , and the beads in the
chain have velocities in the range d / h , . . . , drN about/'1,...,/~g.
486

The configurational distribution function ~, can be found by integrating


/) over all the bead velocities,
( r l , . . . , rN, t ) -- f . . .f :P d f l . . . diCN (36)
The distribution of internal configurations r is given by,
r ( Q 1 , . . . , QN-1, t) - f ~I,'(r~, Q 1 , - . . , QN-1, t ) d r c (37)
where, ~' - ~, as a result of the Jacobian relation for the configurational
vectors [2], [ 0 ( r l , . . . , r N ) / 0 ( r c , Q 1 , . . . , QN-1)I -- 1. Note that the nor-
malisation condition f r dQ1 dQ2 ... d Q N - 1 = 1 is satisfied by ~b. When
the configurations of the bead-spring chain do not depend on the location
of the center of mass, as in the case of homogeneous flows with no concen-
tration gradients, ( l / V ) r = q~, where V is the volume of the solution.
The velocity-space distribution function E is defined by,
7)
(rl,...,rN, r l , - . - , I ' N , t ) -- ~ (38)

Note that E satisfies the normalisation condition f . . . f E die1.., di'N = 1.


Under certain circumstances that are discussed later, it is common to as-
sume that the velocity-space distribution function is Maxwellian about the
mass-average solution velocity,
1
---- ArM exp [ 2 k B T [Trt(rl -- v) 2 -+- . . . nt- m ( r N -- v)2]] (39)

where ArM is the normalisation constant for the Maxwellian distribution.


Making this assumption implies that one expects the time scales involved
in equilibration processes in momentum space to be much smaller than the
time scales governing relaxation processes in configuration space.
Averages of quantities which are functions of the bead positions and
bead velocities are defined analogously to the those in the previous section,
namely,
(X) - f...f X 7) d r 1 . . , drN d ~ l . . , drN (40)
is the the phase space average of X, while the velocity-space average is,
[[x ]] - f f z dl'l . . . dr'N (41)
For quantities X that depend only on the internal configurations of the
polymer chain and not on the center of mass coordinates or bead velocities,
(X) - f X r d Q l d Q 2 . . , d Q N - 1 (42)
487

3.2.2. The equation of motion


The equation of motion for a bead in a bead-spring chain is derived
by considering the forces acting on it. The total force F~, on bead # is,
F~ - Ei F(~), where the F(~), i - 1, 2,..., are the various intramolecular
and solvent forces acting on the bead. The fundamental difference among
the various molecular theories developed so far for the description of dilute
polymer solutions lies in the kinds of forces F(~) that are assumed to be
acting on the beads of the chain. In almost all these theories, the accela-
ration of the beads due to the force F~ is neglected. A bead-spring chain
model incorporating bead inertia has shown that the neglect of bead iner-
tia is justified in most practical situations [37]. The equation of motion is
consequently obtained by setting F~ - 0. Here, we consider the following
force balance on each bead #,

F(h)+F(b)+F(r (i~) -- 0 (# 1, 2 , . . . , N ) (43)


where, F(h) is the hydrodynamic drag force, F (b) is the Brownian force,
F(r is the intramolecular force due to the potential energy of the chain,
and -p(iv)
-# is the force due to the presence of internal viscosity. These are
the various forces that have been considered so far in the literature, which
are believed to play a crucial role in determining the polymer solution's
transport properties. The nature of each of these forces is discussed in
greater detail below. Note that, as is common in most theories, external
forces acting on the bead have been neglected. However, their inclusion is
reasonably straight forward [2].
The hydrodynamic drag force F(h) is the force of resistance offered by
the solvent to the motion of the bead #. It is assumed to be proportional
to the difference between the velocity-averaged bead velocity [~i,u]] and the
local velocity of the solution,

F.(h) = _r [ _ (v. + v,)]' (44)


where ~ is bead friction coefficient. Note that for spherical beads with ra-
dius a, in a solvent with viscosity ~/,, the bead friction coefficient ~ is given
by the Stokes expression: ~ = 67rr/sa. The velocity-average of the bead ve-
locity is not carried out with the Maxwellian distribution since this is just
the mass-average solution velocity. However, it turns out that an explicit
evaluation of the velocity-average is unnecessary for the development of
488

the theory. Note that the velocity of the solution at bead # has two com-
ponents, the imposed flow field vu - v0 + ~(t) 9r~, and the perturbation
of the flow field v~t due to the motion of the other beads of the chain. This
perturbation is called hydrodynamic interaction, and its incorporation in
molecular theories has proved to be of utmost importance in the prediction
of transport properties. The presence of hydrodynamic interaction couples
the motion of one bead in the chain to all the other beads, regardless of
the distance between the beads along the length of the chain. In this sense,
hydrodynamic interaction is a long range phenomena.
The perturbation to the flow field v~(r) at a point r due to the presence
of a point force F ( r ~) at the point r ~, can be found by solving the linearised
Navier-Stokes equation [1,8],

v'(r) - ~ ( r - r ' ) . F(r') (45)

where ~ ( r ) , called the Oseen-Burgers tensor, is the Green's function of the


linearised Navier-Stokes equation,

n(r)-
1
87rq,r
(1 + rr )
r-2
(46)
The effect of hydrodynamic interaction is taken into account in polymer
kinetic theory by treating the beads in the bead-spring chain as point
particles. As a result, in response to the hydrodynamic drag force acting
on each bead, each bead exerts an equal and opposite force on the solvent
at the point that defines its location. The disturbance to the velocity at
the bead v is the sum of the disturbances caused by all the other beads in
the chain, v" - - ~u ~ u ( r ~ - ru). F (h), where, ~ - ~v~ is given by,

~u~--
{ 87rrlsru~
1 (1+ ru~ru~)
r2 , ru~--ru-r~, for##v (47)
]zy
0 for # - - u

The Brownian force F(~b), on a bead # is the result of the irregular colli-
sions between the solvent molecules and the bead. Instead of representing
the Brownian force by a randomly varying force, it is common in polymer
kinetic theory to use an averaged Brownian force,

F (b) -- -kBT ( 0 0r~


In
) (48)
489

As mentioned earlier, the origin of this expression can be understood within


the framework of the complete phase space theory [2,4]. Note that the
Maxwellian distribution has been used to derive equation (48).
The total potential energy r of the bead-spring chain is the sum of the
potential energy S of the elastic springs, and the potential energy E due to
the presence of excluded volume interactions between the beads. The force
F(r on a bead # due to the intramolecular potential energy r is given by,

F(~)_ 0r (49)

In addition to the various forces discussed above, the internal viscos-


ity force F (i~) has received considerable attention in literature [3,38,43]
though it appears not to have widespread acceptance. Various physical
reasons have been cited as being responsible for the internal viscosity force.
For instance, the hindrance to internal rotations due to the presence of en-
ergy barriers, the friction between two monomers on a chain that are close
together in space and have a non-zero relative velocity, and so on. The
simplest models for the internal viscosity force assume that it acts only
between neighbouring beads in a bead-spring chain, and depends on the
average relative velocities of these beads. Thus, for a bead # that is not at
the chain ends,

(-
F ,(iv) = 99 ( r . + l - - r # ) ( r . + l l 2
[r.+l -- r.
rt~)).[~#+1__~#]]

-- qO((r.- r._l)(r. [2rg_l)). [ [ / . _ 1"#-1]] (50)


I r u - ru_ 1
where ~p is the internal viscosity coefficient. A scaling theory for a more
general model that accounts for internal friction between arbitrary pairs of
monomers has also been developed [35].
The equation of motion for bead v can consequently be written as,
0 ln~
- ~ [ ~-i-~]]- v0 - tr A- ~ a u , . F (h) ] - kBT -4- F (~) + F (/') - 0(51)
# 0r~
Since F (h) - k , T (0 In ~ / 0 r u ) - F(r --uF(i~), equation (51) can be rear-
ranged to give,
1 0 ln~ (iv) )
[[~]] - vo + tr + ~ Z "~,~" (--kB T~ + + (52)
u 0r~
490

where -),,~ is the dimensionless diffusion tensor [2],


~,~ - 5u~ 1 + ( f t ~ (53)
By manipulating equation (52), it is possible to rewrite the equation of
motion in terms of the velocities of the center of mass r~ and the bead-
connector vectors Qk,
1
~-/'c]] - v0 + ~ . r~ + 0Qk0r+ r"k )
N ( u,u,kE Bku "/t,u " (kBT OoQkln____~ (54)

1
[[Oj]] - ~" Q j - -~ Z Ajk" ( kBT 0 In 9 0r f(iv) (55)
k
where, Bku is defined by, -Bk~ -- 5k+1,~- 5k~, the internal viscosity force,
fk(i~), in the direction of the connector vector Qk is,

Q'Q* (56)
- I I ]]
and the tensor -Ajb which accounts for the presence of hydrodynamic in-
teraction is defined by,

A.jk -- ~ B--j~,"ft,v-Bku -- Ajkl + ~(~j,k + ~r~j+l,k+l -- ~"~j,k+l - - ~-'~j+l,k) (57)


L,, it

Here, Ajk is the Rouse matrix,


2 for I J - k l - 0,
Ajk-- --1 forlj-kl-1, (58)
0 otherwise
In order to obtain the diffusion equation for a dilute solution of bead-
spring chains, the equation of motion derived here must be combined with
an equation of continuity.
3.2.3. The diffusion equation
The equation of Continuity or 'probability conservation', which states
that a bead-spring chain that disappears from one configuration must ap-
pear in another, has the form [2],
O~ 0
(59)
Ot
491

The independence of 9 from the location of the center of mass for homo-
geneous flows, and the result tr ~ - 0, for an incompressible fluid, can be
shown to imply that the equation of continuity can be written in terms of
internal coordinates alone as [2],
0r 0
0~ - - E o q / [ [ Qj] r (60)
J
The general diffusion equation which governs the time evolution of the
instantaneous configurational distribution function r in the presence of hy-
drodynamic interaction, arbitrary spring and excluded volume forces, and
an internal viscosity force given by equation (56), is obtained by substitut-
ing the equation of motion for ~-(~j~J from equation (55) into equation (60).
It has the form,
0r 0 1 0r ~(i~)
: - j
0qj qJ - Ek [0q 1)
0
kBT ~ OQj Aik" 0r -

(61)
+ ~ j,k OQk

Equations such as (61) are also referred to as Fokker-Planck or Smolu-


chowski equations in the literature. The diffusion equation (61) is the
most fundamental equation of the kinetic theory of dilute polymer solu-
tions since a knowledge of r for a flow field specified by to, would make it
possible to evaluate averages of various configuration dependent quantities
and thereby permit comparison of theoretical predictions with experimen-
tal observations.
The diffusion equation can be used to derive the time evolution equa-
tion of the average of any arbitrary configuration dependent quantity,
X( Q~, . . . , QN-~ ), by multiplying the left and right hand sides of equa-
tion (61) by X and integrating both sides over all possible configurations,
d (X) = E ( ~ ' Q j " OX). kBT E ( A j k " 0 l n r 9 OX
dt j OQj ~ j,k OQk OQj
- 0r r 0X
_ 1E<Ajk.[OQk+,k ]'OQj) (62)
Cj, k
Except for a situation where nearly all the important microscopic phe-
nomena are neglected, the diffusion equation (61) is unfortunately in gem
492

eral analytically insoluble. There have been very few attempts to di-
rectly solve diffusion equations with the help of a numerical solution proce-
dure [9,10]. In this context it is worth bearing in mind that what are usually
required are averages of configuration dependent quantities. However, in
general even averages cannot be obtained exactly by solving equation (62).
As a result, it is common in most molecular theories to obtain the averages
by means of various approximations.
In order to examine the validity of these approximations it is vitally
important to compare the approximate predictions of transport proper-
ties with the exact predictions of the models. One of the ways by which
exact numerical results may be obtained is by adopting a numerical pro-
cedure based on the mathematical equivalence of diffusion equations in
polymer configuration space and stochastic differential equations for the
polymer configuration [29]. Instead of numerically solving the analytically
intractable difusion equation for the distribution function, stochastic tra-
jectories can be generated by Brownian dynamics simulations based on a
numerical integration of the appropriate stochastic differential equation.
Averages calculated from stochastic trajectories (obtained as a solution of
the stochastic differential equations), are identical to the averages calcu-
lated from distribution functions (obtained as a solution of the diffusion
equations). It has now become fairly common for any new approximate
molecular theory of a microscopic phenomenon to establish the accuracy
of the results with the help of Brownian dynamics simulations. In this
chapter, while results of such simulations are cited, details of the develop:
ment of the appropriate stochastic differential equations are not discussed.
A comprehensive introduction to the development of stochastic differen-
tial equations which are equivalent to given diffusion equations for the
probability density in configuration space, can be found in the treatise by
Ottinger [29].

3.2.~. The stress tensor


The expression for the stress tensor in a dilute polymer solution was orig-
inally obtained by the use of simple physical arguments which considered
the various mechanisms that contributed to the flux of momentum across
an oriented surface in the fluid [2]. The major mechanisms considered were
the transport of momentum by beads crossing the surface, and the tension
in the springs that straddle the surface. These physical arguments help to
493

provide an intuitive understanding of the origin of the different terms in


the stress tensor expression. On the other hand, such arguments are diffi-
cult to pursue in the presence of complicated microscopic phenomena, and
there is uncertainity about the completeness of the final expression. An al-
ternative approach to the derivation of the expression for the stress tensor
has been to use more fundamental arguments that consider the complete
phase space of the polymeric fluid [2,41.
A very general expression for the polymer contribution to the stress ten-
sor, derived by adopting the complete phase space approach, for models
without constraints such as the bead-spring chain model, in the presence of
hydrodynamic interaction and an arbitrary intramolecular potential force,
is the modified Kramers expression [21,
~-P - n p E ((r~ - rc)F (r + ( N - 1)npknT 1 (63)
/J

When rewritten in terms of the internal coordinates of a bead-spring chain,


equation (63) assumes a form called the Kramers expression,
or (64)
7"p - -np ~ ( Q j ~ j > + (N - 1)npkuT1
3

It is important to note that the presence of internal viscosity has not


been taken into account in the phase space theories used to derive the
modified Kramers expression (63). When examined from the standpoint of
thermodynamic considerations, the proper form of the stress tensor in the
presence of internal viscosity appears to be the Gicsekus expression rather
than the Kramers expression [39]. Since predictions of models with internal
viscosity are not considered in this chapter, the Giesekus expression is not
discussed here.
In order to evaluate the stress tensor, for various choices of the potential
energy r it turns out that it is usually necessary to evaluate the second
moments of the bead connector vectors, <QjQk ). An equation that governs
the time evolution of the second moments can be obtained with the help
of equation (62). It has the form,
d 2kgT -- 1
dt( Q/Qk) -- ~ " ( QjQk >+ ( QjQk >. ~t + ( Ajk >- -~ ~m {( Qj
0r
+ f(~v)]. A.mk) + ( X j m ' [ ~ 0r + f2s) ]" qk> } (65)
494

The second moment equation (65), which is an ordinary differential equa-


tion, is in general not a closed equation for ( QjQk }, since it invoves higher
order moments on the right hand side.
Within the context of the molecular theory developed thus far, it is clear
that the prediction of the rheological properties of dilute polymer solutions
with a bead-spring chain model usually requires the solution of the second
moment equation (65). To date however, there are no solutions to the
general second moment equation (65) which simultaneously incorporates
the microscopic phenomena of hydrodynamic interaction, excluded volume,
non-linear spring forces and internal viscosity. Attempts have so far been
restricted to treating a smaller set of combinations of these phenomenon.
The simplest molecular theory, based on a bead-spring chain model, for
the prediction of the transport properties of dilute polymer solutions is the
Rouse model. The Rouse model neglects all the microscopic phenomenon
listed above, and consequently fails to predict many of the observed features
of dilute solution behavior. In a certain sense, however, it provides the
framework and motivation for all further improvements in the molecular
theory. The Rouse model and its predictions are introduced below, while
improvements in the treatment of hydrodynamic interactions alone are
discussed subsequently.

3.3. T h e R o u s e m o d e l
The Rouse model assumes that the springs of the bead-spring chain are
governed by a Hookean spring force law. The only solvent-polymer interac-
tions treated are that of hydrodynamic drag and Brownian bombardment.
The diffusion equation (61) with the effects of hydrodynamic interaction,
excluded volume and internal viscosity neglected, and with a Hookean
spring force law, has the form,

_ 0 H kBT 0 0r (66)
0 r _ _ ~ oqj" (~" Qj - --( ~-" Ajk Qk ) r + - - ~ ~ Ajk i)Qj'Oqk
Ot j k j,k

The diffusion equation (66) has an analytical solution since it is linear in


the bead-connector vectors. It is satisfied by a Gaussian distribution,

1 ~ Qj. (er_l)jk " Qk]


(Q1,..., QN-1) - Af(t) e x p [ - ~j,k (67)
495

where Af(t) is the normalisation factor, and the tensor O)k which uniquely
characterises the Gaussian distribution is identical to the second moment,
o ) k - <QjQk) (68)
Note that the tensors O)k are not symmetric, but satisfy the relation
~k -- Jkj- (Further information on linear diffusion equations and Gaussian
distributions can be obtained in the extended discussion in Appendix A
of [22]).
Since the intramolecular potential in the Rouse model is only due to the
presence of Hookean springs, it is straight forward to see that the Kramers
expression for the stress tensor r p, is given by,
v p -- - n p H ~ ajj + ( N - 1)npkBT 1 (69)
J
The tensors ~ j are obtained by solving the second moment equation (65),
which becomes a closed equation for the second moments when the Rouse
assumptions are made. It has the form,
H
d t ~ k -- ~ " ajk -- erjk . ~ t _ 2 k u T A j k 1 - - -
-
~-~[ a j m A m k -b A i m ermk ] (70)

Note that the solution of equation (70) also leads to the complete specifi-
cation of the Gaussian configurational distribution function r
A H o o k e a n d u m b b e l l model, which is the simplest example of a bead-
spring chain, is obtained by setting N = 2. It is often used for preliminary
calculations since its simplicity makes it possible to obtain analytical so-
lutions where numerical solutions are unavoidable for longer chains. For
such a model, substituting for ~rll in terms of r p from equation (69) into
equation (70), leads to following equation for the polymer contribution to
the stress tensor,
~'P + A s rP(1) -- - - n p k B T A H ~/ (71)

where ~'~1) -- d r P / d t - ~ . ~'P - ~'P . ~ i , is the convected time derivative [2]


of ~'P, and AH -- (4/4H) is a time constant. Equation (71) indicates that a
Hookean dumbbell model with the Rouse assumptions leads to a convected
Jeffreys model or Oldroyd-B model as the constitutive equation for a dilute
polymer solution. This is perhaps the simplest coarse-grained microscopic
model capable of reproducing some of the macroscopic rhcological proper-
tics of dilute polymer solutions.
496

In the case of a bead-spring chain with N > 2, it is possible to ob-


tain a similar insight into the nature of the stress tensor by introducing
n o r m a l coordinates. These coordinates help to decouple the connector vec-
tors Q 1 , . . . , QN-1, which are coupled to each other because of the Rouse
matrix.
The connector vectors are mapped to a new set of normal coordinates,
,
, QN-1 with the transformation,

Qj - ~ 1-IjkQk! (72)

where, I-Ijk are the elements of an orthogonal matrix with the property
(I-l-1)jk -- 1-Ikj , such that, ~ 1]mjYImk -- ~jk (73)
m
The orthogonal matrix Iljk , which will henceforth be referred to as the
Rouse orthogonal matrix, diagonalises the Rouse matrix Ajk,
IIjiAjkIIkl - at&t (74)
j,k
where, the Rouse eigenvalues at are given by at - 4sin2(17r/2N). The
elements of the Rouse orthogonal matrix are given by the expression,
2
(75)

The diffusion equation in terms of these normal coordinates, admits a


solution for the configurational distribution function of the form [2]
N-1
r (Q~,..-, Q~v-~) - 1-I Ok (Q~) (76)
k=l
As a consequence, the diffusion equation becomes uncoupled and can be
simplified to ( N - 1) diffusion equations, one for each of the Ck (Q~). Since
the Q~ are independent of each other, all the covariances (Q~Q~) with
j ~ k are zero, and only the ( i - 1) variances dj - (Q~Q~) are non-zero.
Evolution equations for the variances ~ can then be derived from these
uncoupled diffusion equations with the help of a procedure similar to that
used for the derivation of equation (65).
The stress tensor is given in terms of dj by the expression,
TP--Z WjP (77)
J
497

where,

T'jP - - n p H ~ + npkBT 1 (78)

On substituting for dj in terms of v P


j from equation (78) into the evolution
equation for dj, one obtains,

TjP -~- ~j TjP (1) - -


- -
-npkBT.~j ;',[ (79)

where, the relaxation times Aj are given by )~j - ( r aj). Consequently,


each of the "rjP satisfy an equation identical to equation (71) for the poly-
mer contribution to the stress tensor in a Hookean dumbbell model. The
Rouse model, therefore, leads to a constitutive equation that is a multi-
mode generalization of the convected Jeffreys or Oldroyd B model.
It is clear from above discussion that the process of transforming to nor-
mal coordinates enables one to derive a closed form expression for the stress
tensor, and to gain the insight that the Rouse chain with N beads has N
independent relaxation times which describe the different relaxation pro-
cesses in the chain, from the entire chain to successively smaller sub-chains.
It is straight forward to show that for large N, the longest relaxation times
)~j scale with chain length as N 2.
A few important transport property predictions which show the limita-
tions of the Rouse model are considered briefly below. It is worth noting
that since the Rouse model does not include the effect of excluded vol-
ume, its predictions are restricted to dilute solutions of polymers in theta
solvents. This restriction is infact applicable to all the models of hydrody-
namic interaction treated here.
In steady simple shear flow, with ~(t) given by equation (4), the three
independent material functions that characterise such flows are [2],

rip -- npkBT ~ /~j ; t~l -- 2npkBT ~ .kj ; q22 - 0 (80)


J J
It is clear that the Rouse model accounts for the presence of viscoelasticity
through the prediction of a nonzero first normal stress difference in simple
shear flow. However, it does not predict the nonvanishing of the second
normal stress difference, and the shear rate dependence of the viscometric
functions.
From the definition of intrinsic viscosity (1) and the fact that pp ~ N np,
it follows from equation (80) that for the Rouse model, It/]0 ~ N. This is
498

at variance with the experimental results discussed earlier, and displayed


in equation (11). It is also straight forward to see that the Rouse model
predicts that the characteristic relaxation time scales as the square of the
chain length, Ap ~ N 2.
In small amplitude oscillatory shear, ~;(t) is given by equation (6), and
expressions for the material functions G ~ and G" in terms of the relaxation
times )~j can be easily obtained [2]. In the intermediate frequency range,
where as discussed earlier, experimental results indicate that both G ~ and
G " scale as w2/3, the Rouse model predicts a scaling w 1/2 [17].
The translational diffusion coefficient D for a bead-spring chain at equi-
librium can be obtained by finding the average friction coefficient Z for the
entire chain in a quiescent solution, and subsequently using the Nernst-
Einstein equation, D - k n T Z -1 [2]. It can be shown that for the Rouse
model Z = ( N, ie. the total friction coefficient of the chain is a sum of the
individual bead friction coefficients. As a result, the Rouse model predicts
that the diffusion coefficient scales as the inverse of the molecular weight.
This is not observed in dilute solutions. Instead experiments indicate the
scaling depicted in equation (12).
The serious shortcomings of the Rouse model highlighted above have
been the motivation for the development of more refined molecular theories.
The scope of this chapter is restricted to reviewing recent advances in the
treatment of hydrodynamic interaction.

4. H Y D R O D Y N A M I C INTERACTION

Hydrodynamic interaction, as pointed out earlier, is a long range inter-


action between the beads which arises because of the solvent's capacity
to propagate one bead's motion to another through perturbations in its
velocity field. It was first introduced into framework of polymer kinetic
theory by Kirkwood and Riseman [14]. As we have seen in the devel-
opment of the general diffusion equation above, it is reasonably straight
forward to include hydrodynamic interaction into the framework of the
molecular theory. However, it renders the resultant equations analytically
intractable and as a result, various attempts have been made in order to
solve them approximately.
In this section, we review the various approximation schemes introduced
over the years. The primary test of an approximate model is of course
499

its capacity to predict experimental observations. The accuracy of the ap-


proximation however, can only be assessed by checking the proximity of
the approximate results to the exact numerical results obtained by Brow-
nian dynamics simulations. Finally, the usefulness of an approximation
depends on its computational intensity. The individual features and de-
ficiencies of the different approximations will be examined in the light of
these observations.
In the presence of hydrodynamic interaction, and with excluded volume
and internal viscosity neglected, a bead-spring chain with Hookean springs
has a configurational distribution function r that must satisfy the following
simplified form of the diffusion equation (61),

0r 0 H kBT 0 - 0r (81)
Ot - -~j9 OQj " (~" Qj - --~ y~
k
Ayk "Qk)r + r
Y~ OQj "Ajk " OQk
j,k

while the second moment equation (65) assumes the form,


d 2k.r (X k>
dt (QJQk> - ~" (QjQk> + (QjQk) . ~t + ......r
H
- -~-~ [(QjQm-Amk> + <Aim" QmQk>] (82)

Equation (82) is not a closed equation for the second moments since it
involves more complicated moments on the right hand side. This is the
central problem of all molecular theories which attempt to predict the
rheological properties of dilute polymer solutions and that incorporate hy-
drodynamic interaction. The different approximate treatments of hydro-
dynamic interaction, which are discussed roughly chronologically below,
basically reduce to finding a suitable closure approximation for the second
moment equation.
4.1. T h e Z i m m model
The Z i m m model was the first attempt at improving the Rouse model
by introducing the effect of hydrodynamic interaction in a preaveraged or
equilibrium-averaged form. The preaveraging approximation has been very
frequently used in polymer literature since its introduction by Kirkwood
and Riseman [14]. The approximation consists of evaluating the average of
the hydrodynamic tensor with the equilibrium distribution function (34),
and replacing the hydrodynamic interaction tensor Ajk, wherever it occurs
500

in the governing equations, with its equilibrium average Ajk. (Note that
the incorporation of the effect of hydrodynamic interaction does not alter
the equilibrium distribution function, which is still given by (34) for bead-
spring chains with Hookean springs.) The matrix Ajk is called the modified
Rouse matrix, and is given by,
--- 2 1 _ 1 ) (83)
- + h*(v/i j ki ~/iJ - k- 1[ v/Ij - k + 11
where, h* -- a v / ( H / z r k B T ) is the hydrodynamic interaction parameter. The
hydrodynamic interaction parameter is approximately equal to the ratio of
the bead radius to the equilibrium root mean square length of a single
spring of the bead-spring chain. This implies that h* < 0.5, since the
beads cannot overlap. Typical values used for h* are in the range 0.1 <
h* < 0.3 [22].
By including the hydrodynamic interaction in an averaged form, the dif-
fusion equation remains linear in the connector vectors, and consequently
is satisfied by a Gaussian distribution (67) as in the Rouse case. However,
the covariance tensors ~k are now governed by the set of differential equa-
tions (70) with the Rouse matrix Ajk replaced with the modified Rouse
matrix Ajk. Note that this modified second moment equation is also a
closed set of equations for the second moments.
As in the Rouse case, it is possible to simplify the solution of the Zimm
model by carrying out a diagonalisation procedure. This is achieved by
mapping the connector vectors to normal coordinates, as in (72), but in
this case the Zimm orthogonal matrix l-ljk, which diagonalises the modified
Rouse matrix,
N

IIjiAjkIIkl - al~il (84)


j,k
must be found numerically for N > 4. Here, al are the so called Zimm
eigenvalues. The result of this procedure is to render the diffusion equa-
tion solvable by the method of separation of variables. Thus, as in the
Rouse case, only the ( N - 1) transformed coordinate variances dj are non-
zero, and differential equations governing these variances can be derived
by manipulating the uncoupled diffusion equations.
The diagonalisation procedure enables the polymer contribution to the
stress tensor ~'P in the Zimm model to be expressed as a sum of partial
501

stresses ~'jP as in equation (77), but the "FjP n o w satisfy equation (79) with
the 'Rouse' relaxation times Aj replaced with 'Zimm' relaxation times Aj.
The Zimm relaxation times are defined by )~j - ( ~ / 2 H ~j).
From the discussion above, it is clear that the Zimm model differs from
the Rouse model only in the spectrum of relaxation times. As we shall
see shortly, this leads to a significant improvement in the prediction of
linear viscoelastic properties and the scaling of transport properties with
molecular weight in theta solvents. The Zimm model therefore establishes
unequivocally the importance of the microscopic phenomenon of hydrody-
namic interaction. On the other hand, it does not lead to any improvement
in the prediction of nonlinear properties, and consequently subsequent
treatments of hydrodynamic interaction have concentrated on improving
this aspect of the Zimm model.
By considering the long chain limit of the Zimm model, ie., N --+ c~, it
is possible to discuss the universal properties predicted by the model. The
various power law dependences of transport properties on molecular weight,
characterised by universal exponents, and universal ratios formed from
the prefactors of these dependences can be obtained. These predictions
are ideal for comparison with experimental data on high molecular weight
polymer solutions since they are parameter free. We shall discuss some
universal exponents predicted by the Zimm model below, while universal
ratios are discussed later in the chapter.
As mentioned above, the first noticeable change upon the introduction of
hydrodynamic interaction is the change in the relaxation spectrum. In the
long chain limit, the longest relaxation times/~j scale with chain length as
N 3/2 [29], whereas we had found earlier that the chain length dependence
of the longest relaxation times in the Rouse model was N 2.
In steady simple shear flow, the Zimm model like the Rouse model, fails to
predict the experimentally observed occurance of non-zero second normal
stress differences and the experimentally observed shear rate dependence
of the viscometric functions. It does however lead to an improved predic-
tion of the scaling of the zero shear rate intrinsic viscosity with molecular
weight, [~]0 ~ N 1/2. This prediction is in agreement with experimental re-
sults for the Mark-Houwink exponent in theta solvents (see equation (11)).
As with the longest relaxation times, the characteristic relaxation time
Ap ~ N 3/2.
502

In small amplitude oscillatory shear, the Zimm model predicts that the
material functions G ~ and G " scale with frequency as w 2/3 in the interme-
diate frequency range. This is in exceedingly good agreement with experi-
mental results [2,17].
The translational diffusion coefficient D for chainlike molecules at equilib-
rium, with preaveraged hydrodynamic interaction, was originally obtained
by Kirkwood [14]. Subsequently, several workers obtained a correction to
the Kirkwood diffusion coefficient for the Zimm model [21]. The exact re-
sults differ by less than 2% from the Kirkwood value for all values of the
chain length and h*. Interestingly, three different approaches to obtaining
the diffusion coefficient, namely, the Nernst-Einstein equation, the calcu-
lation of the mean-square displacement caused by Brownian forces, and
the study of the time evolution of concentration gradients, lead to identi-
cal expressions for the diffusion coefficient [21]. In the limit of very long
chains, it can be shown that D ~ N -1/2. The Zimm model therefore gives
the correct dependence of translational diffusivity on molecular weight in
theta solvents.
The Zimm result for the translational diffusivity has been traditionally
interpreted to mean that the polymer coil in a theta solvent behaves like a
rigid sphere, with radius equal to the root mean square end-to-end distance.
This follows from the fact that the diffusion coefficient for a rigid sphere
scales as the inverse of the radius of the sphere, and in a theta solvent,
( r2)eq scales with chain length as N. The solvent inside the coil is believed
to be dragged along with the coil, and the inner most beads of the bead-
spring chain are considered to be shielded from the velocity field due to
the presence of hydrodynamic interaction [44,17]. This intuitive notion
has been used to point out the difference between the Zimm and the Rouse
model, where all the N beads of the polymer chain are considered to be
exposed to the applied velocity field. Recently, by explicitly calculating
the velocity field inside a polymer coil in the Zimm model, C)ttinger [30]
has shown that the solvent motion inside a polymer coil is different from
that of a rigid sphere throughout the polymer coil, and that shielding from
the velocity field occurs only to a certain extent.

4.2. The consistent averaging approximation


The first predictions of shear thinning were obtained when hydrodynamic
interaction was treated in a more precise manner than that of preaverag-
503

ing the hydrodynamic interaction tensor. In order to make the diffusion


equation (81) linear in the connector vectors, as pointed out earlier, it is
necessary to average the hydrodynamic interaction tensor. However, it is
not necessary to preaverage the hydrodynamic interaction tensor with the
equilibrium distribution. On the other hand, the average can be carried
out with the non-equilibrium distribution function (67). The linearity of
the diffusion equation ensures that its solution is a Gaussian distribution.
Ottinger [20,22] suggested that the hydrodynamic interaction tensor occur-
ing in the diffusion equation be replaced with its non-equilibrium average.
Since it is necessary to know the averaged hydrodynamic interaction tensor
in order to find the non-equilibrium distribution function, both the aver-
aged hydrodynamic interaction tensor and the non-equilibrium distribution
function must be obtained in a self-consistent manner.
Several years ago, Fixman [11] introduced an iterative scheme (begin-
ning with the equilibrium distribution function), for refining the distribu-
tion function with which to carry out the average of the hydrodynamic o.

interaction tensor. The self-consistent scheme of Ottinger is recovered


if the iterative procedure is repeated an infinite number of times. How-
ever, Fixman carried out the iteration only upto one order higher than the
preaveraging stage.
The average of the hydrodynamic interaction tensor evaluated with the
Gaussian distribution (67)is an ( N - 1) x ( N - 1) matrix with tensor
components, Ajk, defined by,

-- [ H(&j,k) + H(&j+l,k+l)
Ajk -- Ajk I + x/~h* ~/IJ 5' k[
x/'lJ- kl
_ I (8s)
v/l)- k+ iI
where the tensors &,~ are given by,

1 H max(#w)-I
" --^ -- Z ~k (86)
O'l~V- ~ v 0"'1" II~- l/] ]gBr j,k=min(#w)

and the function of the second moments, H(er) is,

3 1 (l_kk 1
H(er) - 2(2rr)3/2 f e x p ( - k - o ' . k) (87)
504

Note that the convention H(~jj)/0 - 0 has been adopted in equation (85)
above.
The self-consistent closure approximation therefore consists of replacing
the hydrodynamic interaction tensor Ayk in equation (82) with its non-
equilibrium average A---jk. As in the earlier approximations, this leads to a
system of ( N - 1) 2 coupled ordinary differential equations for the compo-
nents of the covariance matrix a)k. Their solution permits the evaluation
of the stress tensor through the Kramers expression (69), and as a conse-
quence all the relevant material functions.
Viscometric functions in steady simple shear flows were obtained by
Ottinger [20] for chains with N ~ 25 beads, while material functions in
start-up of steady shear flow, cessation of steady shear flow, and stress re-
laxation after step-strain were obtained by Wedgewood and Ottinger [41]
for chains with N < 15 beads. The latter authors also include consistently-
averaged FENE springs in their model.
Shear rate dependent viscometric functions, and a nonzero positive sec-
ond normal stress difference are predicted by the self-consistent averaging
approximation; a marked improvement over the predictions of the Zimm
model. Both the reduced viscosity and the reduced first normal stress
difference initially decrease with increasing shear rate. However, for long
enough chains, they begin to rise monotonically at higher values of the
reduced shear rate ~. This rise is a consequence of the weakening of hy-
drodynamic interaction in strong flows due to an increase in the separation
between the beads of the chain. With increasing shear rate, the material
functions tend to the shear rate independent Rouse values, which (for long
enough chains), are higher than the zero shear rate consistently-averaged
values. The prediction of shear thickening behavior is not in agreement
with the shear thinning that is commonly observed experimentally. How-
ever, as mentioned earlier, some experiments with very high molecular
weight systems seem to suggest the existence of shear thinning followed by
shear thickening followed again by shear thinning as the shear rate is in-
creased. While only shear thickening at high shear rates is predicted with
Hookean springs, the inclusion of consistently-averaged FENE springs in
the model leads to predictions which are qualitatively in agreement with
these observations, with the FENE force becoming responsible for the shear
thinning at very high shear rates [41,15].
505

The means of examining the accuracy of various approximate treatments


of hydrodynamic interaction was established when the problem was solved
exactly with the help of Brownian dynamics simulations with full hydrody-
namic interaction included [46,48]. These simulations reveal that while the
predictions of the shear rate dependence of the viscosity and first normal
stress difference by the self-consistent averaging procedure are in qualita-
tive agreement with the Brownian dynamics simulations, they do not agree
quantitatively. Further, in contrast to the consistent-averaging prediction,
at low shear rates, a negative value for the second normal stress difference
is obtained. As noted earlier, the sign of the second normal stress difference
has not been conclusively established [3].
The computational intensity of the consistent-averaging approximation
leads to an upper bound on the length of chain that can be examined. As
a result, it is not possible to discuss the universal shear rate dependence
of the viscometric functions predicted by it. On the other hand, it is
possible to come to certain general conclusions regarding the nature of
the stress tensor in the long chain limit, and to predict the zero shear
rate limit of certain universal ratios [20]. Thus, it is possible to show
the important result that the polymer contribution to the stress tensor
depends only on a length scale and a time scale, and not on the strength
of the hydrodynamic interaction parameter h*. In the long chain limit,
h* can be absorbed into the basic time constant, and it does not occur
in any of the non-dimensional ratios. Indeed this is also true of the finite
extensibility parameter b, which can also be shown to have no influence
on the long chain rheological properties [22]. The long chain limit of the
consistent-averaging approximation is therefore a parameter free model.
It is possible to obtain an explicit representation of the modified Kramers
matrix for infinitely long chains by introducing continuous variables in
place of discrete indices [20]. This enables the analytical calculation of
various universal ratios predicted by the consistent-averaging approxima-
tion. These predictions are discussed later in this chapter. However, two
results are worth highlighting here. Firstly, it can be shown explicitly that
the leading order corrections to the large N limit of the various univer-
sal ratios are of order (1/v/N), and secondly, there is a special value of
h* = 0.2424..., at which the leading order corrections are of order ( l / N ) .
These results have proven to be very useful for subsequent numerical explo-
506

ration of the long chain limit in more accurate models of the hydrodynamic
interaction.
Short chains with consistently-averaged hydrodynamic interaction, as
noted earlier, do not show shear thickening behavior; this aspect is re-
vealed only with increasing chain length. Furthermore, it is not clear with
the kind of chain lengths that can be examined, whether the minimum in
the viscosity and first normal stress curves continue to exist in the long
chain limit [20]. The examination of long chain behavior is therefore im-
portant since aspects of polymer solution behavior might be revealed that
are otherwise hidden when only short chains are considered. The introduc-
tion of the decoupling approximation by Magda, Larson and Mackay [18]
and Kishbaugh and McHugh [15] made the examination of the shear rate
dependence of long chains feasible. The dccoupling approximation retains
the accuracy of the self-consistent averaging procedure, but is much more
computationally efficient.

4.3. T h e decoupling a p p r o x i m a t i o n
The decoupling approximation introduced by Magda et al. [18] and Kish-
baugh and McHugh [15] (who use FENE springs in place of the Hookean
springs of Magda et al.) consists of extending the 'diagonalise and de-
couple' procedure of the Rouse and Zimm theories to the case of the self
consistently averaged theory. They first transform the connector vectors
Qj to a new set of coordinates Q~ using the timc-invariant Rouse orthog-
onal matrix Iljk. (Kishbaugh and McHugh also use the Zimm orthogonal
matrix). The same orthogonal matrix Iljk is then assumed to diagonalise
the matrix of tensor components -Ajk. While the process of diagonalisa-
tion was exact in the Rouse and Zimm theories, it is an approximation in
the case of the decoupling approximation. It implies that even in the self
consistently averaged theory the diffusion equation can be solved by the
method of separation of variables, and only the ( N - 1 ) transformed coordi-
nate variances ~ are non-zero. The differential equations governing these
variances can be derived from the uncoupled diffusion equations and solved
numerically. The appropriate material functions are then obtained using
the Kramers expression in terms of the transformed coordinates, namely,
equations (77)and (78).
The decrease in the number of differential equations to be solved, from
( i - 1) 2 for the covariances ~jk to ( N - 1) for the variances dj, is suggested
507

by Kishbaugh and McHugh as the reason for the great reduction in com-
putational time achieved by the decoupling approximation. Prak.~sh and
Ottinger [33] discuss the reasons why this argument is incomplete, and
point out the inconsistencies in the decoupling procedure. Furthermore,
since the results are only as accurate as the consistent averaging approxi-
mation, the decoupling approximation is not superior to the consistent av-
eraging method. However, these papers are important since the means by
which a reduction in computational intensity may be achieved, without any
significant sacrifice in accuracy, was first proposed in them. Further, the
persistence of the minimum in the viscosity and first normal stress curves
even for very long chains, and the necessity of including FENE springs in
order to generate predictions in qualitative agreement with experimental
observations in high molecular weight systems, is clearly elucidated in these
papers.

4.4. The Gaussian approximation


The closure problem for the second moment equation is solved in the
preaveraging assumption of Zimm, and in the self consistent averaging
..

method of Ottinger, by replacing the tensor Ajk with an average. As a


result, fluctuations in the hydrodynamic interaction are neglected. The
Gaussian approximation [24,46,42,48] makes no assumption with regard to
the hydrodynamic interaction, but assumes that the solution of the diffu-
sion equation (81) may be approximated by a Gaussian distribution (67).
Since all the complicated averages on the right hand side of the second
moment equation (82) can be reduced to functions of the second moment
with the help of the Gaussian distribution, this approximation makes it a
closed equation for the second moments.
The evolution equation for the covariances O)k is given by,
H
d 2kBT A-jk ~ [O'jm"--Arnk+ Ajm" O'mk]
H H
- E 9 9 + (88)
( kBT m,l,p
where, the ( N - 1) 2 x ( N - 1) 2 matrix with fourth rank tensor components,
Flp,jk, is defined by,
3v~ h* [ O(j, l, p, k) K(~j,k) + O(j + 1, l, p, k + 1) g(~+l,k+,)
= 4 [ lJ- kl
508

O(j,l,p,k+ 1) K(~,k+l) O(j+ 1,1,p,k)K(~+l,k)]


] (89)
v/Ij - k - 1]3 - ~ i J - k + 1]3
while the function K(er) is defined by the equation,

U(~r) - - 2 f dk k-~lk _(1


(27r)3/2 kk
~-) k exp(-~l k.~r. k) (90)

The function 0(j, l, p, k) is unity if 1 and p lie between j and k, and zero
otherwise,

O(j,l p,k) - {1 i f j ~ _ l , p < k or k ~_ l,p < j (91)


' 0 otherwise
The convention g ( ~ j j ) / 0 - 0, has been adopted in equation (89). Both
the hydrodynamic interaction functions H(~r) and g(~r) can be evaluated
analytically in terms of elliptic integrals. The properties of these functions
oo

are discussed in great detail in the papers by Ottinger and coworkers [22,
24,25,48,47].
All the approximations discussed earlier (with the exception of the decou-
pling approximation) can be derived by a process of succesive simplification
of the explicit results for the Gaussian approximation given above. The
equations that govern the self consistently averaged theory can be obtained
by dropping the last term in equation (88), which accounts for the presence
of fluctuations. Replacing H(~r) by 1 in these truncated equations leads to
the governing equations of the Zimm model, while setting h* - 0 leads to
the Rouse model.
Material functions predicted by the Gaussian approximation in any arbi-
trary homogeneous flow may be obtained by solving the system of ( N - 1) 2
coupled ordinary differential equations for the components of the covari-
ance matrix O)k (88). Small amplitude oscillatory shear flows and steady
shear flow in the limit of zero shear rate have been examined by Ottinger
[24] for chains with N ~_ 30 beads, while Zylka [48] has obtained the ma-
terial functions in steady shear flow for chains with N _~ 15 beads and
compared his results with those of Brownian dynamics simulations (the
comparison was made for chains with N - 12 beads).
The curves predicted by the Gaussian approximation for the storage and
loss modulus, G ~and G", as a function of the frequency w, are nearly indis-
tinguishable from the predictions of the Zimm theory, suggesting that the
509

Zimm approximation is quite adequate for the prediction of linear visco-


elastic properties. There is, however, a significant difference in the pre-
diction of the relaxation spectrum. While the Zimm model predicts a set
of ( N - 1) relaxation times with equal relaxation weights, the Gaussian
approximation predicts a much larger set of relaxation times than the num-
ber of springs in the chain, with relaxation weights that are different and
dependent on the strength of the hydrodynamic interaction [28]. These
results indicate that entirely different relaxation spectrum lead to similar
curves for G ~ and G", and calls into question the common practice of ob-
taining the relaxation spectrum from experimentally measured curves for
G' and G" (see also the discussion in [34]).
The zero shear rate viscosity and first normal stress difference predicted
by the Gaussian approximation are found to be smaller than the Zimm
predictions for all chain lengths. By extrapolating finite chain length re-
~

sults to the infinite chain limit, Ottinger has shown that this reduction
is by a factor of 72% - 73%, independent of the strength of the hydro-
dynamic interaction parameter. Other universal ratios predicted by the
Gaussian approximation in the limit of zero shear rate are discussed later
in the chapter.
A comparison of the predicted shear rate dependence of material func-
tions in simple shear flow with the results of Brownian dynamics simu-
lations reveals that of all the approximate treatments of hydrodynamic
interaction introduced so far, the Gaussian approximation is the most ac-
curate [48]. Indeed, at low shear rates, the negative second normal stress
difference predicted by the Gaussian approximation is in accordance with
the simulations results.
Inspite of the accuracy of the Gaussian approximation, its main drawback
is its computational intensity, which renders it difficult to examine chains
with large values of N. Apart from the need to examine long chains for
the reason cited earlier, it also necessary to do so in order to obtain the
universal predictions of the model. A recently introduced approximation
which enables the evaluation of universal viscometric functions in shear
flow is discussed in the section below. Before doing so, however, we first
discuss the significant difference that a refined treatment of hydrodynamic
interaction makes to the prediction of translational diffusivity in dilute
polymer solutions.
510

The correct prediction of the scaling of the diffusion coefficient with


molecular weight upon introduction of pre-averaged hydrodynamic inter-
action in the Zimm model demonstrates the significant influence that hy-
drodynamic interaction has on the translational diffusivity of the macro-
molecule. While the pre-averaging assumption appears adequate at equi-
librium, it predicts a shear rate independent scalar diffusivity even in the
presence of a flow field. On the other hand, both the improved treatments
of hydrodynamic interaction, namely, consistent averaging and the Gaus-
sian approximation, reveal that the translational diffusivity of a Hookean
dumbbell in a flowing homogeneous solution is described by an anisotropic
diffusion tensor which is flow rate dependent [23,26,27]. Indeed, unlike in
the Zimm case, the three different approaches mentioned earlier for calcu-
lating the translational diffusivity do not lead to identical expressions for
the diffusion tensor [23,26]. Insight into the origin of the anisotropic and
flow rate dependent behavior of the translational diffusivity is obtained
when the link between the polymer diffusivity and the shape of the poly-
mer molecule in flow [12] is explored [31,32]. It is found that the solvent
flow field alters the distribution of mass about the centre of the dumb-
bell. As a consequence, the dumbbell experiences an average friction that
is anisotropic and flow rate dependent. The discussion of the influence
of improved treatments of hydrodynamic interaction on the translational
diffusivity has so far been confined to the Hookean dumbbell model. This
is because the concept of the center of resistance, which is very useful for
simplifying calculations for bead-spring chains in the Zimm case, cannot
be employed in these improved treatments [23].

4.5. T h e twofold n o r m a l a p p r o x i m a t i o n
The twofold normal approximation borrows ideas from the decoupling
approximation of Magda et al. [18] and Kishbaugh and McHugh [15] in or-
der to reduce the computational intensity of the Gaussian approximation.
As in the case of the Gaussian approximation, and unlike in the case of the
consistent-averaging and decoupling approximations where it is neglected,
fluctuations in the hydrodynamic interaction are included. In a sense, the
twofold normal approximation is to the Gaussian approximation, what the
decoupling approximation is to the consistent-averaging approximation.
The computational efficiency of the decoupling approximation is due both
to the reduction in the set of differential equations that must be solved in
511

order to obtain the stress tensor, and to the procedure that is adopted to
solve them [33]. These aspects are also responsible for the computational
efficiency of the twofold normal approximation. However, the derivation of
the reduced set of equations in the twofold normal approximation is signifi-
cantly different from the scheme adopted in the decoupling approximation;
it is more straight forward, and avoids the inconsistencies that are present
in the decoupling approximation.
Essentially the twofold normal approximation, (a) assumes that the con-
figurational distribution function r is Gaussian, (b) uses the Rouse or the
Zimm orthogonal matrix 1-Ijk to map Qj to 'normal' coordinates Q~, and
(c) assumes that the covariance matrix ~k is diagonalised by the same
orthogonal matrix, ie. Ej, k 1-Ijp~k H k q - (Q~Q~q) - dpSm. This leads to
the following equations for the ( N - 1) variances dj,
d 2kBT Aj - H [dj . Aj + Aj . dj]

H H
A]
k
where, Aj = Ajj are the diagonal tensor components of the matrix Ajk,
Ajk- ~ 1-Ilj Alp l-Ipk (93)
1,p
and the matrix ~jk is given by,
Ajk -- ~ 1-Iljl]pk rip,ran l-Imj link (94)
l, m, n, p
In equations (93) and (94), the tensors Ajk and ~rr given by equa-
tions (85) and (89), respectively. However, the argument of the hydrody-
namic interaction functions is now given by,
1 H max(#,v)- 1
^ -- E E Him nkm o'~ (95)
a'tw Itt -- v ! kBT j,k=min(#,v) m

The decoupling approximation is recovered from the twofold normal ap-


proximation when the last term in equation (92), which accounts for fluc-
tuations in hydrodynamic interaction, is dropped. Thus the two different
routes for finding governing equations for the quantities ~j lead to the
..

same result. However, Prakash and Ottinger [33] have shown that this is
512

in some sense a fortuitous result, and indeed the key assumption made in
the decoupling approximation regarding the diagonalisation of Ajk is not
tenable. The Zimm model in terms of normal modes may be obtained from
equation (92) by dropping the last term, and substituting Ajk in place of
A--jk. Of course the Zimm orthogonal matrix must be used to carry out the
diagonalisation in equation (93). The diagonalised Zimm model reduces
to the diagonalised Rouse model upon using the Rouse orthogonal matrix
and on setting h* - 0.
The evolution equations (92) have been solved to obtain the zero shear
rate properties for chains with N ~_ 150, when the Zimm orthogonal matrix
is used for the purpose of diagonalisation, and for chains with N _ 400,
when the Rouse orthogonal matrix is used. Viscometric functions at finite
shear rates in simple shear flows have been obtained for chains with N ~_
100 [33]. The results are very close to those of the Gaussian approximation;
this implies that they must also lie close to the results of exact Brownian
dynamics simulations. The reasons for the reduction in computational
intensity of the twofold normal approximation are discussed in some detail
in [33]. The most important consequence of introducing the twofold normal
approximation is that rheological data accumulated for chains with as many
as 100 beads can be extrapolated to the limit N -+ c~, and as a result,
universal predictions may be obtained.

4.6. Universal properties in theta solvents


One of the most important goals of examining the influence of hydrody-
namic interactions on polymer dynamics in dilute solutions is the calcula-
tion of universal ratios and master curves. These properties do not depend
on the mechanical model used to represent the polymer molecule. Con-
sequently, they reflect the most general consequence of the way in which
hydrodynamic interaction has been treated in the theory. They are also
the best means to compare theoretical predictions with experimental ob-
servations since they are parameter free.
There appear to be two routes by which the universal predictions of
models with hydrodynamic interaction have been obtained so far, namely,
by extrapolating finite chain length results to the limit of infinite chain
length where the model predictions become parameter free, and by using
renormalisation group theory methods.
In the former method, there are two essential requirements. The first is
513

that rheological data for finite chains must be generated for large enough
values of N so as to be able to extrapolate reliably, ie. with small enough
error, to the limit N --4 oc. The second is that some knowledge of the lead-
ing order corrections to the infinite chain length limit must be obtained in
order to carry out the extrapolation in an efficient manner. It is clear from
the discussion of the various approximate treatments of hydrodynamic
interaction above that it is possible to obtain universal ratios in the zero
shear rate limit in all the cases. Four universal ratios that are frequently
used to represent the rheological behavior of dilute polymer solutions in
the limit of zero shear rate are [29],
- lim 77P'~
U,p~ nkBTAI~TP'~ U,R n-~On~Ts(nTrR3g/3)
-

Ur
_ nkBT~l,o
2 Ur ~2,0 (96)
rl~,0 ~,0
where, )~1 is the longest relaxation time, and R 9 is the root-mean-square
radius of gyration at equilibrium. With regard to the leading order cor-
rections to these ratios, it has been possible to obtain them explicitly only
in the consistently-averaged case [20]. In both the Gaussian approxima-
tion and the twofold normal approximation it is assumed that the leading
order corrections are of the same order, and extrapolation is carried out
numerically by plotting the data as a function of (1/v/N). Because of their
computational intensity, it is not possible to to obtain the universal shear
rate dependence of the viscometric functions predicted by the consistent-
averaging and Gaussian approximations. However, it is possible to obtain
these master curves with the twofold normal approximation.
Table 1 presents the prediction of the universal ratios (96) by the various
approximate treatments. Miyaki et al. [19] have experimentally obtained
a value of UoR -- 1.49 (6) for polystyrene in cyclohexane at the theta tem-
perature. Figure 1 displays the viscometric functions predicted by the two
fold normal approximation. The coincidence of the curves for the different
values of h* indicate the parameter free nature of these results. Divergence
of the curves at high shear rates implies that the data accumulated for
chains with N _< 100 is insufficient to carry out an accurate extrapolation
at these shear rates. The incorporation of the effect of hydrodynamic inter-
action into kinetic theory clearly leads to the prediction of shear thickening
at high shear rates even in the long chain limit.
514

Table 1
Universal ratios in the limit of zero shear rate. The exact Zimm values
and the Gaussian approximation (GA) values for U,TR and Ur are repro-
duced from [29], the exact consistent-averaging values from [20], and the
renormalisation group (RG) results from [25]. The twofold normal approx-
imation values with the Zimm orthogonal matrix (TNZ) and the remaining
GA values are reproduced from [33]. Numbers in parentheses indicate the
uncertainity in the last figure.

Zimm 2.39 1.66425 0.413865 0.0


CA 2.39 1.66425 0.413865 0.010628
GA 1.835 (1) 1.213 (3) 0.560 (3) -0.0226 (5)
RG - 1.377 0.6096 -0.0130
TNZ 1.835 (1) 1.210 (2) 0.5615 (3) -0.0232 (1)

In both table 1 and figure 1, the results of renormalisation group calcu-


lations (RG) are also presented [25,47]. As mentioned earlier, the renor-
malisation group theory approach is an alternative procedure for obtaining
universal results. It is essentially a method for refining the results of a low-
order perturbative treatment of hydrodynamic interaction by introducing
higher order effects so as to remove the ambiguous definition of the bead
size. All the infinitely many interactions for long chains are brought in
through the idea of self-similarity. It is a very useful procedure by which
a low-order perturbation result, which can account for only a few interac-
tions, is turned into something meaningful. However, systematic results
can only be obtained near four dimensions, and one cannot estimate the
errors in three dimensions reliably. The Gaussian and twofold normal ap-
proximations on the other hand are non-perturbative in nature, and are
essentially 'uncontrolled' approximations with an infinite number of higher
order terms.
It is clear from the figures that the two methods lead to significantly dif-
ferent results at moderate to high shear rates. A minimum in the viscosity
and first normal stress difference curves is not predicted by the renormalisa-
tion group calculation, while the twofold normal approximation predicts a
515

.0 . . . . . . . . | . . . . . . . . !
0 ~ , .. . . . . . , . . . . . . .-., 9 9

3.5 h =0.15 . . . . h*=0.]5


o h* = 0.25 J 16 o h* = 0.25
3.0
q ....... RG
~o 12 ....... RG
w 2.5

2.0 8 "~
1.5
4 j
1.0[ . . . . . . . . . . . . . . . . . . . . . . . . . . 9 t -, " - r ,~.

0 l---^ :=~-'-~: ...........


10 -1 10 o 101 10 -1 10 o 101

0.005 .........

. O 0 O " " ~ ~ ' " " ~ ~

-0.005
9-
"~
r
-0.010
9-
-0.015

-0.020 ~ o h* = 0.25
...... RG
-0.025 . . . . . . . . . . . . . .
1(,-' " i0 ~ 10'
p
Figure 1. Universal viscometric functions in theta solvents. Reproduced
from [33].

small decrease from the zero shear rate value before the monotonic increase
at higher shear rates. The good comparison with the results of Brownian
dynamics simulations for short chains indicates that the twofold normal ap-
proximation is likely to be more accurate than the renormalisation group
calculations.

5. C O N C L U S I O N S

This chapter discusses the development of a unified basis for the treat-
ment of non-linear microscopic phenomena in molecular theories of dilute
polymer solutions and reviews the recent advances in the treatment of
hydrodynamic interaction. In particular, the successive refinements which
ultimately lead to the prediction of universal viscometric functions in theta
solvents have been highlighted.
516

REFERENCES
~
R. B. Bird, R. C. Armstrong and O. Hassager, Dynamics of Polymeric
Liquids, Vol. 1, Fluid Mechanics, 2nd edn., John Wiley, 1987.
.
R. B. Bird, C. F. Curtiss, R. C. Armstrong and O. Hassager, Dynamics
of Polymeric Liquids, Vol. 2, Kinetic Theory, 2nd edn., John Wiley,
1987.
.
R. B. Bird and H. C. (3ttinger, Annu. Rev. Phys. Chem., 43, (1992)
371-406.
.
C. F. Curtiss and R. B. Bird, Adv. Polymer Sci., 125, (1996), 1-101.
5. C. F. Curtiss, R. B. Bird and O. Hassager, Adv. Chem. Phys., 35 (1976)
31-117.
.
J. des Cloizeaux and G. Jannink, Polymers in Solution, Their Modelling
and Structure, Oxford Science Publishers, 1990.
.
P-G. de Gennes, Scaling Concepts in Polymer Physics, Cornell Univer-
sity Press, 1979.
.
M. Doi and S. F. Edwards, The Theory of Polymer Dynamics, Claren-
don Press, Oxford, 1986
.
X. J. Fan, J. Non-Newtonian Fluid Mech., 17 (1985) 125-144.
10. X. J. Fan, J. Chem. Phys., 85 (1986) 6237-6238.
11. M. Fixman, J. Chem. Phys., 45 (1966) 785-792, 793-803.
12. D. A. Hoagland and R. K. Prud'homme, J. Non-Newtonian Fluid
Mech., 27 (1988) 223-243.
13. J. G. Kirkwood, Macromolecules, Gordon and Breach, New York, 1967
14. J. G. Kirkwood and J. Riseman, J. Chem. Phys., 16 (1948) 565-573.
15. A. J. Kishbaugh and A. J. McHugh, J. Non-Newtonian Fluid Mech.,
34 (1990) 181-206.
16. H. A. Kramers, Physica, 11 (1944) 1-19.
17. R. G. Larson, Constitutive Equations for Polymer Melts and Solutions,
Butterworths, Boston, 1988.
18. J. J. Magda, R. G. Larson and M. E. Mackay, J. Chem. Phys., 89
(1988) 2504-2513.
19. Y. Miyaki, Y. Einaga, H. Fujita and M. Fukuda, Macromolecules, 13
(1980) 588.
20. H. C. Ottinger, J. Chem. Phys., 86 (1987) 3731-3749.
21. H. C. Ottinger, J. Chem. Phys., 87 (1987) 3156-3165.
22. H. C. Ottinger, J. Non-Newtonian Fluid Mech., 26 (1987) 207-246.
517

23. H. C. ()ttinger, J. Chem. Phys., 87 (1987) 6185-6190.


24. H. C. Ottinger, J. Chem. Phys., 90 (1989) 463-473.
**

25. H. C. Ottinger and Y. Rabin, J. Non-Newtonian Fluid Mech., 33 (1989)


53-93.~

26. H. C. Ottinger, AIChE J., 35 (1989) 279-285.


o~

27. H. C. Ottinger, Coll. Polym. Sci., 267 (1989) 1-8.


**

28. H. C. Ottinger and W. Zylka, J. Rheol., 36 (1992) 885-910.


29. H. C. Ottinger, Stochastic Processes in Polymeric Fluids, Springer-
Verlag, 1996.
~176

30. H. C. Ottinger, Rheologica Acta, 35 (1996) 134-138.


31. J. R. Prakash and R. A. Mashelkar, J.Chem.Phys., 95 (1991) 3743-
3748.
32. J. R. Prakash and R. A. Mashelkar, J. Rheol., 36 (1992) 789-805.
33. J. R. Prakash and H. C. Ottinger, J. Non-Newtonian Fluid Mech., 71
(1997) 245-272.
34. J. R. Prakash, in: M. J. Adams, R. A. Mashelkar, J. R. A. Pearson and
A. R. Rennie (Eds.), Dynamics of Complex Fluids, Imperial College
Press-The Royal Society, in press, (1997).
.o

35. Y. Rabin and H. C. Ottinger, Euro-Phys. Lett., 13 (1990) 423-428.


36. P. E. Rouse, J. Chem. Phys., 21 (1953) 1272-1280.
37. J. D. Schieber and H. C. Ottinger, J. Chem. Phys., 89 (1988) 6972-
6981.
38. J. D. Schieber, J. Rheol., 37 (1993) 1003-1027.
39. J. D. Schieber and H. C. Ottinger, J. Rheol., 38 (1994) 1909-1924.
40. G. Strobl, The Physics of Polymers, Springer-Verlag, 1996.
41. L. E. Wedgewood and H. C. Ottinger, J. Non-Newtonian Fluid Mech.,
27 (1988)245-264.
42. L. E. Wedgewood, J. Non-Newtonian Fluid Mech., 31 (1989) 127-142.
43. L. E. Wedgewood, Rheol. Acta, 32 (1993) 405-417.
44. H. Yamakawa, Modern Theory of Polymer Solutions, Harper ~r Row,
New York, 1971.
45. B. H. Zimm, J. Chem. Phys., 24 (1956) 269-281.
46. W. Zylka and H. C. Ottinger, J. Chem. Phys., 90 (1989) 474-480.
47. W. Zylka and H. C. ()ttinger, Macromolecules, 24 (1991) 484-494.
48. W. Zylk~, J. Chem. Phys., 94 (1991) 4628-4636.
519

CONSTITUTIVE EQUATIONS FOR VISCOELASTIC LIQUIDS"


FORMULATION, ANALYSIS AND COMPARISON WITH DATA

A. I. Leonov

Department of Polymer Engineering, The University of Akron,


Akron, OH 44325 - 0301, USA

1. INTRODUCTION

Viscoelastic liquids are capable of accumulating large recoverable strains in


flow. This puts the media in an intermediate position between liquids and solids
and makes their rheological behavior very complicated. Though the particular
properties of elastic liquids have found many useful applications in polymer
processing and modern technology, no deep understanding of the nature of
viscoelasticity has been reached. Since no fundamental relation is believed to
have yet been discovered, at least more than ten popular constitutive equations
(CEs) are in competition at present without any clue to a preferable type. For
several decades, there have been many attempts to derive CEs for polymer
fluids from the viewpoints of mechanics, mathematics and physics.
The first approach, which was adopted by many specialists in mechanics, was
purely rheological. It postulated nonlinear and quasi-linear relations between
the observable variables, the stress tensor cr and the strain rate tensor e. Oldroyd
[ 1,2] pioneered the method and also revealed important principles of invariance.
The concept was further developed by scientists such as Rivlin, Green and
Ericksen and their numerous successors [3]. Later, it was recognized that many
rheological equations derived from different approaches are associated with the
equations proposed by Oldroyd [ 1,2]. A great many rheological equations, both
of differential and integral types have been proposed, and they were able to
describe some properties of viscoelastic liquids. The lack of thermodynamic
analysis is the main disadvantage of this approach. Hence, such important
phenomena as dynamic birefringence, non-isothermal flow, and diffusion
520

cannot be considered simultaneously with rheological constitutive equations.


Also, the viscoelastic constitutive equations are often non-evolutionary.
The second approach was developed mainly by such scientists as Noll,
Coleman, Truesdell, and their colleagues [4-7]. It is the rational mechanics
which searched for the most general form of constitutive relations between
kinematicand dynamic variables. The basic system involving the constitutive
and thermodynamic equations was constructed using strict mathematics, with
the CEs having to satisfy the general principles of causality, material objectivity
and local action. In this way, the properties of all viscoelastic liquids can be
described by a set of hereditary functionals with "fading memory", whose
invariance and thermodynamic consistency were perfectly revealed.
Unfortunately, there is no unique way to specify the memory functionals and
hence predictions are not possible.
The third approach is purely physical and explains the behavior of polymeric
liquids in terms of the intra- and inter-molecular dynamics. At the beginning,
this approach was used to study the behavior of dilute polymer solutions by
Kargin, S lonimsky, Kirkwood, Riseman, Rouse, Zimm, Bueche and others (see
[8] for a review]). For a long time, concentrated polymer solutions and melts
have also been considered as "temporary networks" of entangled chains. Green
and Yobolsky [9], Lodge [10], and Yamamoto [1 1] developed semi-
phenomenological theories that extend the theory for rubber elasticity [12]. This
idea was enhanced by the rise of the "reptation" theories, which are due to the
work of de Gennes [8], Edwards [13], and Doi and Edwards [14]. In the
reptation approach, the creation and decay of the molecular entanglements are
studied by the statistical description of a polymer molecule moving along its
own axis within a "tube" created by surrounding molecules, and then the motion
of the molecule is averaged over high frequency transverse Brownian motion.
Another reptational or statistical approach can be found in the works by Curtiss,
Bird et al. [ 15-18]. Pokrovsky and Vo|kov [ 19,20] proposed a non-reptational
approach based on a generalized Langevin equation for a single macromo|ecule
moving in macromolecular environment. A more fundamental approach to the
linear viscoelastic properties of polymers has been recently developed by
Schweizer [21,22]. This microscopic theory of the dynamics of polymeric
liquids called the "mode-mode coupling" (MMC) approach, omits the specific
assumptions of reptation theories and formulates the statistical properties of
polymers in the most systematic and fundamental way. In contrast to the
rational mechanics, the statistical methods have generated a number of specific
CEs, and have been generally successful for explaining viscoelastic behavior of
polymer liquids in the region of linear or weakly nonlinear deformations. Apart
521

from its poor description of nonlinear data, the physical approaches are also not
free from empiricism in formulations and arbitrariness in overcoming
mathematical difficulties. Additionally, recent mathematical analyses and
numerical simulations revealed examples of unphysical and unstable behavior
of the CEs derived by these methods.
The fourth method of deriving CEs has been introduced by the author [23,24].
It investigated nonlinear viscoelastic phenomena by the methods of the quasi-
linear irreversible thermodynamics with the use of a "recoverable strain tensor"
as hidden parameter. By this approach, a class of Maxwell-type differential CEs
has been proposed under strict stability constraints which are based mainly on
thermodynamics. Later, similar approach was adopted by Dashner and
Vanarsdale [25] to formulate the general class of CEs almost equivalent to the
author's.
Almost all CEs proposed in the literature have a limited ability to describe
start-up, steady state and relaxation phenomena of polymer fluids in standard
rheometric flows, within a relatively narrow region of strain rates usually
employed in the tests [26]. However, there are two frustrating problems: (i)
none of viscoelastic CE proposed could describe the whole set of available data
with one specified set of parameters, (ii) in real modem processing the values of
Deborah number, De, may be at least two orders of magnitude higher than those
in usual rheological tests, and in that flow region almost all CEs exhibit various
instabilities, the reason for which still remains unclear.
There are a lot of speculations in the literature on how the instabilities in CEs
are related to those observed in flows of polymer fluids There are also contrary
opinions about the physical sense of non-evolutionary behavior of viscoelastic
constitutive equations (see, e.g. Refs. [27,28] with contrary views on the
subject). The temptation to relate the instability in CEs to real flow instability
was perhaps caused by the widely spread perception that no rheological
constitutive model is globally stable.
A few books [ 18,26,29-31 ] and a lot of papers on the polymer rheology do not
answer the question as to which CE should be chosen to solve fluid mechanical
problems of polymer processing in the usual case of large De number flows,
where the nonlinear effects of elasticity are important. Therefore, general
studies in this field should be aimed on searching for not only descriptive but
also reliable CEs. Thus in our opinion, the principles of choice of CEs should
also include, besides the descriptive ability of viscoelastic CEs, such
fundamental thermodynamic properties of these as formulation of dissipation
and free energy, along with stability constraints.
522

To study these fundamental properties, even for the relatively simple CEs, one
has to employ a general formalism within the framework of which it is feasible
to find some general stability constraints. Several formal approaches to the
viscoelastic CEs have been proposed relatively recently in the literature. One of
them is the local quasi-linear approach of non-equilibrium thermodynamics
developed by the author [23,24,32]. As mentioned, this approach has resulted in
derivation of a class of Maxwell-like CEs, with some thermodynamic
constraints. Almost 10 years later, the Poisson-bracket variational approach has
also been established. It was first introduced by Grmela [33,34] and then
elaborated by Beris and Edwards [35]. This approach, extending the
Hamiltonian formalism in classical mechanics to the case of continuum
mechanics, also employs the dissipative functional. It was proved [36] that both
the above approaches result in the same "canonical" formulation of general
Maxwell-like CEs. Another two general formal schemes, consistent with the
irreversible thermodynamics, have been developed by Jongschaap et a1.[37,38]
and by Kwon and Shen [39]. The first one employs a matrix representation in
the phenomenological relations between the thermodynamic forces and fluxes,
with additional relations between external and internal (hidden) thermodynamic
variables. The second one uses the notion of evolution of the temporary network
structure. Regardless of all the differences in their detailed schemes, their
equivalence to the general author's approach is evident.
Another objective that should be kept in mind when proposing the CEs for
polymeric liquids, is solving geometrically complicated flow problems under
high deformation rate, as related to industrial processing. Complexity of flow
problems for polymeric liquids is enhanced by the effects of fading memory
which do not exist in viscous liquids or elastic solids. Hence, viscoelastic
polymer fluids demonstrate many unique rheological effects, such as nonlinear
evolution of stress under steady deformation, which cannot be seen in the other
fluids. Because of this, even geometrically simple problems which may be
solved analytically for viscous fluids, have often to be treated numerically in the
viscoelastic case. This necessitates the elaboration of certain criteria for the
selection of CEs for practical applications. The following principles of choice
have been suggested [40,41 ] for the selection of CEs:
(i) Stability. However well an unstable CE can describe rheometric tests, it is
impossible to use it in modeling of polymer processing, since the Deborah
numbers there may be at least two order of magnitude higher and the flow much
more complicated. Extrapolation of the majority of CEs to the region of high
Deborah numbers and 3D flows usually results in several types of instabilities
in numerical flow simulations. These instabilities reflect the mathematical
523

structure of the proposed CEs. In the most cases, they are not related to the
physical instabilities observed in the flows of polymeric fluids, or poor
numerical algorithms, but rather to violations of some fundamental principles.
(ii) Descriptive ability and flexibility. It is now well recognized that polymer
melts with similar linear viscoelastic spectra can show qualitatively different
nonlinear behavior. For the proper description of various flows, this requires
some functions of the kinematic variables, and the associated nonlinear
parameters in the CE (which vanish in the linear limit) to be specified within the
stability constraints. Once these functional forms and parameters are specified
for a particular polymer, the CE must simultaneously describe the entire set of
available experimental data fairly accurately.
(iii) Computational economy. The proposed CE should allow for numerical
calculations in complex flows with as little computational effort as possible. For
example, despite the good descriptive ability, it is rather cumbersome to work
with models in which the elastic potential is specified in terms of the principal
values of a strain measure, and it is usually conceived that working with CEs of
differential type is preferable for numerical calculation than with integral ones.
(iv) Extensibility. Real polymer processing is confronted with a variety of
complications such as compressibility, non-isothermality, wall slip, phase
transitions and separations, chemical effects (degradation, curing), etc. In
principle, the CE of choice should be amenable to extension in order to
accommodate these phenomena.
Among the four principles listed above, the first two can be regarded as the
most fundamental properties which the CEs should possess.

2. FORMULATION OF VISCOELASTIC CONSTITUTIVE


EQUATIONS

There are two types of viscoelastic constitutive equations which are only in
use today for practical applications. These are of differential (Maxwell-like) and
single integral types. Therefore we discuss below only the formulation of these
two types of CEs, since only they have been tested up to present.

2.1. Formulation of general Maxwell like constitutive equations [36]

2.1.1. Thermodynamics and general single-mode viscoelastic approach


We will employ in this section the general approach of irreversible
thermodynamics discussed in various books (see, e.g. [42,43]). The fundamental
hypothesis of"local equilibrium" underlying the method assumes that even in a
524

non-equilibrium process, in any arbitrarily small macroscopic particle of a


medium, it is feasible to operate with the same common thermodynamic
functions, depending on the same variables, as in true thermodynamic
equilibrium. It is also possible to involve in the analysis some non-equilibrium
parameters which vanish in equilibrium; the less these are involved the more
economic is the description. This assumes that the system under study is in a
sense close to a thermodynamic equilibrium. For viscoelastic liquids, the true
thermodynamic equilibrium is at the rest state. The same holds for elastic solids,
but for the liquids there additiona|ly exists an incomplete thermodynamic
equilibrium in the stressed state. Though strong flows of elastic liquids can be
far away from true thermodynamic equilibrium, we can still assume that they
are close to the state of incomplete equilibrium which is characteristic for
stressed elastic solids. In fact, this fundamental hypothesis underlies all the
theoretical treatments of elastic liquids, the molecular approaches, Poisson-
bracket and matrix formalisms included, and reflects our intuitive view of shear
elasticity in liquids.
We assume that in the simple case under study, the state variables of elastic
liquids are" the temperature T and a hidden variable, symmetric second-rank
non-dimensional tensor =c. We assume additionally that the tensor ___eis positive
definite, which to some extent, have been justified independently [36,44]. In the
following, we employ the Eulerian formulation of the constitutive equations
and, without loss of generality, use a Cartesian coordinate system. The physical
sense of tensor c can vary, however. In our own studies [23,24] the tensor =ewas
treated as the Finger elastic (recoverable) strain C, 1, which may be measured
independently. When related to molecular approaches, the tensor c is
proportional to the averaged diadic built up by the end-to-end vector of a part of
a macromolecule between two entanglements. In these approaches, the
configuration tensor is a typical internal variable and the problem of how to
measure it directly is questionable.
Being mostly interested in studies of isothermal flows of isotropic elastic
liquids, we can now introduce as a proper thermodynamic potential, the free
energy density per mass unit, F = F( T, It, 12, 13). Here

/1 - tre, 12 - 1/2(I12-tr__e2), 13 - detc, (1)

are the basic invariants of the tensor g. Then the "thermodynamic stress tensor"
0% associated with the free energy F, as in the quasi-equilibrium elastic case
[45], is introduced as follows"
(o~ = 2pg=.c3F/Oc = 2p[F,c_- + F2(I,c=- s + Ffl,6=] (Fj = OF/OIj) (2)
525

Here 9 is the density and fi is the unit tensor. The only physical reason to
operate with the equation (2) is the explicit assumption [23,24] that polymeric
liquids always possess an "elastic limit", a quasi-equilibrium situation achieved
on very rapid (instantaneous) deformations, where the temporary entanglements
in macromolecules act like cross-links in cross-linked rubbers. Thus, in this
limit, when g --+ _Cl the use of equation (2) can be justified. It should also be
noted that no assumption of incompressibility has been made.
We now briefly discuss the dissipative effects [23,24] which are associated
with flows of elastic liquids. By using the laws of mass and energy conservation
as well as the momentum balance, and separating the total entropy variation in
the flux and origin ("entropy production"), one can easily obtain the Clausius-
Duhem expression for the entropy production Ps in the case under study:

TPs = -~I.VT + tr(=o-.e) - pdF/dt [ T (3)

Here ~/is the thermal flux, __o-isthe actual stress tensor, g is the strain-rate tensor,
and the symbol d/dt means the time derivative in the frame of reference
associated with a moving particle of liquid. The first term on the right-hand side
of equation (3) reflects non-isothermal effects and could be considered as
independent of the other two, which are related to mechanical dissipation.
According to the Second Law, P~ should be positive for all non-equilibrium
processes and vanish at equilibrium. Considering further the isothermal
behavior of elastic liquids. Using equation (2), we can rewrite the dissipative
terms in (3) as follows"

D = TPs[T = tr(cr-e) - tr(_O-e'e-l"1/2d__e/dt), (4)

where D is the mechanical dissipation.


The next step in developing the non-equilibrium thermodynamic approach is
to represent the dissipation as a characteristic bilinear form D = ~ k " Yk where Xk
are thermodynamic "forces" and Yk are conjugated thermodynamic "fluxes". As
thermodynamic forces, we can naturally take the actual, o-, and equilibrium, =O-e,
stresses. Then the thermodynamic flux conjugated to the actual stress =o-is
certainly the strain-rate tensor g. Another thermodynamic flux, ee, conjugated to
the quantity ere is a tensor which is related but is by no means equal to
1 -I
-c .de/dt. This is because the latter does not even satisfy the frame-
2 = =
invariance conditions. This situation demonstrates an ambiguity in definition of
526

the second thermodynamic flux, e~. Mathematically, it associates with


arbitrariness in finding a tensor from a scalar product of two tensors. Thus an
additional physical assumption is needed to define the unknown quantity e~.
This has been proposed in [23] as follows: not only thermodynamic force, cr~ but
also conjugated to this, the thermodynamic flux ee, have to be defined as in the
case o f quasi-equilibrium elastic solids. This assumption seems to match
reasonably well the local equilibrium hypothesis. It also results in the following
o

procedure" to find the solution, e~, from the kinematic relation c=c.__e+e.__c,valid
for elastic solids (or for the total continuum) and hold the expression for the
o

non-equilibrium case. Here c is the co-rotational or Jaumann time derivative of


the tensor c. It is exactly the same procedure that we have used in the definition
of the equilibrium stress tensor ere. In so doing we can define the quantity e, as
the solution of the equation"

f = C.ee + ee'_C_, (5)

which has been obtained in an explicit form [24], too awkward to be reproduced
here. It is easy to see from equation (5) that the term pdF/dt I T in equation (3) is
reduced to tr(o-.e~), which, in turn, can be represented in the form of the second
term in the dissipation equality (4). It should also be noted that this derivation,
being general enough, was confirmed by an independent specific analysis of
kinematics of viscoelastic deformations [24,46,47] when ___e= C~l. If the non-
equilibrium stress _O-pand strain rate ep are defined as follows,

_O-p- g - o-~, __ep- __e-ee

the dissipation inequality due to equation (5) is rewritten in the form:

D - TPs[T - tr(_O-p.e) + tr(o-~.ep). (6)

Equation (6) is now represented as the typical bilinear form D = f~(k'Yk


discussed above, where C~pand ee are independent thermodynamic forces, and ___ep
and ee are independent thermodynamic fluxes. Two independent sources of
mechanical dissipation are now clearly seen from equation (6); both of them
being positive defined and vanishing in the equilibrium. They are" (i) the power
produced by the irreversible stress _O-pon the total strain rate =e (the first term in
(6)), and (ii) the power produced by the reversible stress o-~ on the irreversible
527

strain rate ep (the second term in (6)). Using quasi-linear scheme of irreversible
thermodynamics, these were connected in [23,24,46] by phenomenological
relations with kinetic coefficients represented as some rank four tensors
depending on the internal parameter, tensor g. In presenting these relations, the
Onsager symmetry of kinetic coefficients, proved for quasi-linear case in [43],
has also been used. The structure of the kinetic rank-four tensor had also been
completely revealed [24]. An interested reader can find the details in papers
[24,46,47]. This strict, straightforward and complete approach can be compared
with somewhat controversial matrix approach [43,44].

2.1.2. Maxwell models with quasi-equilibrium stress.


We will consider from now on only the Maxwell liquids for which the actual
stress tensor is equal to the thermodynamic stress, i.e. =o-= fie- (=O-p= 0). In this
case, the first term in equation (6) vanishes. Then the quasi-linear
phenomenological relation between the thermodynamic force, _fie and the
generalized thermodynamic flux, gp, is of the form"

ep,ij = m0kl(T,c)O'e,kl(Y,c). (7)

Here M is a rank-four mobility tensor, an isotropic tensor function depending on


the configuration tensor =e; the stress tensor _o-~now having the form shown in
equation (2). Substituting (7) into the dissipative equality (6) with O-p = 0
represents the latter in the quadratic form of the stress tensor, D = M~st-cr~-=o-~.~t.
This in turn results in the fact that the mobility tensor has to be positive definite
and has the following symmetry properties" it is symmetric in the first two and
second two indices and in transposition of these indices. Also, substituting
expression (7) for ge into equation (4) yields the evolution equation for the
configuration tensor c"

V
c+~c)/0(T) = O, ~c)/O(T) = 2C-ep(T,c). (~j(T,c)/O(T) = 2CikMkist(T,c)O's,(_~._C))
(8)
V
Here c is the upper convected time derivative and ~ c ) is a non-dimensional
isotropic tensor function of tensor c defined as in equation (8) and related to the
dissipative processes in an elastic liquid. Additionally, in the rest state when c--+
=_8(I1--+I2---~3; I3---~1), we have" =d(T,c)---~=0. We assume that this limit transition is
regular, meaning that there is a limit to the linear Maxwell viscoelastic equation
when the intensity of strain rate is very low. In the incompressible case with
528

linear dependencies Cre(=e)and =d(=c), equations (2) and (8) are easily reduced to
the upper convected Maxwell model.
Multiplying the first eqation (8) from the right and from the left by g-~ reduces
the latter to the a "dual form'"

+ s = 0, (b = __e-') (8a)

where b is the lower convected time derivative of the tensor b. Also, in the
incompressible ease with linear dependencies of ~ and ~ ) , eqations (2)
and (8) are reduced to the usual form of the lower convected Maxwell model.
The difference between these is mainly in the dependence of stress on elastic
strain" =o-= G=e corresponds to the case of rubber elasticity, whereas =o-= -Gb can
describe the elasticity of crystals or metals.
Equations (1),(2) and (8) give rise to the energy relation:

r + D - tr(Cre.e), D - tr(g~- Cre.gb)/(20(T)), (9)

where D is the mechanical dissipation.


If the new kinetic rank-four tensor L(e=) is introduced as follows"

Lijkl = CimCjnMnmkl (10)

the dissipative term ~___e)in equation (8) is rewritten in the form"

~j( e=)/O(T) = LoglpOF/c3ckl (11)

Substituting equation (11) into (8) represents the latter in the form which
coincides with the evolution equation obtained by Beris and Edwards (see
equation (2.13) in the second part of Ref.[35]). The tensor L has the structure
[35] similar to that known for the tensor M.
We now analyze the compressibility condition in the general Maxwell model
described by equations (2) and (8). Multiplying equation (8) scalarly by (2=e)1
gives:

1/2d/dt(lnl3) + trep - tr__e; trep - 1/2tr[__c-l.=d(_c_)/0(T)], (12)


where the function =d(c) was defined in equation (8). There is also the mass
conservation equation which can be written in the form"
529

d/dt(lnp/po) = -tre. (13)

Here/90 is the density in the rest state. Then, combining equations (12) and (13)
yields:

d/dt[ln(,O/po. ~-3 ) + trep = O. (14)

Equation (14) shows that P---~Po when c--+_6, since in this limit, trep--+0. It also
shows [23,24] that

if trep = 0, #/9o = 1/-4~-3, (15)

exactly as in the equilibrium limit of elastic solids.


Thus, the evolution equation (8) for tensor ___calso describes the law of mass
conservation in compressible elastic liquid only if the condition tr__ep = 0 is
satisfied. When trep ~ 0, the density variations are not described anymore by the
configuration tensor c but satisfy the kinetic equation (13) or the equation of
mass conservation (12). It means that, in this case, the density is not a state
(thermodynamic) variable and an attempt to improve the situation by making
the assumption that F = F(T,p,e=) contradicts the Murnaghan's definition (2) of
the thermodynamic stress. This was the only reason why in our publications
[23,24,46] we also employed the condition trep = 0.
A simple constitutive equation for the compressible part of stress has been
discussed in [23,24]. It includes the equilibrium pressure and also Kelvin-
Voight modeling of compressible effects. This approach has been recently
extended for non-isothermal linear compressibility phenomena with a bulk
relaxation spectrum when introducing a set of scalar independent hidden
parameters {~k} [48].
Consider now the incompressible case when p =/9o - const. When tr__ep :/: 0,
equations (8) still hold, whereas equations (13) and (14) are reduced to:

1/2d/dt(ln/3) + tr__ep- O; tr__e= O. (16)

Surprisingly enough, in this case, the free energy remains the same, i.e.
F=F(T,II,I2,I3), because the tensor =e is not directly related to the density. By
introducing the modified free energy / ? - F - p lnI 3, the actual stress =0-can be
represented by
530

__o-=-p=~+ ere. (17)

Here the isotropic term 'p'(a Lagrange multiplier) in equation (17) is introduced
to satisfy the condition of incompressibility tre = 0; the thermodynamic tensor
ere being still defined by equation (2). This means that the tensor cr~ is not
defined with the accuracy of an isotropic term as in the equilibrium case, and
isotropic pressure is not an equilibrium one. These features result in the fact
that, even in the incompressible case, the expression (6) for the dissipation is
now not invariant under the transformation, _o-e--~o-~+p__6.
When the constraint trep = 0 is used, equation (15) results in the
incompressibility condition [23,24],/3 = detc = 1. In this case, the modified
^

elastic potential is introduced as" W = W(T, II,I2)-f~(I 3- 1), where W = pie. The
magnitude ~ here is the isotropic pressure. It is used in differentiation in
equation (2) as the Lagrange multiplier, allowing us to consider all cij variables
as independent. Then equation (2) results in the Finger formula for the stress:

- -p6_+ 2 Wle- 2 W2c -l. (18)

Here Wk= OWlOIk and p = ~-/;'2/2. Also in this case,/2 = trc l and the formulae
for free energy and the stress tensor are exactly the same as for elastic solids.
Also, isotropic pressure and the expression for dissipation (6) are invariant
under the transformation: _O-e--)Cr~+ p=6.
We finally stress that as shown in this section, any attempt to extend the
configuration tensor approach to the compressible case is thermodynamically
inconsistent. Since the Maxwell type models involving configuration tensor c
were derived from the evolution equation for distribution function, it simply
means that the later equation should be somehow modified. It is unknown at
present what should be done to rectify the situation.

2.1.3. Maxwell models with non-equilibrium and non-potential stress


tensor[36].
The above thermodynamic derivation of Maxwell-like constitutive equations
was further extended [36] to include formally into consideration the
incompressible Gordon/Schowalter [49], Johnson/Segalman [50] and Phan-
Thien/Tanner [51,52] models. For this reason, instead of the natural, quasi-
equilibrium relation (5), the following non-equilibrium evolution equation for
the tensor c was proposed"
531

c = ~(__C'ee+ ee'__c) (5a)

Here ~ is a numerical parameter (-1 <~< 1, %~0). The case %= 0, related to the co-
rotational derivative, can be also easily considered but we shall omit this. Then
following the same procedure, described in the derivation of equation (8), we
can finally obtain [36]:

__o
-= -p__6+ __ere Crex= Cre(g)/~, _c- ~(e.e + =e.___c)+ &(T,g)/O(T) = 0, (19)
D = tr(Crex-&).

In equations (19), written here in so-called canonical form, the formula for the
extra stress tensor Crex is given due to some thermodynamic reasons [26]. Note
that when ~,~+1, equations (19) have no perfect elastic limit and therefore they
are non-equilibrium.
One can also ignore the thermodynamic origin in the stress formulation ere(C)
and consider formally the second formula in equation (19) as a general non-
potential stress-elastic strain relation O-~x= Crex(C,~). It can be formally done even
in the case ~,=+1, where there is a limit to elastic solid behavior, ~(T,_e_)--+=0.
However, in this case, it is always possible to consider such a loading-unloading
procedure that will create work from nothing [53]. It means that this approach
inevitably leads to the theoretical eternal motion machine and therefore is
thermodynamically forbidden.

2.1.4. Examples of single mode Maxwell-type constitutive equations.


We now illustrate how the particular Maxwell-type CEs can be obtained from
the above general equations by specifying the terms ~ and ere in equations (19)
and (2), or F in equation (1).
When

~b~= g-_6, (2p/G)F = I~-3, ere = G__e, (20)

where p and G are the shear modulus and density, equations (19) represent an
interpolated Maxwell or the Gordon/Schowalter [49] - Johnson/Segalman [50]
model. When ~=+1 or ~=0, it is respectively called the upper/lower convected
or co-rotational Maxwell model.
When
532

0 = 0o(T)f(I1), g~ = g-=b, (2p/G)F = I~-3, _O-e=G___c, (21 )

it becomes the general Phan-Thien/Tanner model [51,52], where f(I1) was


proposed as a linear or exponential, increasing function of I1,. Again, when ~=1,
we particularly call it an upper convected Phan-Thien/Tanner model.
When

0 = Oo(T)f(IIe), IIe=tr(__e2), ~= 1, ~=__c-__6",(2,o/G)F=Ii-3, O-e=Gg, (22)

where f is a decreasing function of lie, it reduces to the White-Metzner CE [54],


which does not belong to the general class of quasi-linear CEs.
The FENE model (see, e.g., Section 8.5.3 in Ref. [26]) can be written as:

Y_,=1, 0==K(I~)c-=d, (2p/G)F=(Ic3) 1nK(I,), o-~=GK(I~)c,


K(I, ) = (Ic- 3 )/([c-ll ), (23)

where Ic=Rcz and Re is the ultimate finite dumbbell length.


When

~=1, =dr=otc2+(1-2ot)c-(1-c~)__d(0<c~<l), (2p/G)F=I~-3, __o-~=Gs (24)

it represents the Giesekus model [55], with a numerical parameter or.


The case

~=1, =d~=B(I~)(c-__6), (2p/G)F=(3/~)lnB(I1), _O-e=G__c/B(I1),


B(I1)--1+~(I,-3)/3 (25)

corresponds to the canonical representation [36] of the Larson differential


model [56].
The general class of the author's incompressible CEs [23,24,47] is represented
as:

~=1, r l /2c[b~(c-I~ 5/3 )-b2(c~-12~_/3)], O-e=2Wlc-2W2c-~ (26)

Here W(I~,Iz) (=pF) is the strain energy function for incompressible case. The
positive functions bi(Ii,I2) should have a proper linear viscoelastic limit and
their positive definiteness suffices the positive definiteness of the dissipation.
The convexity constraints,
533

Fl>0, F2>0, F11Fzz>F12


z, (Fi=c3FIc3Ii,Fij=c3Filc3Ij) (27)

imposed on the general form of potential F were also suggested [23,24,47]. The
important implication of inequalities (27) and the proper use of this class of CEs
are discussed in detail in [40]. In the simple case of bl=b2=l, with the neo-
Hookean potential for F, it reduces to the simplest "Leonov model" which
includes no nonlinear parameters.

2.1.5. Multi-mode Maxwell models.


All the above single mode viscoelastic constitutive equations of Maxwell-type
are usually extended to the multi-mode case. We briefly discuss below only
models with potential stresses in the common incompressible case. The typical
N-mode extension is as follows"

N N

F - y'Fk(T, ck),
k=l
cy - ~ r
---ex -= --- ex,k
(T,c )
=k ~
(28)

Here Fk, Crex,kand Ck are the free energy, extra stress tensor and configuration
tensor, respectively, in "k"th relaxation mode. Additionally, the general form of
evolution equation presented in equation (18) holds for every configuration
tensor s In the limit of very low Deborah number, this approach shows the
linear viscoelastic behavior, with a discrete spectrum of relaxation times { Ok}. It
means that every nonlinear relaxation mode is generated by the corresponding
linear viscoelastic mode. The generalization to the multi-mode case can be
justified only if the various relaxation modes are well separated, i.e.

01>>02>>...>>ON (29)

In this case, it is reasonable to assume that the various relaxation modes act
independently. All the known experimental datatestify in favor of inequalities
(29). Additionally, the guess-independent Pade-Laplace method (see e.g. [57])
reveals the effective discrete linear relaxation spectrum in accord with (29).

2.2. Formulation of nonlinear single integral constitutive equations


From a wide class of viscoelastic CEs of the integral type, only the single
integral ones have been experimentally tested. In the common incompressible
case, its general form is represented as [58]"
534

o%= ~ [qg,(I,,I2,t- x)C- qg2(II,I2,t- "r.)C-l]dx (30)


--00

Here _O-ex is the extra stress tensor, C is the Finger total strain tensor for
incompressible media, whose time evolution is described as follows"

v
C-dC/dt-C.Vv-(Vv_)
_ _ _._ _~_ ~
r -C-O;
___ =
C[
I = T
=8
- -
(31)

Here I l= trC, I2 = trC -1, and ~, and q32 are generally independent functions. We
can also introduce many other measures of deformations. One of them, the
Hencky measure, H = (l/2)lnC, will be used below.
Experiments show that the simplified time-strain separable version of equation
(31),

6k (Ii,/2,t-x) = m(t-x)q)k(I1,/2) (k=l,2) (32)

can be introduced. Here re(t-x)- dG(t-x)/dx, and G(t) is the relaxation modulus.
Equations (30) and (31) are not the only single integral form of CEs. When, for
example, the mixed convected time derivative is used, the CE can also be
represented in an integral form (see, e.g., Ref [26]). Also, Kaye [57] and
Bernstein et al. [60] proposed the potential form of equations (30) and (31):

(O, =(2p/G)OP/DI~, ~p2=(2,o/G)0~'/012; (33)

or in the time-strain separable case"

q)~=(2p/a)3F/Ol~, tp2=(2p/a)OF/OI2. (34)

Here G is the elastic Hookean modulus. Equations (33) or (34) constitute the K-
^

BKZ class of single integral CEs. The potential F in equations (33) denotes the
thermodynamic free energy with relaxation effects taken into account.
For the time-strain separable viscoelastic CEs with potential F, the basic
functionals such as the free energy W, the extra stress tensor o-e and the
dissipation D are of the form [61 ]:
535

W = (p/G) iF ( I, ,I 2 )m(t - z)dx (35)


-oo

t
o-~ = (2p/G) IC. c3F/~C(I,,I 2 ) m ( t - , ) d , (36)
--O0

D - tr(crce)-dW/dt = Co~G) IF (I~,I 2 ) [ d m ( t - x) / dxld~ (37)


--oo

A comparison of equations (2) and (8) with equations (35) and (36) clearly
shows that CEs of the differential type where the dissipation and free energy are
generally independent, are more flexible for rheological modeling than CEs of
the integral type where those quantities are roughly proportional.
We now describe some particular CEs of separable single integral type by
specifying q~, and q~2 in equation (30) or the potential F for the K-BKZ type in
equations (34).
Wagner et al. [62] proposed their first specification as:

q~l=fexp(-nl 4I - 3 )+(1-J)exp(-n2 ~/I - 3 ), I=flll+(1- /2; r


(38)

where f, nl, /</2 and, fl are positive fitting parameters. Later, Wagner and
Demarmels [63] introduced a new factorable version:

q 2: h(I~,/2) = l/(ot-,/~), z = (/1-3)(I2-3). (39)

Here again, a and, , are positive fitting parameters. We further refer to the
specification (38) as the Wagner model I, and (39) as the Wagner model II.
Luo and Tanner [64] proposed the modification of the CE originally presented
by Papanastasiou et al. [65] in the form:

qo,= oth(Ii,I2), qo2= ~:oth(IiJ2); h(IiJ2) = [ot-3+flll+(1-tS')I2]1, (40)

where c~, /3 and ~: are fitting parameters and K>0. When K=0, this model
becomes identical to the one proposed by Papanastasiou et al. [65]. One can
easily see that no thermodynamic potential relation exists (q012 ~ q~21) in all of
the above three CEs (38)-(40).
536

The Lodge model [10], which is reduced to the integral presentation of


UCM[1],

q01=l, q)2=0, (41)

is a particular case of the separable K-BKZ class (34) when the neo-Hookean
potential is used.
In order to describe better the viscometric data, Larson and Monroe [66]
suggested the following 4-parametric form ofF for the K-BKZ class:

(2p/G)F = (3/2ct)1 n[ 1+~(I-3)/3], I = (1-fl)ll + ~ / 1 + 21312 - 1, (42)


Ot=Ko+K2tanq[K1A3/(1+A2)], A=I2-I1

Here Ko, K1, ~:2 and fl are numerical fitting constants.


Another potential form was derived by Currie [67] as a close approximation of
the Doi-Edwards reptation model [ 14]. It is written as

(2p/G)F = (5/2)1 n[(J- 1)/7], J = ll + 2x/I2+ 1 3 / 4 , (43)

and contains no nonlinear fitting parameters.


A linear combination of simple potential forms,

2p/~=. G! t- G2 e x p ( - t
exp( -~- ). In[1 + ~( I , - 3)] + 21302 -~2 )" In[1 + 13(12 - 3)] +
2or0
G3 exp( - t (44)
20--7 O3 )" ( I , - 3)

was also introduced by Yen and McIntire [68] as a partially time-strain


separable version of the general potential presented by Zapas [69].Here the
relation between the potential /~ and CE is shown in equations (33), Gi's and
0,'s are moduli and relaxation times, respectively, and a and fl are nonlinear
fitting parameters.
It should also be noted that the thermodynamic potentials proposed in the
theory of rubber elasticity can in principle be applied in the case of K-BKZ
class of integral CEs as well as for CEs of the differential type. In many cases,
they are presented as a function of principal values Ci of the Finger tensor rather
537

than invariants of the total Finger deformation tensor. Below are several
examples of these potentials.

(i) Ogden potential [70]:

2PF= ~ G,, (C.,,~


1 2 +C~",~2 + C ~,,~2
3 _ 3) (45)

Here Gtn are numerical parameters which can be negative or positive, and Gn are
shear moduli. The potential (45) becomes identical to the Mooney potential [71]
if it contains only two modes corresponding to a 1=2 and cx2= -2.

(ii) Valanis/Landel potential [72]:


3

2pF= i~[aff-Q, (lnff-C , - 1 ) + 131n.fQ-~] (46)

where c~ and 13 are parameters with dimensionality of modulus.

(iii) The BSTpotential [73 ]

2pF-- A I ~ + B
_ _ o m
IE, I E - 1 (C~,,/2
~
, + C ,~2
2 + C ,3/ 2 ) .
n n (47)
n n

Here A and B are the parameters with the dimensionality of modulus, whereas n
and m are numerical parameters.
As mentioned, the non-potential viscoelastic CEs proposed in the literature,
are unphysical, since when applying very fast deformations, it is possible to
create a perpetual motion machine from a hypothetical material subordinated to
this type of CE [53]. Nevertheless, we will analyze below this type of CEs too,
considering it as mathematical abstraction for the superficial data curve fitting.
To make the following analyses more efficient, we now introduce a unified set
of notations for both differential and single integral types of CEs which employ
only upper convected time derivatives in the evolution equations. The lower
convected time derivative (~=-1) in equation (19), can be equivalently rewritten
in the form of upper one (~=1) [36], as shown by equation (8a). We introduce a
modified pressure term defined as:
538

for differential CEs


P t (48)
+ jq)212m(t- 1:)d1: for integral CEs
--cO

Then using the Cayley-Hamilton identity and the invariance of rheological


variables (say, the extra stress o-~) under arbitrary addition of isotropic terms in
the case of incompressibility, we represents both the classes of CEs as follows:

____o-=-p'=d+ O-e, O-~= 2,~__.SF/SR,


for differential CEs
o (49)
=e [_~m(t_,)E(c)dx for integral CEs

E = 2p[q01c + q02(Ilc - c 2) + q0113~

Here 8/8g is in general the partial Frechet derivative with respect to c, and with
the definition for q)i in equations (2) and (30). In this notation, c becomes the
total Finger strain tensor __Cin the case of integral CEs, and thus for the CEs of
integral type and the author's CEs, I3=1, q)3=0 automatically. Even though the
set (40) is written for hyper-viscoelastic equations, the non-potential
viscoelastic formulations can also be included into consideration.
The (elastic or total) strain tensor c is the solution of the following evolution
problem:
v t el,__o = =8 for differential CEs
(50)
= = = = [
Ct=t,
= C] t = t t = =8, ~(c) - =0
__ =
for integral CEs

Here ~bis the dissipative term which vanishes for integral CEs, and from now on
we consider only upper convected time derivatives even for differential models.

3. STABILITY OF VISCOELASTIC CONSTITUTIVE


EQUATIONS

3.1. Background
In numerical simulations of viscoelastic flows, degradation of the numerical
solution or lack of convergence of computational schemes has been frequently
539

observed for large or even modest values of Deborah numbers. It is thought that
the main reason for this numerical malady is an improper choice of a CE which
possess some bad mathematical properties (e.g., see Ref. [74], p.314). Therefore
the analysis of such properties of various viscoelastic CEs as their stability and
boundedness seems to be a pre-requisite for any successful numerical modeling.
It will be shown in this Section that these properties are in turn, originated in
our capability to incorporate basic thermodynamics laws in formulation of the
CEs.
We discuss in this section recent results of stability analyses for isothermal
formulations of CEs for viscoelastic liquids. To begin with, we briefly describe
the basic aspects of stability analysis in general, and two types of instabilities
related to the formulation of viscoelastic CEs in particular.
The general purpose of the stability analyses is to find the conditions of
boundedness for a solution of a set of functional equations subordinate to some
additional (usually initial and boundary) conditions. In our case, the total set of
equations consists of momentum balance and continuity equations coupled with
viscoelastic CEs. No-slip boundary conditions are usually applied at the rigid
surfaces and the common set of kinematic and dynamic boundary conditions is
used at the unknown free surfaces. It is assumed a "basic" solution of the set
(generally, 3D and time dependent) to exist, subordinate to initial and boundary
conditions. The solution is said to be stable in the Lyapunov sense, if "small"
disturbances imposed on the solution at time to, which satisfy the homogeneous
boundary conditions, remain to be small at any time t (to<t<oo). The solution is
said to be stable relative to the arbitrary "large" initial disturbances if they
remain to be bounded at any time t (to<t<oo). The particular type of the latter
stability is the boundedness of the basic solution itself. It is in principle also
possible to analyze the stability of the solution relative to small or large
disturbances in boundary conditions and in parameters ("robustness"). These
types of stability analyses are not considered below.
The stability "under small disturbances" is usually considered by analyzing a
linearized total set of equations, if such a linearization is possible. In spite of
their linear character, the initial boundary-value problems for the evolution of
disturbances are very difficult to solve, even in the simplest cases. A great deal
of simplification can be achieved when spacetime scales of disturbances are
considerably less than those scales for the basic solution. In this case, one can
analyze locally, i.e. in the vicinity of any spacetime point, the linear stability
problem, using elementary waves of disturbances and ignoring the boundary
conditions. This is the method of "frozen coefficients", which is illustrated in
Section 3.2.
540

The purpose of the stability analyses discussed in the following sections is not
to describe the real physical instabilities observed in flows of viscoelastic
liquids, but rather to reveal some constraints which should be imposed on the
formulation of CEs to prevent them from the occurrence of unphysical
instabilities. Therefore, we call the stability considered in the subsequent
analyses not the stability of viscoelastic flows but the stability of CEs. These are
related to the formulation of CEs.
The formulation of nonlinear viscoelastic CEs is not as straightforward as it
appears at first sight. For many specifications of CEs in the above two classes,
numerical modeling of high Deborah number flows has displayed heavy
unphysical instabilities, regardless of the "true physics" employed in their
derivation. Two types of instabilities related to the formulation of CEs were
observed for large or even modest values of Deborah numbers, and have been
analyzed in the literature. These are the Hadamard and dissipative instabilities.
The case of Hadamard stability is the most understood. We define the
complete set of equations for viscoelastic liquids as Hadamard stable (or
evolutionary, or well-posed) when the solution of the boundary-value Cauchy
problems for the set at any time provides the complete initial conditions for
determining the solution at subsequent instants in time [75]. Thus, the
Hadamard stability allows one to continue the solutions in the positive direction
of the time axis. When this is impossible, very quick blow-up instability occurs,
with extremely short wave disturbances, which results in progressive failure in
numerical calculations: the finer the mesh, the worse will be the degradation of
the results [76]. In many cases, one can treat the Hadamard instability as a
blow-up type increase in the amplitude of initially infinitesimal waves of
disturbances as the wavelength tends to zero. In viscoelastic liquids, this type of
instability can be associated with a nonlinear rapid response of CEs. Hence, this
type of instability depends on such quasi-equilibrium properties of the CEs as
the type of differential operator in the evolution equation for differential models
and the elastic potential in the hyper-viscoelastic case.
Rutkevich [77] initiated the study of Hadamard stability for viscoelastic CEs.
Later, Godunov [78] analyzed some aspects of the stability in more detail. Some
significant results were obtained relatively recently by Dupret and Marchal [75]
and Joseph and co-workers (see, e.g., Joseph's monograph [76]). These studies
which have analyzed the Hadamard instabilities in particular flows for
particular specifications of differential viscoelastic CEs, can be summarized as
follows. Joseph and co-workers [79,80] and independently Dupret and Marchal
[75] proved that the interpolated Maxwell model, which involves the mixed
time derivative in the evolution equation, is Hadamard unstable, except for the
541

cases of upper and lower convected time derivatives. The ill-posedness of the
Gordon/Schowalter [49] or Johnson/Segalman [50], and the original Phan-
Thien/Tanner [51,52] CEs is subject to this type of instability. Using the general
method of characteristics, Dupret and Marchal [75] also showed that the White-
Metzner model [54] is nonevolutionary, which was again justified by Verdier
and Joseph [79], who employed a perturbation method in their analysis. It was
found that the dependence of the relaxation time on the invariant of strain rate
tensor is the cause of Hadamard instability in the White-Metzner model.
Analyzing the White-Metzner model in elongational flow, Verdier and Joseph
[81] also noticed a type of dissipative instability that occurs whenever the
extensional strain rate exceeds the half of the reciprocal relaxation time.
The method of characteristics is the most general mathematical tool for the
Hadamard stability analysis. This method is, however, sometimes complicated
and cumbersome. Fortunately, in the most interesting cases, it can be simplified
by using the "frozen coefficient" method (see, e.g. [76]), when analyzing only
extremely short and high frequency wave disturbances propagating with a finite
speed. Those cases are related to all the CEs of the quasi-linear differential type
as well as the time-strain separable single integral CEs. Then, the linear stability
analysis of the problem can be studied locally, without considering boundary
conditions. Although following Kreiss' examples [82], this local stability
condition is neither necessary nor sufficient for the overall stability in the
general case of nonlinear partial differential equations, for quasi-linear
differential (and time-strain separable single integral) equations, the local
stability analysis with the method of frozen coefficients can be employed
without loss of generality [83].
The studies of Hadamard stability have also a long history in the theory of
nonlinear elasticity, where many general results were obtained and understood
in great detail. Among the many general stability conditions suggested, the
simplest is known as the Baker-Ericksen inequality [84]. Its physical sense is
that the maximal principal stress always occurs in the direction of the maximal
principal strain. In the case of hyperelastic solids, the thermodynamic stability
criteria called "GCN + conditions" were also established long ago (see, e.g.,
Section 52 in Ref. [4]). Their physical sense is the convexity of the elastic
potential with respect to the Hencky strain measure. This has been also known
as a condition for strong monotonicity of stress with respect to strain. In
nonlinear elasticity, the Hadamard stability criteria of field equations
correspond to the conditions of strong ellipticity which coincide with the
stability requirements known for dynamic problems. The weaker conditions of
ordinary ellipticity deliver the stability constraints for static states [85].
542

The GCN + condition is closely associated with the condition of strong


ellipticity. The lack of symmetry in the representation of second-rank variables
causes a more restrictive condition for the strong ellipticity than for the GCN +
condition [4]. Therefore, such inequalities as the Baker-Ericksen and the GCN +
may be treated as necessary conditions for the strong ellipticity or the Hadamard
stability. For the important case of isotropic incompressible hyperelastic solids,
Zee and Sternberg [85] have recently found the necessary and sufficient
conditions for the strong ellipticity (Hadamard stability) of CEs in a form of
algebraic inequalities imposed on the first and second derivatives of the elastic
potential. For the compressible case, these results have been obtained earlier
[86]. For viscoelastic CEs, general results on global Hadamard stability, i.e.
stability for any type of flow and for any Deborah number, were recently
obtained for both general classes of quasilinear Maxwell-like CEs [36] and
time-strain separable single integral CEs [87]. It was found later that these
algebraic criteria for Hadamard stability were confused with the necessary and
sufficient conditions for thermodynamic stability, that is, GCN +. These
convexity conditions for thermodynamic potential in the hyperelastic case
impose weaker constraints on CEs than the criteria for Hadamard stability. The
complete results on the Hadamard instabilities for both the classes of
viscoelastic CEs under study were then found and published in paper [88] for
incompressible viscoelastic CEs. For compressible case, the complete criteria
for Hadamard stability of viscoelastic CEs were recently found by Kwon [89].
In contrast to the Hadamard instability, another important type of instability, of
dissipative type, results from the formulations of the non-equilibrium
(dissipative) terms in CEs. This type of instability can occur even if the
dissipation is positive definite. The studies of dissipative instabilities, initiated
in paper [36] for general Maxwell-like CEs, were motivated by the fact that the
upper convected Maxwell model which is globally Hadamard stable, displays
the unbounded growth of stress in simple extension when the elongation rate
exceeds the half of the reciprocal relaxation time. For any regular flow with a
given history, a sufficient condition for the dissipative stability, close to the
necessary one, was proposed in [36] for CEs of the differential type. Then the
necessary and sufficient condition for single integral CEs was found in paper
[90]. It was also noticed in those papers that in many viscoelastic flows, neither
strain nor stress histories are known, but rather a complex mixture of both.
Several patterns of pathological behavior as predicted by some popular
specifications of Maxwell-like CEs, related to dissipative instability, were
exposed in 1D flows [91 ]. It was found that it is necessary for the dissipative
543

stability that both the steady flow curves in simple shear and in simple
elongation have to be monotonously and unboundedly increasing.
The combined Hadamard and dissipative stabilities have also been analyzed in
paper [88] for incompressible viscoelastic CEs. The results of this analysis will
be discussed at the end of this Section.

3.2 Hadamard stability criteria for viscoelastic constitutive equations


We now briefly outline the mathematical procedure [88] to obtain the
conditions for global Hadamard stability for quasi-linear differential and time-
strain separable single integral CEs in incompressible isothermal flow of
viscoelastic liquid. It has been proved several times [36,75,79,80] that the
evolution equation (19) with the mixed convected time derivatives (~=+1) is
Hadamard unstable. Therefore, we consider below only the case of upper
convected time derivative in the evolution equation (19) of the differential type
and single integral CEs of the form (36). Then the total set of equations in this
stability problem consists of viscoelastic CEs (48)-(50) complemented by the
equations of momentum balance and incompressibility:

pdv_/dt =-Vp +V-o, V.v=O. (51)

We assume that the set has a solution {__c,v,p} which satisfies some proper initial
and boundary conditions, and impose on the solution extremely short and high
frequency, infinitesimal waves of disturbances"

{8g, 8g, 8v, 8p} = ~{ ~, g, v, p }.exp[i(k.x-mt)/g 21 (52)

Here c is a small amplitude parameter (leJ << 1), c, v, p and p are (generally
complex) amplitudes of the corresponding disturbances, k is a wave vector, and
co is the frequency.
Using just the local linear stability analysis, we can easily find the following
"dispersion relation", i.e. the dependence of the frequency co on the wave vector
k and the parameters of the basic flow:

1 ~,)2~2 _ I B~j'n~V~kjVmk~ for differential models

2 [_~m(t - t 1)Bij,~mdt 1 9v i k j V m k n for integral models

kjvj -0. (53)


544

Here ~ = co- k.v is the frequency of oscillations, with Doppler's shift on the
basic velocity field v taken into account. The last equation in (53) means that
due to the incompressibility condition in (51), the wave vector of disturbances k
is orthogonal to the amplitude of velocity dis.turbance v. The fourth rank tensor
Bijmn is defined in the principal axes of tensor __cwith principal values ck as:

Bijnm- (~,n~n + ~ln~,n )Gij + ~j~nnLim,


Gij = (Ci q- Cj)[(q)l q- q)2(I1 - C i - Cj)], (no sum!) (54)
Lij = 4q)nfifj + 2q~12CiCj(211 - Ci- Cj) - 2q)21(CiCj-1+ ci'lfj)

Formulae (54) show that the tensor Bjjmn depends linearly on the constitutive
functions q0k and their second derivatives q0kj with respect to the basic invariants
Ik, independently of whether the approach is potential or not.
It is now seen that due to equation (52) and the definition of ~ , the
requirement for the stability is that the left-hand side of equation (53) should be
positive. Therefore, the necessary and sufficient condition for global Hadamard
stability is reduced to the following:

1 ~-~2V2
- BijnmvikjVmk n > 0 (55)
2

For the integral type constitutive equations, the condition (55) provides the
stability during relaxation, which is included in the global stability requirement.
On the other hand, the convexity of potential F, or the GCN + condition, can be
represented as follows"

-- 02 F = 4Cmq eip
B ~jmn[30]3,,m > 0, B ijmn -- Oh ~j0h,m 0c qn 0C pj

Here hij is the Hencky strain measure, h__ = (1/2) I n c . To guarantee the
thermodynamic stability, the inequality in (56) should be satisfied for any
arbitrary symmetric tensor ,fl,j with the condition of incompressibility, trl~ = 0.
In the potential (hyper-viscoelastic) case, the identity, Bijnm = B ijmn, holds. The
comparison between the inequalities (55) and (56) shows that due to the
symmetry of the tensor fl,j, the condition (56) is included in the inequality (55),
i.e. the condition (56) imposes weaker stability constraints than the inequality
545

(55). It means that the conditions for the Hadamard stability are stronger than
those of GCN +.
Employing the algebraic procedure which has been used in hyper-elasticity
[85], one can finally obtain the necessary and sufficient conditions for the
global Hadamard stability as the set of algebraic constraints imposed on the
functions q~k:

(i) lai > O, ~i = ((Pl + (p2Ci)~k/CjCk (i~j~k), q>O,


21
(ii) r q- 21-ti> O, q i =(Ii-Ci)(qOl+q02Ci)+2(Ii2-212-Ci 2- s. )[q)llW(q)12+q)21)ci-l-q)22Ci 2] (57)
Ci
(iii) [ ,+21ai + j + 2 ~ j ] 2 > g k - 2 ~ t k (iCj~k).

The additional constraint in (i), the positive definiteness of the tensor c, holds
by definition for the integral CEs. It was also proved for the Maxwell-like CEs
of differential type [36,102], however, only for the flow situations with a given
history.
The above approach to the global Hadamard stability has been recently
extended by Kwon [93] on the compressible case. The new quality which
occurs there is the possibility of longitudinal wave propagation. In the
incompressible case, the speed of the longitudinal wave approaches infinity,
whereas the speed of the transverse wave is finite. Hence, perturbation of basic
solutions by the longitudinal wave was not considered in this stability analysis.
The result was that the wave vector is always orthogonal to the vector of the
main velocity field. However, in the compressible case, the speeds of both
waves have finite values. Thus for stability, the initially infinitesimal amplitude
of disturbing waves of either type (or a mixed type) should remain small all the
time. It means that the conditions of Hadamard stability in compressible case
are more rigid than those for incompressible one. It was demonstrated [89] on
the simple example of Mooney-Rivlin potential with additional term dependent
on density.
Two sufficient conditions for the incompressible case have also been
proposed:

(1) The author's condition (27): the thermodynamic potential F for the author's
class of viscoelastic CEs is a monotonously increasing convex function of
invariants I1, and I2.
546

(2) Renardy's condition" the thermodynamic potential F for the K-BKZ class of
CEs is a monotonously increasing convex function of ~ and

Although Renardy's condition has been proved only for the K-BKZ class, it
also holds for the Maxwell-like CEs with upper convected time derivatives [88].
Since the author's condition (1) is stronger than Renardy's (2), it also guarantees
the global Hadamard stability for the K-BKZ class. The above sufficient
conditions are much more easier to employ than the necessary and sufficient
conditions for Hadamard stability (57). Therefore they are very useful for a brief
evaluation of the stability for new formulations of CEs. For the compressible
CEs, one sufficient condition for the global stability has also been suggested
[89], but it is too complicated to use.

3.3 Dissipative stability criteria for viscoelastic constitutive equations


As mentioned, there can be another source of instability originated from
specification of dissipative terms in viscoelastic CEs. For viscoelastic CEs of
differential type, this instability may happen due to an improper formulation of
the dissipative term ~ (or ~b when ~ = 1) in equations (19), even for the
Hadamard stable CEs with positively definite dissipation. For single integral
CEs, the instability results from fading memory effects in equations (30) and
(36). Although the global criteria for dissipative stability of viscoelastic CEs are
far from being complete (if it is in general possible), we discuss in this Section
two specific criteria that have been proven. In the case of compressible flow, no
theorem on dissipative stability is known yet, but the following theorems are
presumably valid also for the compressible CEs.

3.3.1 Criterion I of dissipative stability


Theorem 1.1 (the case of CEs of the differential type [36]).
Consider the set of upper convected Maxwell-like CEs (8) with the positive
dissipation D = D(T, Ii, 12, 13) defined in equation (9). Let the free energy F be a
non-decreasing smooth function of three invariants Ik. If for any positive
number E, the asymptotic inequality

O > E'll ell when Ilcll oo (11 11- (trc2)m) (58)

holds, then in any regular flow, the configuration tensor =e and the stress tensor
ere are limited.
547

Theorem 1.2 (the case of single integral CEs [90]).


In any regular flow, the functionals of free energy (35) and dissipation (37) are
bounded, if (and only if) the thermodynamically or Hadamard stable potential
function F(H1,H2,H3), expressed in terms of principal Hencky strains Hk,
increases more slowly than exponentially.
In the theorem 1.2, principal values of Hencky strain tensor and Finger tensor
for the total deformation are related as"

Hi = (1/2)lnCi, or =/1= (1/2)lnC, trH = 0. (59)

Detailed proofs and definitions are given in the papers [36,90]. While the
theorem 1.1 has been proved for differential CEs as a sufficient condition close
to the necessary one, theorem 1.2 provides the necessary and sufficient
condition for boundedness of single integral CEs.
The above theorems were motivated by the fact that the globally Hadamard
stable upper convected Maxwell model displays the unbounded growth of stress
in simple extension when the elongation rate exceeds the half of the reciprocal
relaxation time. As the consequences of the above theorems, (i) the upper
convected Maxwell model which violates Criterion I, and (ii) the K-BKZ class
with a potential F represented as an increasing rational polynomial function of
basic invariants Iu, are dissipative unstable. Therefore, the Mooney and the neo-
Hookean potentials as well as the potentials for the K-BKZ class of CEs which
are subordinate to Renardy's sufficient evolution criterion also violate Criterion
I of dissipative stability. Since the satisfaction of only Criterion I cannot prevent
the viscoelastic CEs from severe dissipative instability, an additional criterion
for dissipative stability has been introduced.

3.3.2 Criterion II of dissipative stability [88]


For the stability of Maxwell-like and time-strain separable single integral CEs,
it is necessary that both the steady flow curves in simple shear and in simple
elongation have to be monotonously and unboundedly increasing with respect
to the strain rate.
It has been demonstrated [91] that the violation of Criterion II results in
"blow-up" instability or even negative principal values of tensor __c in simple
shear. Therefore the subordination to the combined criterion "I+II" was
assumed in [88] to be presumably sufficient for the dissipative stability of both
the differential Maxwell-like and the time-strain separable single integral CEs,
at least in simple flows.
548

3.4. Application to viscoelastic CEs. Discussion


Both the Hadamard and dissipative types of instability for such two broad
classes of viscoelastic CEs have been discussed in this Section. These are the
quasi-linear differential and factorable single integral models with instantaneous
elasticity, which are the only ones in practical use today. The problem of global
Hadamard stability for these two classes of CEs seems to be completely
resolved in the isothermal, incompressible and compressible cases. This
problem was reduced to that well known in the nonlinear elasticity, where the
complete set of necessary and sufficient conditions of stability was formulated
in algebraic form. It has been demonstrated that the proposed analysis of
Hadamard stability for the two classes of viscoelastic liquids is reduced to the
analysis of strong ellipticity. The physical sense of this is very evident: the
studies of Hadamard stability involve very rapid disturbances which create only
elastic response in viscoelastic liquids.
In the case of dissipative stability, the global analysis is far from being
completed, if it is generally possible. However, two distinct patterns of
dissipative instability have been revealed, which are related to (i) the
boundedness of stress, free energy and dissipation in a start-up flow problem
under a given strain history (Criterion I), and (ii) the monotonously and
unboundedly increasing steady flow curves in simple shear and simple
elongation (Criterion II). Furthermore, it was assumed that the subordination of
CEs to the combined criterion "I + II" is presumably sufficient for the
dissipative stability in the simple flows.
There is a tough problem as to how to distinguish the unstable behaviors
caused by poor modeling of CEs and the observed physical instabilities which
those equations should also describe. However, the long history of various
branches of continuum mechanics and physics teaches us that the occurrence of
either Hadamard instability or/and ill-posedness in ID situations without such
important physical reasons as phase transitions, etc., is a distinct sign of
inappropriateness in the CEs. Thus we can treat the instabilities demonstrated in
this section as being associated not with the real instabilities observed in flows
of polymer melts, but rather with the improper modeling of various terms in
CEs. In numerical simulations of complex flows with unstable CEs, when the
flow rate becomes high enough, the occurrence of various types of unphysical
instabilities is inevitable. Even in the range of moderate Deborah numbers, the
existence of singular points in flow geometry such as the comer singularity in
die entrance region, is sufficient to spoil the entire numerical procedure.
549

All the results of the stability analyses found in various studies for popular
viscoelastic CEs, are summarized in Table 1. An interested reader can find the
details of calculations in references also provided in the Table 1. It is
noteworthy that CEs derived from molecular approaches such as the Larson and
the Currie models, exhibit the most unstable behavior. Surprisingly enough,
none (to the authors' knowledge) of the time-strain separable single integral
models are evolutionary. Appendix A represents the explanation of the reasons
for that given by Simhambhatla [94]. He analyzed the time-strain separability
concept for CEs and concluded that the Hadamard unstable CEs of time-strain
separable type cannot properly describe the experimental data of stress
relaxation after step-wise loading. The instabilities revealed in Ref. [94] exactly
correspond to the results reviewed in this Section. It is astonishing that many
CEs become Hadamard unstable even in viscometric flows.
For the CEs of differential type, only three stable specifications exist. These
are the FENE, the upper convected Phan-Thien-Tanner models, and the author
class of CE's (8), (26) under convexity constraints (27). However, the FENE
and the upper convected Phan-Thien-Tanner models predict zero value for the
second normal stress difference in simple shear flow, which contradicts the
experimental evidence for polymer melts and concentrated polymer solutions.
It should be noted that all the necessary and sufficient conditions obtained for
single-mode CEs become, strictly speaking, only sufficient for the multi-modal
approach. Even though the necessity is not proved, it is thought that due to the
inequalities (29), i.e. well separateness in the relaxation times, the exact
conditions for Hadamard stability exposed above for a single mode CE, will be
closed to necessary for multi-mode approach. It is also evident that the
threshold of instability would only be delayed to some higher Deborah number
region in the multi-modal approach, if any single-mode is unstable.
For some viscoelastic CEs, regularization of ill-posedness may be achieved.
E.g., it is well-known that adding a small Newtonian term to the stress stabilizes
Hadamard unstable CEs. However, for complex flow simulations, this may not
be enough to suppress numerical instability, and when the Newtonian term
becomes larger, the description of the CE will deviate from the experimental
data. In the case of Hadamard stable but dissipative unstable CEs which violate
the Criterion II, one can also propose the more fundamental procedure of
stabilization by changing the elastic potential. For example, the Giesekus model
with or<l/2 in equations (24) will become evolutionary in both global
Hadamard and dissipative senses, if the potential is assigned to satisfy the
convexity constraint (27).
550

Table 1 shows that the combined stability criteria impose very tough
constraints on viscoelastic CEs. Therefore, the serious question arises as to
whether there exists a CE or a class of CEs which can properly describe all the
available rheometric data for concentrated polymer solutions and melts, when
satisfying all the stability constraints. We will describe such a class in the next
Section.

Table I
Stability of viscoelastic constitutive equations
Model (Eq. #) Type of CE Type of instability References
Upper convected Quasilinear Dissipative e.g.[36]
Maxwell (20) (~=1) differential unstable(Criterion I)
Interpolated Quasilinear Hadamard unstable [36,75,76,80]
Maxwell(20) differential (~,1)
(Johnson/Segalman
Gordon/Schowalter)
General Phan-Thien/ Quasilinear Hadamard unstable [36,75,76,80]
Tanner (21) differential (~.l)
Upper convected Quasilinear Hadamard stable; [88]
Phan-Thien-Tanner differential dissipative stability
(21) (~=1) depends on
dissipative term
White-Metzner(22) Nonlinear differential Hadamard unstable, [75,81]
dissipative unstable
(Criterion I)
FENE(23) Quasilinear Globally Hadamard [88]
differential and dissipative stable
Giesekus(24) Quasilinear Dissipative unstable [91]
differential (Criterion II)
Simplest Leonov(26) Quasilinear Dissipative unstable [91]
(bl, = b 2 = 1) differential (Criterion II)
Leonov class (26) Quasilinear Globally Hadamard [88]
under stability and differential dissipative stable
constraints (27)
Larson (25) Quasilinear Hadamard unstable [88]
differential
dissipative unstable [91]
(Criterion II)
Wagner I (38) Separable single Hadamard unstable [88]
integral
Wagner II (39) Separable single Hadamard unstable [88]
integral
Papanastasiou (40) Separable single Hadamard unstable [90]
(K=0) integral
551

Luo-Tanner (40) Separable single Hadamardunstable [88]


integral
Lodge(41) Separable single Dissipative unstable [90]
integral K - B K Z (CriterionI)
K-BKZclassunder Separable K-BKZ Dissipative unstable [90]
Renardy's condition (Criterion 1)
Larson-Monroe Separable K-BKZ Hadamardunstable [88]
potential (42)
dissipative unstable [90]
(Criterion II)
Currie potential (43) Separable K-BKZ Hadamard unstable [88]
dissipative unstable [95]
(Criterion II)
Yen-Mclntire (44) Quasi-separable Dissipative unstable [90]
K-BKZ (Criterion 1)

4. MODELING OF POLYMER FLUIDS WITH STABLE


CONSTITUTIVE EQUATIONS

Following mostly the paper [40], we demonstrate in this Section the class of
differential CEs [23,24,36] which is able to consistently describe simple flow
data for such basic polymers as HDPE II, PS I, PIB P-20, PIB L-80 and LDPE
Melt I/ IUPAC A/ IUPAC X, while complying with the global isothermal
stability constraints. For simplicity, only one or two nonlinear parameters
additional to the discretized linear viscoelastic spectra are introduced for the
description of data.
Instead of the simple "Leonov model" which uses only the parameters of the
discretized linear viscoelastic spectrum, we employ in the following sections a
highly nonlinear specification of the general class of Maxwell-type CEs
proposed by the author [23,24,36]. This specification subordinated to the
convexity conditions (27) guaranties the both Hadamard and dissipative
stabilities. Comparison of the descriptive ability of other models with
experimental data is not attempted in this Section.

4.1. Selection o f a descriptive subclass from the author's CEs


For the sake of simplicity, only the constitutive modeling of viscoelastic
liquids with incompressible Maxwell-type CEs is considered below, with the
equations shown for a single relaxation mode. The model equations are given
by equations (2), (8), (26) and (27).
552

We consider firstly the modeling of dissipative terms in evolution equations


(8) and (26). The following simple forms of equation (26) have been considered
[40]:

i) b~ = b(I~,/2,T)/20(T), b2 = 0;

ii) b~ = 0, b2 = b(I~,/z,T)/20(T); (60)

iii) b~ = b2 = b(I~,Iz,T)/20(T).

Here, 0 = 0(T) is the relaxation time in the linear Maxwell limit. The
specification i) with the Neo-Hookean elastic potential, results in a decreasing
branch of the flow curve in simple shear for b = 1. This poor quality, resulting
in dissipative instabilities, can be rectified by specifying more sophisticated
functional dependencies either for b(I~,I2,T), or the elastic potential. However,
this form was rejected due to the inconvenience for practical modeling. With the
specification ii), various simple flows can be described quite accurately for
several polymers. The only problem with ii) is the weak maxima of Nl predicted
during start up shear flow, in comparison with experimental observations. No
way of rectifying this malady was found and we therefore reject this
specification also. The form iii) for which the evolution equation (8), (26) can
be written as:

20(T)c +b(I, ,12,T) [c2 + c.(I2-I,)/3 - __6]= __0 (61)

is the only one discussed and tested in the literature. It allows for plane
deformations in simple shear and endows the resulting equations with the
proper quality and flexibility for modeling a wide variety of polymers. Detailed
comparisons with experimental data have been made [40] with this
representation.
In order to relate the extra stress tensor to the elastic Finger tensor during the
deformation history, a functional form for the elastic potential W(I1,/z,T) = poF
must be provided. Here, Po is the density, and F is the specific Helmholtz free
energy. The following fairly general elastic potential:

3G(T)
W(II,I2,T) - 2(n+ 1) {( 1-[3)[(I~/3)n+1-1] + (1-13)[(I,/3)n +'-1 ], (62)
553

has been suggested in [40]. Here G(T) is linear Hookean elastic modulus, and 13
and n are numerical parameters. With the evolution equation (61) and potential
relation (62), the constitutive equations are Hadamard and dissipative stable for
0_< 13 <1 and n >0 (see the sufficient condition O for Hadamard stability and
criterion 'I +II' for dissipative stability) except when the criteria for "fluidity
loss" are met (see below). Equation (62) yields the Mooney potential for n = 0,
and the neo-Hookean potential for n = 13= 0. The extra stress tensor can then be
written due to equation (18) as:

flex = (1-13) (I1/3) n __C-l] (12/3) n C"1. (63)

In the multi-mode approach employed below the same values of nonlinear


constants for each mode are used. This gives a few-parametric description of the
data. Also, the conditions of Hadamard and dissipative stability used for every
mode will obviously satisfy the stability criteria for the complex model.
For modeling isothermal experimental data, we need specifications for the
function b(1t,12), as well as a simple reduction of the potential W(It, I2) suggested
in equation (62). For practical modeling, b(I1,/2) can be thought of as a
deformation-history dependent scaling factor for the linear relaxation times. The
simplest choice is to let b = 1, which is known as the standard "Leonov model".
Henceforth, the CE with b=l and the neo-Hookean potential (n = 13= 0) is
further referred as the "simple model". While this simple choice assures the
proper linear viscoelastic limit, and can also be expected to describe weak
nonlinearities, it may not suffice for the description of highly nonlinear
phenomena. For instance, some polymers (e.g., LDPE with a high degree of
branching) show great hardening in simple elongation, while others (e.g.,
HDPE, PS) do not. In any simple flow, if b(Ii,/2) is chosen to decrease gradually
(e.g., using weak power law) with an increase in the magnitude of 11 and 12,
there will be hardening relative to the simple model. A rapid fall in b(Ii,/2) (e.g.,
exponential) will cause the steady state components of c to be double-valued up
to a critical value of the strain rate, with one stable branch. Beyond this critical
value, there is no steady-state solution and the components of c increase
unboundedly as in an elastic solid. This is the concept of "fluidity loss"
analyzed in detail by the author [23,32]. On the other hand, an increasing b(Ii,/2)
will cause relative softening. The more rapidly b increases, the more gradual
will be the variation of the steady-state value of c_with an increase in the strain
rate.
The choice of b(I~,/2) is not as difficult as it appears. The recommended
procedure is to first perform some preliminary calculations for various flows
554

with the simple model. Then, if there is disagreement with experimental data, an
appropriate functional form of b function required to bring the calculations into
qualitative agreement with the data, can be systematically determined. The
reason this procedure is straightforward is that for this class of equations, the
effects in various flows are well separated in the sense that there is some
measure of flexibility in modeling their qualitative behavior independently. This
is a direct consequence of the fact that whether the simple deformations are
steady or non-steady, the following relationships hold for the invariants of the
elastic Finger tensor c:

I~ = 12 in simple shear and planar elongation

11 > 12 in simple elongation, and (64)

11 < I2 in equi-biaxial extension.

These relations also hold for the invariants of the total Finger deformation
tensor C. This is a remarkable feature of the evolution equation (61). The
experience [40] in modeling the viscoelastic behavior of several polymers
(LDPE, HDPE, PS, low and high molecular weight PIB) with the function
b(I~,/2) showed that simple power law or exponential functions of the invariants
with one or two adjustable parameters are sufficient for an accurate quantitative
description of all the available data. However, unlike the elastic potential (62),
no any unified form for b(I1,/2) has been found which could be used for the
description of data for all polymer melts.
Because of the flexibility which the modeling of the dissipative term permits,
one can operate with fairly simple forms of the general elastic potential (62),
such as the neo-Hookean and Mooney potentials for the description of the usual
rheometrical data. However, in simple shear, with this approach, cl2 generally
reaches a limit value of unity at high Deborah numbers leading to a saturation
of the shear stress and therefore to dissipative instability. One remedy is to
extend the discrete relaxation time spectrum at the small relaxation time end in
order to effect stability until the region where physical instabilities appear. A
simpler approach is to work with the existing discrete relaxation modes
obtained from the usual linear experiments, and to allow the parameter n in the
potential relation (62), to be a small positive number (e.g., n = 0.1). In this way,
we can preserve all the predictions in the rheometric regimes with simple
potentials and also effect the non-saturation of the stress at very high Deborah
numbers. Still, for concrete recommendations for quantitatively modeling high
555

Deborah number flows, more data in the region of incipient physical


instabilities would be welcome.

4.2. Component equationsfor simple flows


For convenience, the equations and initial conditions for simple flows for a
single Maxwell mode are presented below. These equations and formulae
should be employed in a multi-mode approach for all the predictive calculations
in comparison with data.

4.2.1. Simple shear


The evolution equations take the form"

20dCll /dt + b(I)(c~ + c 212- 1) - 4,~0c 12


20dc12 / dt + b(I)Cl2 (Cll + C22 ) -- 2~C22 (65)
Cl.C..,
..-l+c 212 9~ I 1 -I 2 -I-1 + c 1 + C
1 22 9

The system of stresses is:

cy12 - G(I / 3) n C12 ; N 1 - G(I / 3) n ( e l l _ C2 2 ) ;

N 2 - G(I/3)n[(1 -- ]3)(C22 -- 1)+ ]3(1-- C11)] (66)

Here, ~, is the shear rate, O"12 is the shear stress, Nl and N2 are the first and
second normal stress differences, respectively.

Startupflow
The initial conditions for startup flow are:

cijl t=o- 8ij. (67)

The steady-state solution of equations (65) is of the form"

Cll -- - ~ Z / 4 Z + 1; r - ~ / 4 z + 1; Ol). - ~ / z ~ - 1 / ( z + 1);


Z ( I ) - (I / 2)(I - 1) 2 - 1. (68)
556

Here, I can be obtained by solving the implicit equation"

Z ( I ) - 41 + (20~ / b(I)) 2 . (69)

Stress relaxation
Here, ~, = 0. For stress relaxation following cessation of shear flow, the initial
conditions for equations (65) are:
c,j[ t_-o- cij]v,t , (70)

where ts is the shearing time prior to cessation of flow. For the stress relaxation
following the imposition of a step strain 7o, the initial conditions for equations
(65), are:
ell ]t = O * - 1 + 7 2.
o, ]
c22 t-o* -1; c 12 ]t=o* -70 9 (71)

Creep and recovery


For creep, let a shear stress ,:3-~ be applied at time t = 0. Let the shear strain at
time t = +0 after the jump be %. Then the initial conditions for the kinematic
variables are given by equation (71). The value of ~,o is obtained by substituting
for C12 from equation (71) into equation (66) and setting O"12 - - CY~ For the
multi-modal case, ~ ~2 is the total stress defined as the sum of the sub-stresses in
the various modes. Then the shear rate, ~; is calculated by setting dcrlz/dt = 0 in
equation (66) and substituting for the time derivatives of the kinetic variables
from equation (65). In the multi-modal case, r~2is the total shear stress.
Evidently,

d 7 / dt - ~ (7] t=0" - 70). (72)

For N modes, the set of 2N+ 1 differential equations, two per mode in equation
(65) and one in equation (72), can then be solved by using e.g., a Runge-Kutta
scheme with an automatically adapting step size. The conditions for recovery
following unloading can be similarly derived.

4.2.2. Simple elongation


The evolution equation is:

X-~dX/dt+b(I~,I2)(X 2-X-~)/(6~.0)-~;; I~-~2+2X-~ 12


_ ~-2 +2~.
557

(73)

where ~ is the elongation rate. The elongation stress is:

G"E = G [ ( 1 - 13)(1, / 3) n (X 2 - X-') + 13(I2 / 3)(X - X-2 )]. (74)

Startup flow
The initial condition for startup flow is:

Lit_ o - 1. (75)

The steady-state solution for X is obtained from the implicit equation:

b(I, ,I 2 )(~2 _ X-' )(1 - k-') = 6/~0. (76)

Stress relaxation
Here, ~=0. For stress relaxation following cessation of elongation flow, the
initial conditions for equations (73) is:

~l t=0-- ~l+,t , (77)

where te is the time of extension before cessation of flow.


For stress relaxation following the imposition of a step Hencky strain eo, the
initial condition is:

)~l t=o" - exp(eo)" (78)

The procedure for creep and recovery calculations is analogous to that for
simple shear.

4.2.3. Planar extension


The evolution equation is:

~-' d~ / dt + b(I)(X 2 - ~-2) / (40(T)) = ~p; I, - 12 - I - )~2 + )C2 + 1. (79)

Here ~p is the planar extension rate. The system of planar extension stresses is
558

o p, - 0,, - ~22 - G(I / 3)" ( 1 - 213)(X2 - )v-2)


0 p, - 0 22 - 0 33 - G(I / 3)" [13)v2 + ( 1 - 13))v2 - 1]. (8o)
Startup flow
The initial condition for startup flow is"

)q t - o - 1 . (81)

The steady-state solution for 9~ is obtained from the equation:

b(I)(L 2 - X-2) - 4~p0. (82)

Stress relaxation
Here, /~p= 0. For stress relaxation following cessation of planar extension
flow, the initial condition for equation (79) is:

~1t=0-- ~l tp,~p, (83)

where tB is duration of biaxial extension before cessation of flow. The initial


condition for stress relaxation following the imposition of a step biaxial
Henckey strain eBo may be written as:

~1 t=O -- exp(e Po ). (84)

Creep and recovery calculation can be performed as for the other simple flows.

4.2.4. Equi-biaxial extension


The evolution equation is:

L-'dX / dt + b(I,,I2)(X 2 - ~-,)(~2 + 1) / ( 1 2 0 ) - ~B;


I~ - 2X 2 + X -~" I2 - 2X-2 + X~ ~ (85)

Here, ~ B is the biaxial extension rate. The biaxial stress is"

~B - G[(1 - 13)(1, / 3) n (X: -- ~-4 ) ..1_ ]~(i 2 / 3) n (~4 __ X-2 )]. (86)
559

Startup flow
The initial condition for startup flow is"

)q t=o- 1. (87)

The steady-state solution for L is obtained from the equation:

b(I,,I2)(~, 2 - X-4)(X2 + 1 ) - 12/~,0. (88)

Stress relaxation
Here, ~B= 0. For stress relaxation following cessation of biaxial flow, the
initial condition for equation (85) is"

~] t=0-- )L]g,,tB, (89)

where tB is duration of biaxial extension before cessation of flow. The initial


condition for stress relaxation following the imposition of a step biaxial
Henckey strain eBo may be written as"

XIt-o - exp(~:Bo)" (90)

Creep and recovery calculation can be performed as for the other simple flows.

4.3. On the comparison with experimental data


High density polyethylene HDPE-II, polystyrene PS-I, polyisobutylene (PIB)
P-20, a relatively high molecular weight PIB, Exxon Vistanex L-80, and low
density polyethylene LDPE Melt-I have been chosen in paper [40] to compare
the predictions of the above constitutive equations with data. A numerous
amount of data and calculations involved in the comparison demonstrated
generally a great success in our modeling. The interested reader can find a lot
of useful details in Ref.[40].
For four first polymers in the tested group, the specification of the function
b(Ii,I2) was uniform and proposed as follows:

b(h) = exp[m(/j3 - 1)]. (91)


560

Eq.(91) means that the four first tested polymers demonstrate softening
behavior at high strain rates.
For describing the data for LDPE Melt I with the author's class of CEs, some
preliminary calculations were first performed with the simple model, as
suggested in the section 3.1 [40]. With this approach, the description of shear
flows was quite accurate. However, the biaxial extension damping function was
overpredicted, while the hardening effects in extension flow were
underpredicted. To rectify the observed discrepancies with the simple model,
the relations (60) suggest that b(11,I2) should be a decreasing function of (IJlz)
(see section 3.1 for the physical sense of the function b). A simple choice was
made as:

b(11,I2) = (I2/11) '~ . (92)

The parameter 'm' was chosen to be 1.4, for properly describing the extension
stress growth data. It was a hope [40] that with formula (93) for b(I1,I2) all the
available experimental data could be described reasonably well. Indeed, the
calculations according to this choice of b(I1,I2) could describe properly almost
all the data for the Melt I but they failed to describe the planar elongation tests
[107]. The reason for this was that in the planar elongation, as in the simple
shear, 11 = 12. Thus the hypothetical rheological behavior in the planar
elongation, as predicted by formula (92), is softening. However, this prediction
contradicts the hardening phenomena in planar elongation, observed
experimentally [ 107]. Therefore in Appendix B, a new, more physically related
attempt is presented to describe the whole set of data for LDPE Melt I.

5. CONCLUSIONS

The behavior of two common classes of viscoelastic constitutive equations


(CEs) for polymer melts and concentrated polymer solutions was discussed.
These are general Maxwell-like and single integral CEs with instantaneous
elasticity.
The formulation of both classes of CEs was analyzed. The Maxwell-like CEs
usually employ some hidden tensor variables with different physical senses.
Therefore, in spite of the generality in formulation, their evolution equations
and stress relations have different features, depending on the theoretical
approach used. Some artifacts related to formulation of the CEs were also
exposed. Such an important effect as compressibility was discussed.
561

General results on stability for both classes of CEs were demonstrated, which
included stability analyses of both the Hadamard and dissipative types. Results
of the stability analyses were applied to popular CEs.
The descriptive capability of a class of Maxwell-like CEs was demonstrated,
whose formulation satisfies all the stability constraints.
It should be noted that the data [98] for equi-biaxial extension were obtained
with using lubricated squeezing technique. Recent publications [113,114]
reported that experiments with this technique can involve undesirable side
effects, such as distortion of sample shape [113] or uncontrolled thinning of
lubricant layer [114]. Therefore there is still a need for independent equi-
biaxial extension data for polymers.

REFERENCES

~ Odroyd, Proc. Roy. Soc., A200 (1950) 1063.


2. Oldroyd, Proc. Roy. Soc., A245 (1958) 278.
3. Rivlin, Research Frontiers in Fluid Dynamics, 144, Interscience, New
York, 1965.
o Truesdell and W. Noll, The Non-Linear Field Theories of Mechanics,
Springer, New York ,1992.
. Coleman, Arch. Rat. Mech. Anal., 17 (1964) 1.
6. Coleman, Arch. Rat. Mech. Anal., 17 (1964) 230.
7. Coleman and V.J. Mizel, Arch. Rat. Mech. Anal., 29 (1968) 105.
8. de Gennes, Scaling Concepts in Polymer Physics, Cornell University,
Ithaka, New York, 1979.
, Green and A.V. Tobolsky, J. Chem. Phys., 14 (1946) 80.
10. Lodge, Trans. Faraday Soc., 52 (1956) 120.
11. Yamamoto, J. Phys. Soc. Japan, 11 (1956) 413; 12 (1957) 1148; 13
(1958) 1200.
12. Treloar, Physics of Rubber Elasticity, 3rd ed., Oxford University Press,
Oxford, 1975.
13. Edwards, Proc. Phys. Soc., 92 (1967) 9.
14. Doi and S.F. Edwards, J. Chem. Soc., Faraday Trans. II, 74 (1978) 560,
918, 1818; 75 (1978) 38.
15. Curtiss and R.B. Bird, J. Chem. Phys., 74 (1981) 2016.
16. Bird and J.R. Deaguiar, J. Non-Newton. Fluid Mech., 13 (1983) 149, 161.
17. Curtiss, R.B. Bird and O. Hassager, Adv. Chem. Phys., 35 (1976) 31.
18. Bird, R.C. Armstrong and O. Hassager, Dynamics of Polymeric Liquids,
Vol. I & II, 2nd Ed., John Wiley and Sons, New York, 1987.
562

19. Pokrovsky and V.S. Volkov, Visokomolekulamie Soed. (Russian), A20


(1978) 2700.
20. Volkov, Int. J. Polym. Mater., 9 (1982) 115.
21. Schweizer, J. Chem. Phys., 91 (1989) 5802, 5822.
22. Schweizer, J. Non-Cryst. Solids, 131-133 (1991) 643.
23. Leonov, Rheol. Acta, 15 (1976) 85.
24. Leonov, J. Non-Newton. Fluid Mech., 25 (1987) 1.
25. Dashner and W.E. Vanarsdale, J. Non-Newton. Fluid Mech., 8 (1981) 59,
69.
26. Larson, Constitutive Equations for Polymer Melts and Solutions,
Butterworth, Boston, 1988.
27. Hunter and M. Slemrod, Phys. Fluids, 26 (1983) 2345.
28. Ramamurthy, J. Rheol., 30 (1986) 337.
29. Tanner, Engineering Rheology, Clarendon Press, Oxford, 1992.
30. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefringence,
Springer, New York, 1983.
31. Pearson, Mechanics of Polymer Processing (Part I: Polymers, Mechanics
and Rheology), Elsevier, New York, 1985.
32. Leonov and A.N. Prokunin, Nonlinear Phenomena in Flows of
Viscoelastic Polymer Fluids, Chapman & Hall, London, 1994.
33. Grmela, Physica, D21 (1986) 179.
34. Grmela, Phys. Lett., A 130 (1988) 81.
35. Beris and B.J. Edwards, J. Rheol., 34 (1990) 55,503.
36. Leonov, J. Non-Newton. Fluid Mech., 42 (1992) 323.
37. Jongschaap, Physica, A94 (1978) 531.
38. Jongschaap, K.H. de Haas and C.A.J. Damen, J. Rheol., 38 (1994) 769.
39. Kwon and S. Shen, Rheol. Acta, 23 (1984) 217; 24 (1985) 175.
40. Simhambhatla and A.I. Leonov, Rheol. Acta, 34 (1995) 259.
41. Leonov, Polym. Internat., 36 (1995) 187.
42. De Groot and P. Mazur, Non-Equilibrium Thermodynamics, North-
Holland, Amsterdam, 1962.
43. Gyarmati, Non-Equilibrium Thermodynamics. Field Theory and
Variational Principles, Springer, New York, 1970.
44. Hulsen, J. Non-Newton Fluid Mech., 38 (1990) 93.
45. Murnaghan, Amer. J. Math., 59 (1937) 235.
46. Leonov, Ann. N.-Y. Acad. Sci., 410 (1983) 23.
47. Stickforth, Rheol. Acta, 25 (1986) 447.
48. Leonov, On the constitutive equations for non-isothermal bulk relaxation,
accepted for publication in Macromolecules, 1997.
563

49. Gordon and W. R. Schowalter, Trans. Soc. Rheol., 16 (1972) 79.


50. Johnson, Jr. and D. Segalman, J. Non-Newton. Fluid Mech., 2 (1977)
255.
51. Phan Thien and R. I. Tanner, J. Non-Newton. Fluid Mech., 2 (1977) 353.
52. Phan Thien, J. Rheol., 22 (1978) 259.
53. Larson and K. Monroe, Rheol. Acta, 23 (1984) 10.
54. White and A. B. Metzner, J. Appl Polym. Sci., 7 (1963) 1867.
55. Giesekus, Rheol. Acta, 21 (1982) 366.
56. Larson, J. Rheol., 28 (1984) 545.
57. Simhambhatla and A.I. Leonov, Rheol. Acta, 32 (1993) 259.
58. Rivlin and K.N. Sawyers, Ann. Rev. Fluid Mech., 8 (1971) 17.
59. Kaye, College of Aeronautics, Cranford, U. K., Note No. 134 (1962).
60. Bernstein, E. A. Kearsley and L. J. Zapas, Trans. Soc. Rheol., 7
(1963)391.
61. Kwon and A.I. Leonov, Rheol. Acta, 33 (1994) 398.
62. Wagner, T. Raible and J. Meissner, Rheol. Acta, 18 (1979) 427.
63. Wagner and A. Demarmels, J. Rheol., 34 (1990) 943.
64. Luo and R.I. Tanner, Int. J. Num. Meth. Eng., 25 (1988) 9.
65. Papanastasiou, L. E. Scriven and C. W. Macosko, J. Rheol., 27 (1983)
387.
66. Larson and K. Monroe, Rheol. Acta, 26 (1987) 208.
67. Currie, in G. Astarita, G. Marrucci and L. Nicolais, "Rheology", Vol.1,
Plenum, New York (1980).
68. Yen and L.V. McIntire, Trans. Soc. Rheol., 16 (1972) 711.
69. Zapas, J. Res. Natl. Bur. Std., 70A (1966) 525.
70. Ogden, Proc. Roy. Soc., A326 (1972) 565.
71. Mooney, J. Appl. Phys., 11 (1940) 582.
72. Valanis and R.F. Landel, J. Appl. Phys., 38 (1967) 2997.
73. Blatz, S.C. Sharda and N.W. Tschoegl, Trans Soc. Rheol., 18 (1974) 145.
74. Crochet, A.R. Davies and K. Walters, Numerical Simulation of Non-
Newtonian Flow, Elsevier, Amsterdam, 1984.
75. Dupret and J.M. Marchal, J. Non-Newton. Fluid Mech., 20 (1986) 143.
76. Joseph, Fluid Mechanics of Viscoelastic Liquids, Springer, New York,
1990.
77. Rutkevich, J. Appl. Math. Mech., 33 (1969) 30, 573; 34 (1970) 35.
78. Godunov, Elements of Continuum Mechanics, Nauka, Moscow, 1978.
79. Joseph, M. Renardy and J.C. Saut, Arch. Rat. Mech. Anal., 87 (1985)
213.
80. Joseph and J.C. Saut, J. Non-Newton. Fluid Mech., 20 (1986) 117.
564

81. Verdier and D.D. Joseph, J. Non-Newton. Fluid Mech., 31 (1989) 325.
82. Kreiss, Numerical Methods for Solving Time-Dependent Problems for
Partial Differential Equations, Les presses de l'Universite de Montreal,
Montreal, 1978.
83. Strang, J. Diff. Eq., 2 (1966) 107.
84. Baker and J.L. Ericksen, J. Wash. Acad. Sci., 44 (1954) 33.
85. Zee and E. Sternberg, Arch. Rat. Mech. Anal., 83 (1983) 53.
86. Knowles and E. Sternberg, Arch. Rat. Mech. Anal., 63 (1977) 321.
87. Kwon and A.I. Leonov, J. Non-Newton. Fluid Mech., 47 (1993) 77.
88. Kwon and A.I. Leonov, J. Non-Newton. Fluid Mech., 58 (1995) 25.
89. Kwon, J. Non-Newton. Fluid Mech., 65 (1996) 151.
90. Kwon and A. I. Leonov, Rheol. Acta, 33 (1994) 398.
91. Kwon and A. I. Leonov, J. Rheol., 36 (1992) 1515.
92. Hulsen, J. Non-Newton. Fluid Mech., 38 (1990) 93.
93. Renardy, Arch. Rat. Mech. Anal., 88 (1985) 83.
94. Simhambhatla, The Rheological Modeling of Simple Flows of Unfilled
and Filled Polymers, Ph.D. Dissertation, the University of Akron, 1994.
95. Kwon, Studies of Viscoelastic Constitutive Equations and Some Flow
Effects for Concentrated Polymeric Fluids, Ph.D. Dissertation, The
University of Akron, 1994.
96. Einaga, K. Osaki, M.Kurata, S. Kimura, and M. Tamura, Polym. J., 2
(1971)550.
97. Laun, Rheol. Acta, 17 (1978) 1.
98. Khan, R.K. Prud'homme and R.G. Larson, Rheol. Acta, 26 (1987) 144.
99. Soskey and H.H. Winter, J. Rheol., 28 (1984) 625.
100. Takahashi, K. Taku, and T. Masuda, J. Soc. Rheol., Japan, 18 (1990)18.
101. Laun, J. Rheol., 30 (1986) 459.
102. Leonov and A.N. Prokunin, Rheol. Acta, 19 (1983) 137.
103. Laun, Stress and recoverable strains of stretched polymer melts and their
prediction by means of a single integral constitutive equation. In"
Rheology, vol.2, Plenum Press, New York (1980).
104. Munstedt and H.M. Laun, Rheol. Acta, 20 (1981) 211.
105. Vinogradov and A.Ya. Malkin, J. Polym. Sci. A-2, 2 (1964) 2357; 4
(1966) 135.
106. Leonov and A.N. Prokunin, Rheol. Acta, 19 (1980)393.
107. Laun and H. Schuch, J. Rheol., 33 (1989) 119.
108. Wagner, J. Non-Newton. Fluid Mech., 4 (1978) 39.
109. Laun, Rheol. Acta, 21 (1982) 464.
110. Wagner and H.M. Laun, Rheol. Acta, 17 (1978) 138.
565

111. Giacomin, R.S. Jeyaseelan, T. Samurkas and J.M. Dealy, J. Rheol., 37


(1993) 811.
112. Meissner, Trans. Soc. Rheol., 16 (1972) 405.
113. Takahashi, T. Isaki, T. Takigava, and T. Masuda, J. Rheol., 37 (1993)
827.
114. Kompani, D.C. Venerus, and B. Bernstein, Development and evaluation
of lubricated squeezing flow technique. In: Proc. XIIth Int. Congr. on
Rheology, A. Ait-Kadi, J.M. Dealy, D.F. James, and M.C. Williams, Eds.
August 18-23, 1996, Quebec City, Canada, p.754.

Appendix A

On the invalidity of strain-time separability at quick time scales


(M. Simhambhatla [94])

Strain-time separability, i.e. the factorability of the material response to


nonlinear step strains into time and strain dependent parts, has been widely used
as a convenient basis for the specification of viscoelastic CEs. The proponents
of this assumption claim justification based on experimental observations.
However, it is shown here that the same experiments for polymer melts and
solutions require that the principle of strain-time separability be violated at
small times following the application of step shear strains, in order to guarantee
the Hadamard stability.
According to strain-time separability, the stress response, to a step shear strain
applied from the rest state is:

o(t,y) : yh(y)G(t), (A 1)

where h(7) is the shear damping function, and G(t) is the linear (Maxwellian)
relaxation function. This means that the stress relaxation curves (log(o(t)) vs.
log(t)) for various applied step strains will be parallel. This observation of
seemingly parallel stress relaxation curves has been reported for several
polymer melts and solutions (e.g., see [96], [97]).
Consider strain-time separable CEs with a perfect elastic limit. For these
equations, the step stress cl in response to a step strain 7, will be

a(7 ) - 7h(7 ) ~ G, (A2)


i=l
566

Here, N is the number of Maxwellian modes, and Gi the linear relaxation


modulus for the 'i'th mode. Now consider the experimentally determined
dependence Th vs. T for LDPE Melt I [98] shown in Figure A1. The maximum
in this dependence implies that 6 vs. T should also have a maximum (equation
(A2)). Such maxima for Th vs. T appear for all the experimental data that we
have come across (e.g., see [96-100]). It is however, easy to see that the
decreasing branch of the dependence 8 vs. T, is unstable in the Hadamard
sense.
In order to clarify this, consider a Cartesian coordinate system (x,y) with the
'x' axis parallel to the direction of the shear displacements 'u', and the 'y' axis
normal to the shear planes. The equation of motion for simple shear can then be
written as"

p0v/0t = 3cr/o~ (A3)

With v - Ou/c~, and 3' = Ou/0y, equation (A3) can be rewritten in the elastic limit
as;

/~U/C~ 2= M(]t)c32u/o~ 2 (A4)

Here 'u' is the displacement in the 'x' direction, and M(T) - c36/0y. Let uo be a
basic solution satisfying equation (A4). Let us impose a small disturbance on
the basic solution, so that

u - Uo + ~ fi exp[i(ky-mt)/a21 (A5)

where ~ is the small amplitude parameter, 'k' the wave number, and co the
frequency of disturbance. Substituting for 'u' from formula (A5) into equation
(A4) and taking into account only the lowest order terms in g, yields"

(1)2 = M(To)k2/p. ( M(y o ) - d r ~ 03, Ir,, ) (A6)

If M(y o) > 0, m is real and the disturbances in equation (A5) will not grow with
time. However, if M(To) < 0, we have:

co - +i~/(]m(y o)lk =/p. (A7)


567

2.0 r - - ......... I ...... ",~ ........... I . . . . . . . [ ......... 1" ....

1.5
J~ " 0
MeltI

v
_L--
1.0

o5 [
o Experirnenfal Poinfs
----- h(7) = ,57exp(--.31T)+.43exp(--.10-67)

L ........... I .... t, . . . . . . . . . . . . . . . . . . . L ................. J ........................................

5 i0 ~ 5.. 20 25 50 55
Shear Stroin, -y

Figure A 1. Experimentally observed maximum in the plot 7h(7)vs. 7 for LDPE


Melt I [96].
When the imaginary part of o is negative, formula (A5) indicates that even
infinitesimal high frequency disturbances will grow in amplitude rapidly and
unboundedly in time. Because there are high frequency disturbances in the
spectrum of natural noise, the decreasing branch of the dependence t~ vs. 7 will
result in severe instabilities during experimental measurements.
However, since the experiments have been conducted for large step shear
strains, and no instabilities have been reported, it is evident that the principle of
strain-time separability should be violated at small times following the
imposition of these large shear strains. This violation should be manifested in
the form of an upturn in the dependence cy(t) vs. t at small times following the
imposition of a step strain 7 when moving along the time axis in the negative
direction, in order to have a monotonic increase in the dependence ~ vs. the
step strain Y. Interestingly, this is preciselywhat was observed by Einaga et al.
[96] (see Figure 2) and Takahashi et al. [100]. In experiments where yh(7) vs.
7 has a maximum, but no loss of parallelism is seen in the curves ty(t) vs. t , it
568

is presumably because data could not been obtained at very short times
following the imposition of the step strains (assuming no wall slip).

........ I .......... ~ ........ ~-

10 4

t0 3

0,,,.
v
o
p..
to2
v
r~
"10
(3)

,J~
101
r

i 011 ................................. I .............. . . . I ...... .. ................ i .. . . . . . .

1 1 0 `= 10 2 10 s 10 4

Figure A2. Illustration of violation of strain-time separability at short times for


20% PS in Arochlor [96].

It should be noted that the deviations from the master curve of vertically shifted
stress relaxation curves are consistent with the requirement of Hadamard
stability.
Two conclusions can be drawn. The first is that strain-time separability holds
only in an approximate sense, if we are prepared to neglect the material
response to quick disturbances. The second is that any strain-time separable CE
is certainly unstable in the Hadamard sense in the limit of very rapid
disturbances, due to the universally observed maximum in the dependence ~,h(~/)
vs. ~,in those time regimes. This indicates that we cannot simultaneously
describe the experimental data in the rheometrical regimes, and have the
stability at high Deborah numbers for any strain-time separable CE. It is evident
that the above instabilities have no physical basis, but are simply rooted in the
improper extrapolation of strain-time separability to quick time scales. High
Deborah number flow simulations using these CEs should therefore be avoided.
569

Appendix B

On rheological description of LDPE Melt I by stable CEs


M. Simhambhatla and A.I. Leonov

According to Exxon material data, LDPE Melt I has molecular weight


Mw=460000 and polydispersity ratio Mw/Mn=22. Linear viscoelastic spectrum
for this material at 150~ is shown in Table B 1.
Table B 1

Linear__v_i.sc0e!ast_.!c_s_pectrum o f,L Dp E Melt i .(.15oOc)[9:7]


i 1 2 3 4 5 6 7 8

0i, sec 10 3 102 101 10 ~ 101 10 .2 10 .3 10-4

G,, Pa 1.00xl0 ~ 1.80x102 1.89x103 9.80x103 2.67x104 5.86x104 9.48x104 1.29X!0!

In order to describe all available data for the LDPE Melt I we will use the
simple non-Hookean potential and two-parametric expression for the dissipative
term b"

sinh[v(I, - 3)1
b(I, ) - exp[-13(I, - 3)] + -1. (0<13 < 1, v > O) (B1)
vO, - 3 )

Here [3 and v are some numerical fitting constants. Formula (B 1) is in fact a


modification of a similar expression proposed in paper [102] to describe
simultaneously the hardening phenomena in the simple extension of LDPE Melt
I with following softening. The first exponential term in (B1) describes the
hardening phenomena in polymer melts due to orientation of macromolecules.
The simple molecular model and explanations are given in paper [23] and book
[32]. The second term in formula (B 1) reflects the softening phenomena in the
flow of high oriented melts as described by the Eyring formula in the activated
rate processes. As shown in paper [102], the softening is attributed to the
thermo-mechanical degradation with a decrease in the molecular mass during
extension flow of an LDPE melt. This was revealed by intrinsic viscosity
measurements performed on the specimens left over from the experiments. This
points to irreversible effects which are outside the scope of a purely rheological
description. Interestingly, Wagner had to make a "structural irreversibility"
570

assumption [108] to properly describe recovery following extension flow for


LDPE Melt I.
To describe simultaneously the flow data for LDPE Melt I, the parameters 13
and v in (B 1) were chosen after a fitting procedure as"

[3 - 0.15, v - 0.03. (B2)

With the parameters shown in (B2), formula (B1) initially show only
hardening and only then, after decay of the first term, it describes "irreversible
softening". It should be also noted that in simple shearing, with typical data for
shear rates available in rotating instruments, the value b in formula (B 1) is all
the time near unity. This fact preserves the description of simple shear data
demonstrated in paper [40]. Nevertheless, we manifest in the following also
some important comparisons between our calculations and data for simple
shearing too.
Figure B 1 indicates a good description of the steady shear viscosity and first
normal stress difference over a wide range of shear rates. Descriptions of the
transient stress growth during shearing experiments are depicted in Figure B2.
The discrepancy for N~ at short times is probably due to the instrumental
problems for these measurements [97].

7 7
10 ~__...~_v . ~ u l ~ l l 10

Melfl o o
I 10 6
I ~eference Temp. 150 C - 10
6

13..
~..,.

o 10
5 10 5 c-,-

~j_~-~7~,,
,--~7_~~^,-, ~
121
O
>:, 4
104 ~
-~
.,..~ 10
"I3
0 Q
I
3 u~
m
5= 103 10 ,-,
t~
ol

c~ 02 _ 9
10 f-]
O

C~
1 O
10 .,. I I I I 1 I I 10 ~.~.

10-4- 10-.3 10-2 10 -1 10 ~ 101 10 2 10 `3 10 4

Reduced Shear R a t e , 7G T (s 1)

Figure B 1 .Steady shear viscosity and the first normal stress coefficient of LDPE
Melt I at reference shear rate of l s -1 [97] at various temperatures (symbols).
571

a 10 5 i i
n Melt l 0
Reference Temp. 150 C **t:~:~O ~
c~ z-

,~

9 -I
~ ~ o3 .-vaT - 1s

r" t..

~ 10 2
I0-' 10 0 10 1 10 2

Shear Strain, 7

Figure B2. Transient stress growth normal stress coefficient for LDPE Melt I
[97,109]. Various symbols correspond to different temperatures.

We now consider shear creep and recovery. The coincidence with


experimental data is good for both the shear strain and N1, for the creep
condition, as shown in Figure B3. Strain recovery and normal stress decay
following unloading after the creep experiment in Figure B3 are shown in
Figure B4. Here, the predictions are accurate for strain recovery but N1 is
underpredicted.
Large amplitude oscillatory shear data, at low to moderate frequencies are
indicative of dissipation during cyclical nonlinear deformation. Figure B5
shows a good agreement with the data obtained by Giacomin et al. [ 111 ] for the
batch-labeled IUPAC X.
We also demonstrate the capability of the model to describe the extension
experiments. Figure B6 demonstrates an excellent agreement between
calculations and data [112] for extension stress growth, with the use of
formulae (B1) and (B2) for simple elongation. Figure B7 shows the
comparison between our calculations and data [107] for the stress growth
coefficient in planar extension.
572

6
10 I I 10

Melt l
o
Reference Temp. 150 C
10 4
_ 105
o "~ - 10 Pa
12
d z
o--
0
N1
10 _ 10 4 ~" U,
121
l-

c"
(/3
10 0 _ 10 3

2
10 -~ u i i i 10
10 -2 10 -1 100 101 102 10 5

Reduced time, f / a T (s)

Figure B3. Shear strain and first normal stress difference in creep under
constant shear stress for LDPE Melt I [110]. Various symbols correspond to
experiments at various temperatures.
102 I
. . . .

I I I
105

1 104
10 - N, t

> z
0
~ 0 _ 10 3 ,~
-la
r ~~ 10
Q
r
o_
El
k_

(z)
10
-1
o 102
Reference Temp. 150 C

10 -2 I I I I 101
10 -2 10 -1 100 101 102 103

Reduced time, f / a T (s)

Figure B4. Shear strain recovery and first normal stress difference after
unloading following the experiment in Figure B3 [110]. Various symbols
correspond to various temperatures.
573

80 I I 1 I I I 1

60 - 70 = 5.0

Frequency = 1Hz
40 _

O~
-~ 2o
(o)
e 0
r~
-20

o~ -40 IUPAC X
o
-60 150 C

-80 I I I I I I I
-40 -30 -20 -10 0 10 20 50 40
-t
Shear Rate (s )

100 I I I I I I I

75
7o = 10.0
OQ
5O
Frequenc =
v 25 _

f/l
I1)
~ 0
(b)

-25 -

&)

rn -50 IUPAC X -
o
150 C
-75 _

I I I I I I I
-100
-80 -60 -40 -20 0 20 40 60 80

-1
Shear R a t e (s )

Figure B5. Large amplitude oscillatory shearing of LDPE IUPAC X [ 111 ].


574

6I
I I I I

10 0.I
t 1.0 0.01

I
0
13_
1
..---2
.+r 10
v

+#

f Melt I (I50 ~
10

I I I I I
10 -I 100 101 102 103 10 4

Time (s)

Figure B6. Tensile stress growth coefficient vs. time for LDPE Melt I [112].

Finally, we also attempted to describe with the use of formulae (B1) and (B2),
the data [98] for equi-biaxial elongation for LDPE Melt I. The result of
comparison between our calculations and data is shown in Figure B8. It is
evident that the calculated curve predicts more hardening in the rheological
behavior of the Melt I as compared to the data. The possible reason for that was
discussed in the Section 5 of the paper's main text.
575

106 I

-1 -1
0.05 s 0.01 s
0
0

0~
E~
13_
5
10

.kp

+o_

Melt I (125 ~C)

4
10

t .. I

100 101 102 103


Time (s)

Figure B7. Planar tensile stress growth coefficient vs. time for LDPE Melt I
[107].

1.0

0
c 07
0
"~ C:
c" 0
0
= 0.5
x

~ l.a.
12I

~ ~o3
IUPAC A (150 ~

.1 '
,_ 9 I

0.1 0.3 0.5 0.7 1.0 3.0


Strain

Figure B8. Equi-biaxial damping function for LDPE Melt I [98].


577

SCALING APPROACH IN SOLVING PROBLEMS OF


COMPLEX VISCOELASTIC FLOWS WITH MULTI-MODE
CONSTITUTIVE EQUATIONS OF DIFFERENTIAL TYPE

A. I. Leonov*, J. Padovan**

*Department of Polymer Engineering


**Department of Mechanical Engineering
The University of Akron,
Akron, OH 44325 - 0301, USA

1. INTRODUCTION

There are two types of constitutive models for nonlinear viscoelastic


phenomena in polymer fluids in use today. One is of the differential and the
other of the single integral type. The differential models are descriptively more
rich and flexible. This is because they are able to independently model both the
dissipative and elastic, non-dissipative terms in constitutive equations (CEs).
However, they usually operate with a set of independent Maxwell modes
generated from discrete points of the linear relaxation spectrum. Since each
independent mode is generally described by a set of six partial nonlinear
differential equations, this approach tends to be computationally intractable
when solving 3D unsteady problems.
Significant efforts were spent through the years to make specific formulations
of the CEs. The interested reader can find many specifications of C E s o f both
differential and integral types in recent books [1,2]. This includes comparisons
with experimental data and useful discussions. Several general approaches to
the derivation of single mode Maxwell-like CEs of the differential type were
recently developed. The first, Poisson bracket formalism, which uses variational
techniques, was introduced by Grmela [3-5] and extensively used by Beris and
Edwards [6,7] for purposes of creating a general unified approach to Maxwell
type CEs. These authors employed a variational derivation of CEs with the
configuration tensor c treated as a hidden variable. In Refs. [7,8], this problem
578

was treated locally by means of non-equilibrium thermodynamics, and the same


results as Beris and Edwards were obtained for Maxwell type CEs. Moreover, in
Ref.[8] a "canonical" formulation of general Maxwell type CEs was proposed
and used for the general stability analysis of CEs (see, e.g. [10]). In the multi-
mode approach, the set of hidden tensor variables {__el,__c2, ..., cN }is usually
employed. The ability of a class of multi-mode CEs (see, e.g. [2]) to describe
consistently simple polymer flows was recently demonstrated [11] when
satisfying all the stability constraints. For the sake of simplicity, functionally
mode independent modeling of dissipative and elastic terms was employed and
successfully used in paper [11 ].
The main goal of this chapter is to demonstrate that for multi-mode CEs,
modal independence immediately results in a "scaling theorem" which in cases
with known flow histories, reduces the solution of any 3D unsteady and non-
isothermal problem to the computations for a single Maxwell mode. Although it
is easy to extend the scaling theorem for the compressible case, for simplicity,
only the incompressible version is considered. This is to avoid discussions [8]
of compressible formulations for many viscoelastic CEs lumped together in the
unified approach.
The chapter is organized as follows. In the next Section 2 we briefly discuss
the general unified formulation [8] for multi-mode approaches. Then in Section
3, some similarity assumptions are made which make it possible to prove the
scaling theorem. In Section 4, two examples illustrate the use of the scaling
approach for reducing computations for complex viscoelastic flows. Some
concluding remarks are presented in Section 5 of the paper.

2. F O R M U L A T I O N

The general unified isotropic formulation proposed in Ref.[8] for N


independent Maxwell modes with the set of hidden variables {Cn}, consists of:
(i) evolution equations for each n-th mode,

On(T)[c~tc n nt- ( v . V ) c n - Cn'(D q- o.1 C_n- ~n(Cn'e + e'c__ja)] q- ~n(~n,Cn) -- 0


(-1 < ~ n < 1) (1)

whose most important feature is that the velocity field v is the same for each
mode;
(ii) the n-th modal extra stress tensor formulation:

e
O'n -- Gn(T)sn(~n, Cn ) (2)
579

and
iii) the expression for the total stress tensor:

c~ - -p~_ + o e e e (3)
= _ -- ' (5" = 'Y'~ (5"
-- --n
n

Here 't' is time, v is the velocity vector, N and e are the vorticity and strain rate
tensors, ~ is the stress tensor, p is the isotropic pressure, fi is the unit tensor,
0n(T)and Gn(T) are the relaxation time and elastic (Hooke's) modulus for n-th
mode, ~,n is a numerical parameter for each mode, and T is temperature. The
elastic terms in the CEs include the evolution operators within the square
brackets in equation (1) and the formulation of the stress tensor with the aid of
equations (2) and (3). Here _Sn (~,,, _C_n)are some isotropic tensor functions of
tensors Cn for each n-th mode. The dissipative terms in the CEs are represented
by the set of isotropic tensor functions ~n(~,, ~ ) of tensors c for each n-th mode.
If the values of parameters n are equal to either -1, or 0, or +1, the evolution
operator in equation (1) converts to the lower convected, corotational, or upper
convected tensor time derivative, respectively. It is also generally assumed that
the above isotropic tensor functions _S_n(r _C_n) and ~. (~n, --C_n) satisfy the
dissipative inequality and provide the CEs with a regular limit to the linear
viscoelastic case.
The momentum balance and the continuity equations are of the form:

plat x + (x._V)x] = Vp + e + pgk, (4)

V.v =0 (5)

Here p is the density and g is the gravity acceleration directed along the unit
vector k.
Finally, there is also the heat equation, which using the entropy viscoelasticity
assumption (e.g., see [2]) takes the form"

Cvp[c3tT + (v.V)T] = V.K-VT + J.tr(~.e). (6)

Here C~ is the heat capacity under constant volume, J is the thermal heat
equivalent of power, and K is the thermal conductivity which is generally
represented as a second rank tensor. In the case of high-elastic cross-linked
580

rubbers, the stress induced anisotropy of thermal conductivity was well


documented [12-14]. For polymeric liquids, it was assumed that K = K(T, =el,
c2,..., ___CN)[ 15]. To the authors' knowledge, the flow induced anisotropy in heat
conductivity has never been tested experimentally, seemingly because of
evident experimental difficulties.
Using equations (1)-(6), we consider the general 3D and unsteady problem for
a flow of visco-elastic liquid in a domain gl c 9t 3, confined in the boundary
Og~= c3g/1 w c3t~2, with the initial and boundary conditions"

t = 0, x e~: v(0,x)=V(x), _C_n(0,x)=fi, T(0,x)=To(x);


(7)
x ec3gll" v=U(T,x), T=T~(t,x); x (Kc~"~2: cy=S(t,x), T=T2(t,x)

Here To, T~ and T2 are known temperature fields, V and U are known solenoidal
vector fields, and S(t,x) is a known boundary stress. We assume below that a
solution of the problem (1)-(7) exists, and will make simplifying assumptions to
the above formulations.

3. SIMILARITY ASSUMPTIONS AND SCALING THEOREM

We now make three similarity assumptions which reduce the set of CEs (1) -
(3) and Eq.(6) to a more simplified form.

Assumption 1. Thermo-rheological simplicity"

0n(T) = Otn0(T), Gn(T) = [BnG(T). (8)

Here an and 13n are numerical factors, and 0(T) and G(T) are some characteristic
relaxation time and modulus, taken e.g. either from the first relaxation mode, or
from a dominant mode, or considered as mean values averaged over the linear
relaxation spectrum. The assumption (8) usually holds for polymer fluids in a
restricted time-temperature region.

Assumpion 2. Mechanical mode independence:

gn: 8,n= 8,; Sn(~,Cn)= S(~,Cn); ~(8,,_C_n) = ~(~,_C__n) (9)

Thus instead of 2N independent functions _Snand ~ , equation (9)just assumes


that there are only two independent non-dimensional isotropic tensor functions
581

s ( ~ , c ) a n d ~ ( ~ , c ) . It requires in particular, that the parameters appearing in these


functions be the same for any n-th mode (e.g., see [ 11 ]).

Assumption 3. Thermal mode independence:

K(T,c_I,e:,...,_C_N)= R~--"K ~( T , C n )~ n / ~ n , R=(~-~ 13n /C/,n )-1 (10)


n n

The foregoing decomposition provides for mode independent thermal effects.


This formulation yields two physically valid limiting cases, namely:
(i) the isotropic heat conductivity,

K = n:o(T)6__, (11)

(ii) the anisotropic heat conductivity for solids with finite elasticity,

K=K(T,C-'). (12)

Here C -~ is the Finger strain tensor. It should be noted that w h e n , r the


evolution equations (1) have the elastic limit: _C_n---~C1. This may occur either for
intense viscoelastic flows, or for a process of solidification (0(T)--+~).
We now prove the following

Scaling theorem.
Assume that the flow history (i.e. the velocity field v(t,x)) is known. Let the
assumptions 1-3 with equations (9)-(11) hold. Assume that for a certain value
n-k, a solution { 7~,_fi,~,_~} has been obtained. It depends on coordinate x and a
"stretched" time ~, and satisfies the s i n g l e m o d e set of equations:

O, ~=+ (fi . V ) . 6=- 8 . ff3 + ff3 . 6 - ~ ( 6 . 8 + 8 . 8) + ~ ( ~ , ~ ) / O(ffZ) = O (13)

6 - 1 / 2[_V__fi + (V____fi)T ], O3 -- 1 / 2[V__fi - (V___fi)T 1, (14)

8 - -~ + G(T)s(_~), (15)

ao[0,T + (_ft. V__)_fi]- V_V_.6 + bgpk, (16)


582

V__._fi- 0, (17)

bC vp[a, T + (_ft. __V)T]= R E . [K(~', a)- VT] + J. G(T)- tr[~(a) .~1, (18)

with the initial and boundary conditions"

- o, x n: __a(O,x)- _9(x), _a(O,x)- 5, T- To (x)


X e~ 1" ~- O(~,X), T- "]-'1(~,X) ; (19)
X E O ~ 2" a- S(l:,X), T- T2('l:,X)

Let the stretched time x, parameters 'a' and 'b', and functions ~, IA, ]71, "I'2'
and ~ in equations (19), be defined as follows"

z-t/ak, 2
a = 1/~--'~[3nan, b _ l / Z ] 3 n, ]Tj (a:,X) - Tj (t / ak , x )
n n
^

V ( x ) - a k ___V(x), lJ(z, x_)- a k U ( t / a k , x ), =S(x,x_)= b S ( t / a k ,_x) (20)

Then the multi-mode solution of equations (1)-(6) for the initial boundary
problem (7) is found as follows:

v(t, x) - __fi(x,_x) / a k, T(t, x) = T(T, x), p(t, x ) - ~(z, x) / b,


c (t,x)-~,('ta
=n =
k / a n ' x), =
e =G(T)~--,13nS( c )
= =n
(a:-t/ak) (21)
n

Preliminary remarks.
1. There are two important features of the above CEs: (i) as mentioned, the
velocity field is mode independent and given, and (ii) due to the scaling
assumptions 1-3, all the equations (13)-(18) are also independent of the mode
number ' n'.
2. The following scaling properties are exposed in equations (20) and (21): (i)
the time in each n-th Maxwell mode is scaled by the relaxation ratio an (= 0n/0)
as: ~n= t/an; (ii) the velocity in each n-th mode is scaled by the ratio an as:
_fin(1;n,X)--anV(t,x); and (iii) the non-dimensional n-th modal extra stress
tensor is scaled by the modulus ratio, [~n (=Gn/G).
3. It should be noted that the velocity field, v(t,x) in nonlinear multi-mode CEs
cannot be found from the known solution of the problem (13)-(19) for a single
583

Maxwell mode. This can be easily proved for steady simple shearing or the
simple elongation case.
Proof of the scaling theorem.
Using the above features, it is easy to show that the functions defined in
formulae (21) through the single mode solution of the initial boundary problem
(13)-(19), satisfy equations (1)-(6) and initial and boundary conditions (7) for
the multi-mode approach.
Firstly, the evolution equation (1) for n=k immediately follows from equations
(13) and (14) after substituting there z = t/ak, and _fi(t,x) and ~(z,x) from
equation (21). Because value 'k' was taken arbitrarily, equation (1) follows
from equations (13) and (14) for arbitrary 'n'.
Secondly, introducing into equation (16) the expressions: x = t/ak, __fi(t,x) and
15(~,x) from equation (21), as well as the formulae for 'a' and 'b' from equation
(20), yields"

1 gP
--[Otv+(v'g)v]-
- Z~n ~ +g - [o(r)s(c k)]-t- ~-'f~ k (22)
~Y'~a~ ~ n
n n n

This equation holds for any (and all) values of 'k'. Then the momentum balance
equation (4) can be retrieved for the multi-mode approach by multiplying
equation (22) by 13kand summing over all 'k'.
Thirdly, introducing into equation (18) the respective variables from (20) and
(21) yields:

CvakP (23)
~[0tT + (v-V)T] = RV. [K(T,c k ). VT] + akG(T)J 9tr[s(c k ).e]
n

Multiplying equation (23) by 13k /ak taking the sum over all 'k', and using
equation (10), yields the expression (6) for heat equation in the multi-mode
approach.
To finalize the proof, the multi-modal version of the continuity equation (5), as
well as the initial and boundary conditions (7) must be retrieved. In particular,
equations (5) and (7) follow directly from equations (17) and (19) when
employing equations (20) and (21).
Since in the reality neither flow, nor stress fields are known, the requirement of
the theorem about the known velocity field seems at first sight quite artificial.
584

Yet, the situation with known flow field commonly happens when using any
iterative numerical scheme. This will be clarified in the following Section.

4. EXAMPLES

In this Section, we illustrate the scaling approach and a possible computational


economy it gives for calculations of complex flows.

4.1. Start-up simple shear flows.


As an easy example, we consider first an non-inertial and isothermal solution
of start up problem for simple shearing. In this case, the loading of polymer
liquid begins at time instant t = 0 initiating from the rest state, when a constant
shear rate ~, is suddenly imposed. In typical computations, the value of ~ is
confined in an interval [3; min, "Y max] where ~, min < < "Y max. Here ~ min usually
belongs to the region of linear viscoelasticity, and ~, max is attributed to the
highest available experimental data. We now make a choice of values { ~ k} as:
~t o -- ~ min < "~ 1 < ~ 2 < . . -< ~t n = ~t max. For any ~, from the interval, the solution
of the problem for extra stress tensor in an 'i'-th mode is represented due to the
scaling theorem as:

erie(t, ~ ) - Gis(t/Oi, ~' Oi). (24)

Here s is the non-dimensional mode-independent tensor function of two


variables, and 0i the relaxation time. In multi-mode approach, we assume the
ordering: 01 > 02 > ... > ON where N is the number of modes.
It is wise to start making calculations with the mode with highest relaxation
time 0~. Then making 'n' calculations of the initial problem for the single 1-th
mode with the values ~, k ( k = 0, 1, 2,...,n) in the whole time interval t > 0, we
can obtain 'n' functions"

~le( t, ]t k) = Gls(t/01, ~' k01) (k = 0, 1, 2 , . . . , n) (25)

To calculate the extra stress tensor in any, say j-th mode, we do not need to
solve once again the initial problem, but simply use the scaling approach:

_%~(t, ~, ) = (Gj/G~)~]~(t0~/0j, ~ 0j/0~) - Gjs(t/aj01, aj ~ 01) (26)


( j = 1, 2, ..., N).
585

Here aj = 0j/01. Note that since in this example, aj< 1, we need only to
extrapolate the values of or1e, already computed, for different values of ~ which
are not equal to q?k. Thus, to calculate the extra stress tensor using the scaling
approach we just need to make 'n' calculations for a single (say, first) mode for
all chosen values of shear rates r k and then restore the values of extra stress
tensor by interpolation.
It seems that (especially) for homogeneous simple shearing, there is no
computational advantage, as compared to the conventional procedure. Indeed,
using scaling one needs to compute in this case, n+l coupled ordinary
differential equations for the first mode and also perform a time consuming
interpolation procedure. If Ks is the amount of operations required for solving
the start up problem for a single model with a single value of ~, k and Kj is the
amount of the interpolation operations, the total amount of operations, K, using
scaling is: Ks~ = (n~c+l)Ks + Ki. On the other hand, the conventional method of
computations is performed with the following amount of operations: Kcon
NKsn~o.. Here n~c and ncon are the numbers of shear rates ~ k we need to perform
the complete set of computations. In order to compare the numbers of
operations when using these methods, we note that the values of n~c and n~on
should be generally quite different from each other. The value noo, can be taken
arbitrarily and related usually to the compared data. To make the interpolation
valid, the value of ns~ however, should be twice-three time more than the
amount of modes, N, i.e. ns~ = (2-3)N. Thus the efficiency of scaling over
conventional method is estimated by the ratio:

KsJK~on ~ [(2-3)Ks + Ki/N]/(Ksn~on). (27)

Since approximately Ki ~ N, there is no expected computational economy when


using scaling in simple shearing. The same will happen for other simple 1D
flows. The reason for this is that the computations of ordinary differential
equations are very cheap today.

4.2. Scaling in computations of complex flows


For the simplicity, we consider here only isothermal and steady, but generally
3D complex flows, with non-slip boundary conditions. A steady contraction
flow in transition region between two non-symmetrically connected tubes of
geometrically arbitrary cross sections serves as a good example of such flows.
The level of velocity field v(x) is characterized here by the value of flow rate Q
586

commonly defined as integral of v(x)over any cross-sectional area. The values


of flow rate Q are convenient to mark the values of Deborah(Weissenberg)
numbers related to any real flow field. This is because the scalar linear
functional Q{v(x)} is a single valued, however in general, non-monotonous. We
will use below the steady viscoelastic CEs (1)-(3) under assumptions (8) and
(9), along with the momentum balance equation (4) (ignoring inertia and gravity
force) and continuity equation (5). In order not to miss some possible flow
bifurcations, the flow field should be gradually increased from a lowest to a
highest possible values. These values are correspondingly characterized by the
values of flow rate, Qmin and Qmax. Here the value Qrmn commonly corresponds
to the linear viscous flow regime. The value of Qmax corresponds to expected
high Deborah (Weissenberg) number, and is usually limited by the occurrence
of some numerical instabilities. Thus within the interval [Qmin,Qmax] we can
generate the set {Qi} (Qi < Qi+l, i = 0, 1, 2,..., I), so that Qo=Qmi, and Qi=Qmax. In
interesting cases when there are some regions of non-monotony in the
functional dependence Q{v(_x)} a special treatment should be taken. This is
outside the scope of our simple example.
To solve the problem, for any value Qi from the interval, we employ the
iterative algorithm [16] which has been successfully used for solving steady
contraction flows in capillaries. Applying this algorithm to the above set of
equations yields"

( V im-1 ._V)=Cm,i - c=n,i


m .co m-1 +co m-1 9Cnmi- r
---i ---i = , = , .e.
=l
m-1 +e m-1 .c m ) +
=i =n,i
(28)
m
~(r i ) / O n -- _0' (n = 1 , 2 , . . . , N )

N
o'im _ --pim ~ + e , m l + qa (e m
i - eim-1 ), 13"ie,m - Z Gn-S ( ~ ' c-n,i
m ); (29)
n=l

m e,m ~7 2 (V m m-1
~-Pi -- g ' o . + l"la i - vi ), V ' V im -- 0. (m > 0) (30)

Here equation (28) describes the evolution equation for n-th mode, equation
(29) is the formulation of the stress tensor in the N-mode model, with the
artificial Newtonian term (with 'viscosity' rla) included, and equation (30)
describes the momentum balance and continuity equations. In equations (28)-
(30), the subscript 'i' indicates the flow with the flow rate Qi and the superscript
'm', the number of iteration.
587

The above iterative scheme works as follows. For a given 'i', the "initial
oo oo
guess" for the virtual velocity field is obtained as" v ~ Xi vi_l,. Here vi_ 1, is
the solution of the problem with the flow rate Qi-1 and )v~ is a numerical
parameter ()vi > 1). Since the vector field v ~ is still solenoidal, there is the
relation: Qi = )vi Qi-1. Thus to guarantee that given flow rate values Qi-1, and Qi
should be close enough, one should take the value of )vi just slightly above 1.
Using the value _,v~, the strain rate and vorticity tensors for the initial guess, =,e.~
and mi~ are easily calculated. Then equations (28) should be solved with some
proper (usually given upstream) conditions to find the space distribution of
values for N six-component tensors c 1 ., and concomitantly, the extra stress
=n,l

tensor, ~i e'~. The last step in the first iteration is the determining of the first
iteration for the velocity field, vl, and the pressure p l at the first iteration.
These are found as the solution of the linear non-homogeneous Stokes problem
(30) under constraint Qi = const. Then one should repeat the iterative procedure
till it converges, if it possible. Obviously, for any value 'i', the artificial viscous
terms in equations (29) and (30) vanish at the convergence, m --+ oo. The
described iterative scheme works well [ 16], if the CEs have some good stability
properties. It allows to find effectively the velocity and stress distributions for
very high level of Deborah numbers (flow rates Qmax). It is evident that the
scaling approach can be applied to the computations at any step of iterative
procedure, since the velocity field here has been found from the calculations on
the previous step.
To solve equations (28)-(30), some numerical schemes employing various
discretization (e.g. pure FEM, or combined FEM and upstream) methods can be
applied. These are not the subject of discussion in this paper. Obviously in this
case, the highest computation burden is the numerical solution of the evolution
equations (28) for the N tensors =n,i
c m on any iterative step ' m ' with the given
flow rate Qi. This is because for any given values 'n', 'm' and 'i', equations (28)
is generally the set of six coupled nonlinear partial differential equations. Once
again, it is wise to start solving equations (28) with the highest relaxation mode,
n=l. Then using the scaling approach, one can find the distributed values of
tensors Cn,
im and extra stress tensor =13ie'm for any level of flow rate Qi and the
iterative step 'm', as follows"
588

m m
__Cn,i - Cl,i
{a n -vm-1 } (an
-- 0
n
/ 0 9 n-
1,
1,2,
"",
N)

N
r _ Z ans(r mi )
Cn, (31)
n=l

Here we need to compute equation (28) only once for n=l and then using an
cl,i {anvm-1}
interpolation procedure, obtain the values of the functionals =m - from

a search of already computed values. This is because in equations (31), an < an-1
(On < 0n-l). Since the flow rate Q is a linear functional of velocity v, the
respective values of flow rates corresponding to the velocity a~y_iv-can be found
as Qni = anQi. Since an<l, SO Qni<Qi. Thus the value of Qni belongs to an already
computed domain, [Qj,Qj+I] (j+l<i). Evidently, the described procedure
preserves the nonlinear relationship between the stress and flow fields.
We now estimate the efficiency of scaling type of computations as compared
m. be the amount of numerical computations
to the conventional one. Let s Kn,i,
for solving equation (28) for a given 'n'-th mode, at the given value of flow
rate, Qi, and for a given iteration, 'm'. To simplify the estimates, we assume that
these values are independent of 'n', 'i' and 'm', i.e. s Knmi ~ Ks. Let Kret be the
m
amount of operations for retrieving the distributions Cn,i (x). Finally, let Kvp be
the amount of numerical operations for solution of the Stokes problem (30).
Then neglecting the amount of operations needed for tabulations of extra stress
tensor, the total amount of operations for computing the problem using scaling
is:

Ksc ~ [Ks + Kret'(N-1)] KvpM'I. (32)

Here M is the amount of 'm' iterations within a computational precision, I is the


total number of intervals where Qi= const, and N is the amount of relaxation
modes (typically, N~3-10). We now can employ in equation (32) the inequality"

Kret'(N-1) < < Ks, (33)

which holds for complex flows. Inequality (33) simply displays the fact that in
complicated computations, the number of operations for retrieval of already
589

calculated data is considerable less than that for straightforward computations


m
of the distributions Cn,i (x).
On the other hand, using the conventional calculations, the amount of
computations, K~on, can be approximately found as"

Kcon~ Ks.N.KvpM.I. (34)

Then comparing formulae (32) and (34) in the cases when inequality(33) holds,
one can find that

Ks/'Kcon ~, 1 / N . (35)

Equation (35) shows that for calculations of complex flows with multi-mode
viscoelastic CEs, the scaling approach can provide with a significant
computational economy as compared to the straightforward computations. The
key element here is the inequality (33) which holds only for complex flows.
Finally, it can be demonstrated that almost the same computational economy
can be achieved with the scaling approach for solving non-steady problems for
complex viscoelastic flows. Here, the scaling approach can give more economy
when using more stable implicit methods.

5. CONCLUSIONS

1. We have demonstrated that employing the similarity assumptions 1-3 in


general formulations of Maxwell type multi-mode CEs (1)-(3) allows for
scaling velocity and time in each Maxwell mode, by the mode's relaxation time,
and for scaling stress by the mode's modulus.
2. We proved that with known velocity field, these scaling properties reduce the
solution of a general non-isothermal initial boundary problem for multi-mode
Maxwell type CEs to solving a similar problem for a single Maxwell CEs.
3. We demonstrated a computational advantage of scaling approach for solving
problems of complex viscoelastic flows. Though today's computational tools do
not allow to obtain the above single mode solution in all its generality, there is
no doubts that this will be possible in near future.
4. As shown, the scaling approach might have its primary advantage as
problem's complexity grows, i.e. when dealing with 2D/3D complex geometries,
multi-mode models and unsteady problems.
590

5. We have proven the scaling theorem for very broad formulations of multi-
mode CEs of Maxwell type. However, the very existence and stability of
solutions of the CEs will highly depend on such fundamental properties as the
thermodynamic consistency and stability of CEs.

ACKNOWLEDGMENTS

This work was supported in part by the NSF Grant (No CTS-932-0037). The
authors would also like to thank Sunil Acharia for valuable discussions and Jay
Jeong for help.

REFERENCES

1. R.G. Larson, Constitutive Equations for Polymer Melts and Solutions,


Butterworth, Boston, 1968.
2. A.I. Leonov and A.N. Prokunin, Nonlinear Phenomena in Flows of
Viscoelastic Polymer Fluids, Chapman & Hall, New York, 1994.
3. M. Grmela, Phys. Lett., A 111 (1985) 36.
4. M. Grmela, Physica, D 21 (1986) 179.
5. M. Grmela and P.J. Carreau, J. Non-Newton. Fluid Mech., 23 (1987) 271.
6. A.N. Beris and B.J. Edwards, J. Rheol., 34 (1990) 55.
7. A.N. Beris and B.J. Edwards, J. Rheol., 34 (1990) 503.
8. A.I. Leonov, J. Non-Newtonian Fluid Mech., 42 (1992) 323.
9. R.J.J. Jongschaap, K.H. de Haas and C.A.J. Damen, J. Rheol., 38 (1994) 769.
10. Y. Kwon and A.I. Leonov, J. Non-Newton. Fluid Mech., 58 (1995) 259.
11. M. Simhambhatla and A.I. Leonov, Rheol. Acta, 34 (1995) 259.
12. H. Tautz, Experim. Techn. Phys., 7 (1959)1.
13. K.H. Hellwege, I. Henning, and W. Knappe, Kolloid-Z. Polym., 188
(1963)121.
14. L.N. Novichenok and Z.P. Shulman, Thermophysical Properties of
Polymers (Russian), Nauka i Tekhnika, Minsk, 1971.
15. A.I. Leonov, Rheol. Acta, 15 (1976) 85.
16. R.K. Upadya and A.I. Isayev, Rheol. Acta, 25 (1986) 80
591

A T H E O R Y OF F L O W IN S M E C T I C LI Q U I D CRYSTALS
F. M . L e s l i e
Mathematics Department, University of Strathclyde, Livingstone Tower,
Richmond Street, Glasgow G1 1XH, Scotland
1. I N T R O D U C T I O N
The continuum theory for nematic liquid crystals proposed by Er-
icksen [1] and Leslie [2] has significantly improved our understanding of
flow phenomena in these anisotropic liquids. Initially it explained flow
alignment and non-Newtonian flow effects exhibited by these liquids, and
thereafter contributed greatly to our insight into a variety of instabili-
ties that can occur in these fascinating fluids under the application of
magnetic and electric fields, thermal gradients, and flow [3]. Also the
realisation that changes in alignment of the local anisotropic axis can
induce flow, which in turn influences the changing alignment, had seri-
ous implications for applications using these materials in display devices
[4-6], this leading to a greater appreciation of the theory.
More recently potential for fast switching display devices has led to
an increasing interest in smectic liquid crystals, and particularly smec-
tic C liquid crystals. These liquid crystals display translational order as
well as orientational order, and clearly their intrinsic layering presents
new problems in terms of the formulation of an appropriate mathemat-
ical model. Motivated largely by applications in display devices, Leslie,
Stewart and Nakagawa [7] have recently proposed a theory for smectic
liquid crystals, which may prove useful for the modelling of such appli-
cations. Essentially it invokes two simplifications in order to reduce the
mathematical complexity, assuming that the layer spacing in these ma-
terials remains constant, and also that the tilt of the alignment of the
anisotropic axis with respect to the layer normal also remains fixed. For
many situations these assumptions appear reasonable, but clearly there
are others where they are too restrictive. The assumption of fixed layer
spacing largely rests on the notion that variations in layer thickness re-
quire considerable energy, and consequently it seems reasonable when the
592

smectic is not under undue strain. Also the constant tilt presumably ex-
cludes certain thermal and pre-transitional effects. However, the theory
has already proved useful in a practical context [8,9], and in this article
we attempt to summarise its predictions relevant to rheology.
For readers not familiar with general properties of liquid crystals,
the books by de Gennes and Prost [10] and Chandrasekhar [11] give full
accounts of physical properties, while that by Collings [12] provides a
fairly gentle introduction.
2. S M E C T I C THEORY
This section presents a summary of the equations proposed by Leslie,
Stewart and Nakagawa [7] for smectic C liquid crystals. The theory is
somewhat restricted in that it assumes that the smectic layer spacing re-
mains constant, although the layers may bend and deform, and also that
the tilt of alignment with respect to the layer normal remains unchanged.
However, the theory differs from its predecessors in that it does allow for
finite bend and deformation of the layers, and for non-linear displace-
ments and flow. Like the theory for nematics it rests on the concepts of
classical mechanics.
With the above assumptions one can readily describe the layered
configurations by employing two orthogonal unit vectors, one the unit
layer normal a, and, following de Gennes [10], the second unit vector c
is perpendicular to the layer normal indicating the direction of tilt of the
alignment. Clearly this second vector is tangential to the smectic layer.
Hence the two vectors or directors are subject to
a.a-c.c--1 , a.c-0. (1)
Also, in the absence of any singularities or defects in the layering, as both
Oseen [13] and de Gennes [10] argue one must also impose the constraint
curl a = 0. (2)
A more familiar constraint arises from the customary assumption of in-
compressibility, and so the velocity vector v is subject to
d i v v - O, (3)

and of course the density p is constant.


The theory essentially rests on the balance laws of classical continuum
mechanics, the balances of linear and angular momentum. In Cartesian
tensor notation the balance of linear momentum is given by
593

pvi -- pFi "Jr-tij,j, (4)


F denoting the body force per unit mass, t the stress tensor, and the
superposed dot the material time derivative. However, the balance of
angular momentum takes the less familiar form
pK~ + eijktkj + l~j,j = 0 (5)
including terms generally omitted, K the external body moment per unit
mass and l the couple stress tensor, but an inertial term associated with
intrinsic local angular momentum is omitted since it is generally consid-
ered negligible. Since thermal effects are ignored in the problems dis-
cussed, it is not necessary to give an equation representing a balance of
energy. Here we employ Cartesian tensor notation whenever necessary,
and therefore a repeated index is subject to the summation convention,
a comma preceding an index represents a partial derivative with respect
to the corresponding spatial coordinate, and 5ij and eijk denote the Kro-
necker delta and alternator respectively.
The static version of the theory rests on the assumption of a local
stored energy function that depends upon the two directors and their
gradients, being quadratic in the latter. This energy takes the form
2w - + + 2K a . cjaj. ck

-1- 2K~eipqCpaqai,kCk "~ K~(ci,i) 2 -t- K~ci,jci,j -t- K ~ c i , j c j c i , k c k


(6)
-t- 2K~ci,jcjci,kak + 2K~eipqCpaqCi,~ak + 2K~Cci,icjaj,kck
+ 2K~Cci,iaj,j,
the coefficients constants. The static equations can now be obtained by
appeal to a virtual work formulation [7,14], and one finds that
OW OW
tij - -pSij Jc flpepjkak,i -- ~ C k , i + tij,
cOak,j ak,i OCk,j
(7)
/
~ij -- ~papSij -- ~ i a j -[- eipq (ap
OW
+ Cp
ow )
\ Oaq,j OCq,j
where the pressure p and the vector fl arise from the constraints (3) and
(2), respectively. The tensor t denotes the dynamic viscous stress, but
the corresponding term in the couple stress is omitted since it can be
shown to be zero by essentially a thermodynamic argument requiring the
594

rate of viscous dissipation always be positive.


The viscous stress t is the sum of a symmetric part
tij -- ~oDij + ~lapaqDpqaiaj + ~2(Dikakaj + Djkakai)
+ ~3CpCqnpqciCj + ~4(DikCkCj + DjkCkCi)
+ ,~~n~(~icj + ~j~i) + ~1 ( A ~ j + Aja~)+ ~(C~cj + Cj~,)
+ )~3cpAp(aicj + a j c i ) + al(Dikakcj + n j k a k c i + n i k c k a j (8)

+ Djkckai) + ~2[2apcqDpqaiaj + apaqDpq(aicj + ajci)]

+ ~[2~.c~D.~cicj + c.c~D.~(aicj + ajci)] + ~l(C.aj + C j ~ )


+ r2(Aicj + Ajci) + 2r3cpApaiaj + 2T4cpApcicj,
and a skew-symmetric part
ti~ - )~l(Djkakai - Dikakaj) + )~2(Djkckci - Dikckcj)
+ ~ c , ~ D . ~ ( ~ i ~ j - ~jci) + ~ ( A j a ~ - Ai~j) + ~ ( C j c ~ - C~cj)
+ )~6cpAp(aicj - a j c i ) + V l ( D j k a k C i - DikakCj) (9)
+ ~ ( D j k c ~ , - Di~c~aj) + ~a.a~D,~(~,c~ - ~jc.)
+ ~c.c~D.~(~icj - ~jc,) + ~(Ajci - Aicj + Cj~i - C,~j).
where
2Dij - vi,j + vj,i , Ai - ai - Wijaj , Ci - ci - Wijcj, (10)
2Wij -- vi,j - vj,i
and the coefficients are constants.
The above theory is that for chiral smectic C liquid crystals, or what
are also called ferroelectric smectic liquid crystals, being invariant to only
proper orthogonal transformations of the coordinate axes. Put differently
such materials can distinguish between right-handed and left-handedness.
However, should one restrict the theory to non-chiral materials, those
invariant to transformations including the full orthogonal group, one must
choose
g~ = g~ =0. (11)
As Carlsson, Stewart and Leslie [15] discuss, there are grounds for set-
ting the former coefficient zero even in chiral theory, and therefore the
distinction between theory for chiral and non-chiral materials rests es-
595

sentially on one elastic term that gives rise to twisted configurations in


which the c director rotates about the layer normal, much as the direc-
tor in cholesterics or chiral nematics rotates about a given axis to form
helical configurations. It is of interest to note that the viscous stress is
identical for both chiral and non-chiral smectics.
An inspection shows that the above equations are invariant to the
simultaneous change of sign in the two directors a and c, this being con-
sidered the symmetry appropriate to smectic C liquid crystals. However,
if the material symmetry allows the independent change of sign of the
two directors, then the constitutive equations simplify, the elastic coef-
ficients K~, g ~ ~ and g ~ ~ in equation (6) and also the ~ and r terms in
the equations (8) and (9) having to be set equal to zero.
It is possible to re-arrange the above equations in forms more conve-
nient for calculations that follow. To this end the intrinsic body moment
in equation (5) can readily be expressed as
"iss -- eijk(ajg~ak + cjgk)
eijktkj ~c , (12)

wherein

g~ = --2()~lDikak + )~3CiCpaqDpq -F )~4Ai -1- A6CiCpAp -]- r2Dikck


"[- T3CiapaqDpq + T4CiCpCqDpq --~ T5Ci) ,
~tc = --2()~2DikCk 4- )~5Ci -F T1Dikak + 7"5Ai).

Similarly in cases of interest the external body moment can be written


p g i - eijk(ajG~ + cjG~). (14)
As a consequence some manipulation allows one to rewrite the balance
of angular moment (5) as two equations

,j coal -]- Ga -F {ta -}- 7ai + tzci -]- eijkflk,j -- 0


(ow) ,j
OW
+ + Z + + - o,
(15)

where the scalars V, # and r are arbitrary multipliers stemming from the
constraints (1). Also, with the aid of equations (115) the balance of linear
moment (4) can be expressed as
pi~i = pFi + a~,ak,i + G~ck,i - ~,i + (~ak,i + ~ c k , i + tij,j, (16)
596

where
#=p+W. (17)
The forms (15) and (16) are more convenient for our purposes in the
following sections. Also, employing the forms (12) allows one express the
viscous dissipation inequality as
t i8j D i j __
g a A i - gi
*~C C
i >_ O, (18)
which leads to some restrictions upon the viscous coefficients.
To conclude this section we attempt to clarify the physical role of
the vector/3 arising from the constraint (2). If we consider planar layers
subject to flow or external fields, equations (7) show that they are subject
to moments

~.i --- ~.ijaj -" ~ p a p a i -- ~i q- eipq ap Oaq,j -~- Cp-OCq,j" aj. (19)

Thus, if the layer normal coincides with the z-axis, one has
= + ..., 4 - + ..., - 0 + ... (20)

simply giving the/3 contributions explicitly. Hence, if the layers are con-
strained to be planar by for example confining parallel plates, this vector
provides a means by which the couple stress transmits the necessary
torques to maintain the parallel layering.
3. S I M P L E S H E A R F L O W
Given the layered structure always present in smectic liquid crystals
it is not clear as to how this class of liquid crystal will respond to flow in
general, there being little by way of clear experimental evidence to guide
us. Here, therefore, we choose to consider the simplest options, two cases
in which the layered structure appears to be compatible with the imposed
flow. In our first example our results appear to be consistent with our
premise, but in our second the question does remain somewhat open.
3.1 P l a n a r layers p a r a l l e l to t h e b o u n d i n g p l a t e s
Consider a smectic C liquid crystal confined between two parallel
plates with the smectic layers everywhere parallel to the plates, one of
which is at rest but the other moves with velocity V along a straight line
in its own plane. With Cartesian axes such that the z-axis is normal to
the plates and the x-axis parallel to the imposed motion, it is natural to
examine solutions of the equations of the previous section in which
597

a-(0,0,1) , c- (cosr162 , v- (u(z),v(z),O), (21)


this choice clearly consistent with the constraints (1), (2) and (3). As
Gill and Leslie [16] discuss, with this choice equations (16) quickly yield
[~z = (r/~ + r/2 cos 2 r + ~/2 sin r C v ' - c~, (22)
tyz - (r/1 + r/2 sin 2 r + r/2 sin r cos Cu' - c2,
where the prime denotes differentiation with respect to z, Cl and c2 are
constants, and the viscosities r/1 and r/o are given by
2~71 - / z o + ~2 - 2A1 + A4 , 27/2 - ].t4 -[- ~5 -[- 2A2 - 2A3 -4- A5 -4- A6, (23)
this assuming that the flow stems solely from the relative motion of the
plates, there being no imposed pressure gradients, or external body forces
or moments, apart from gravity which is readily absorbed into the pres-
sure p. Consideration of special cases of the above equations quickly leads
one to the conclusion that T]I and 771-~-772are both positive. The equations
for angular momentum (15) lead only to
K~r + (T 1 -- T5)(U' sin r v' cos C) - 0, (24)
this assuming a choice of the vector fl of the form
-- (]31(Z) , /~2(Z) , 0), (25)
being more than reasonable for the problem under consideration.
Before proceeding further, it is of interest to look for solutions repre-
senting uniform flow alignment of the c-director in shear, and clearly we
have two such options
r or 7r , v ' - 0 . (26)
Also, consideration of small perturbations leads one to select
r if (1"5 --7"l)Ut > 0, (/)--71" if (7"1 -- 7"5)U' > 0, (27)
essentially rejecting the alternatives on grounds of stability. Thus one can
have uniform flow alignment tilted forward in the direction of shear or
backwards, depending upon whether 75 is greater or less than I"1 [16,17].
The former case is consistent with one's intuition, and influences the
material symmetry chosen, namely that our equations should be invariant
to the simultaneous change of sign of the two directors. If one were to
opt for invariance to the change of sign of a or e independently, no flow
alignment would be predicted for this the simplest of shear flows involving
smectic C liquid crystals.
598

Returning to the general formulation one quickly obtains from equa-


tions (22)
~I(T]I -~- ~2)U t - - Cl(~l -~- r/2 sin 2 r c2r/2 sin r 1 6 2 (28)
,11(,/1 + r/2)v' -- c2(r/1 + y2cos 2 r clr/2 sinr r
giving the velocity gradients as functions of the alignment. Also elimina-
tion of the velocity gradients from equation (24) leads to
g~r + (rl - T5)(Cl sine- C2 COSr -- 0, (29)
which readily integrates to give
K~r '2 + 2(T5 -- rl)(Cl cosr + c2 sin r -- c3, (30)
where c3 is a constant.
The boundary conditions to be satisfied in general are
u(d) - V, u(-d) - v(d) - v(-d) - 0, r - r , r - r (31)
the gapwidth being 2d with the origin chosen midway between the plates,
and r and r two prescribed angles. It readily follows from equations
(28) that

771(?']1 "~- ?~2)V -- c 1 f__ (Z]l -~- ?']2


J- d
si"2r - c2 r12 F sin r cos
d
Cdz ,
(32)
0 C2 f__ (771 -~- 7]2 r Cl ?']2 sin r Cdz ,
J- d d
providing two equations for the constants Cl and c2. If the choice of
angles r and r is such that
r - -r (33)
and the angle r is an odd function, then the above implies that c2 is zero,
and

771(~]1 -~- ~72)V -- 2Cl (Z]l -~ Z]2 sin 2 r (34)

so that Cl is clearly positive.


While we know rather little concerning likely values for the vari-
ous material parameters, they appear in the above equations simply as
three combinations, the two viscosities rh and 7/2 and the group (T5 --
rl)/K~r/1. Consequently a numerical integration of the equations is rel-
atively straightforward. Gill and Leslie [16] give details for two cases,
599

with the angle r either symmetric or asymmetric. For the former the
angle r tends to either zero or r in the centre of the cell, dependent upon
whether T5 is greater or less than T~. For asymmetric solutions the results
are more interesting with the twist in the centre of the cell unwinding as
the shear rate increases, the flow causing the twist to concentrate near
the plates with a growing region of uniform alignment in the middle of the
cell. This result appears to be consistent with the early experiment by
Pieranski, Guyon and Keller [18]. Note that for this problem the equa-
tions are equally valid for chiral or non-chiral materials, the difference
between the two, if any, being in the initial equilibrium configuration.
3.2 P l a n a r layers p e r p e n d i c u l a r to t h e b o u n d i n g p l a t e s
Our second example considers what is commonly called the bookshelf
geometry with the layers planar but normal to the plates, and a shear
imposed parallel to the layers would appear to be compatible with the
smectic layering, as Carlsson, Leslie and Clark [17] discuss. In this case we
examine solutions referred to Cartesian axes with the z-axis perpendicular
to the plates of the form
a-(0,1,0) , c- (cosr ,sine(z)), v- (u(z), v(z), 0), (35)
this again satisfying the constraints (1), (2) and (3). Here, however,
we must confine attention to non-chiral materials, since chiral materials
would require the inclusion of a y dependence in the angle r and pre-
sumably also in the flow components, which would lead to an analysis
beyond the scope of an initial investigation of the simpler options.
As Gill and Leslie [16] discuss, the equations of linear momentum
(16) with the above choice yield
t'~z - #1 (r + #2(r cosr cl, (36)
t' z = + osr
where again Cl and c2 are constants, and there are no imposed pressure
gradients or fields. The viscosity functions in the above equations are
given by
2#1(r = ~Uo+ ~ 4 ~- )k5 "4- 2A2 cos 2r + 2#3 sin 2 r cos 2 r
2~t2(r -- tel -4- I"1 -4- "r2 -4- T5 -4- 2(n3 -4- 7"4)sin 2 r (37)
2~t3(r --/Zo 4-/z2 A- 2A1 + A4 + (~t4 -~- ~5 -- 2A2 + 2A3 d- A5
-I- )~6) sin 2 r
500

Here also one can solve the equations (36) to obtain the velocity gradients
in terms of the angle r and the imposed shear stresses. The equations of
angular momentum (15) reduce to

f ( r 1 6 2 -4 21 dr(C)
de r _ (~5 + ~2 cos 2r - (T1 -~- 7"5)v' c o s r 0, (38)
where
f(r - g ~ + g~ cos 2 r + g~ sin 2 r (39)
but this entails a choice of the vector/3 and the scalar 7 of the form
t3-(~(z)y , 0 , ~(z)y) , 7=~(z)-y~(z), (40)
the prime again denoting differentiation with respect to z.
There are two respects in which this solution differs from the former,
both giving rise to grounds for not pursuing it in any great detail at
least for the present. One is that the material parameters appear in
greater numbers, and our lack of knowledge of their relative magnitudes
therefore inhibits progress. The second is that the surprising dependence
of the vector/3 on the y coordinate means that the resulting torques
can become large, and therefore the material may seek to relax to a less
strained state, say by forming domain structures. Some guidance from
related experimental studies would not go amiss in this respect.
To close this section we consider the alignment dictated by shear flow
in this geometry. Here the options are
r162 rico , v'-0, (41)
where r is the acute angle defined by
cos2r - -~5/~2 , provided ~5 <_ [~21, (42)
otherwise uniform alignment is not possible in shear flow in this geometry.
Note that the viscosity )~5 is clearly positive as a consequence of the
inequality (18), as Carlsson, Leslie and Clark [17] discuss. Of the solutions
(41) we select on grounds of stability
r162162 , if)~2u'<0 , r162162 , if)~2u'>0,(43)
as Gill and Leslie [16] and Carlsson, Leslie and Clark [17] show. For
such uniform alignment without transverse flow it is of interest to note
from equations (36) and (37) that the transverse shear stress is in general
non-zero, this an aspect perhaps more readily amenable to experimental
observation.
601

4. F L O W I N S T A B I L I T I E S
Continuum theory for nematic liquid crystals quickly gained credibil-
ity through its success in describing satisfactorily a number of instabilities
that occur in these materials, when subject to external influences includ-
ing flow. Here our aim is to show that similar flow induced instabilities
are likely in smectic liquid crystals, being predicted by the present theory,
and naturally one hopes to stimulate relevant experimental studies. In
this section it suffices to give a simple illustration discussed by Leslie and
Blake [19].
We return to the geometry of our first example in the previous section
and thus consider smectic C layers confined between parallel plates, the
layers parallel to the plates, and subjected to simple shear flow. Here,
therefore, with the same choice of Cartesian axes, the z-axis normal to
the plates and the x-axis parallel to the imposed flow, we again examine
solutions of the form (21), and find as before that the governing equations
are
g~r + (T1 -- rs)(u' sin r v' cos C) - 0,
(r/~ + y2 cos 2 r + 7/2v' sin r cos r c~, (44)
(711 -~- 712 sin 2 r + ~/2u' sin r cos r - c2,
where v/1 and v/2 are again given by equations (23). Choosing the origin
midway between the plates, a distance 2d apart, the relevant boundary
conditions are
u(d) - Y , u(-d) = v(-4-d) - O , r162 (45)
where V is the velocity of the moving plate, and r is either zero or r.
Straightforwardly the equations (44) have simple solutions subject to
conditions (45) in which
r162 , u- Y(z + d)/2d , v-O, (46)
t~z - (~71 + r12)l//2d , tyz - O.
As remarked above the solution anticipated is
r 0 , if r~ > T~ , r , if 7"5 < T1, (47)
being unstable otherwise. Consider small perturbations to the solution
(46) in which
r r + r , u - Y ( z + d ) / 2 d + ~t(z) , v - ~(z), (48)
602

and one obtains the following equations for these perturbations


g~$" + (~ - ~)~os r ~') - O,
(49)
(~1 + ~ ) ~ ' - a , ~1~' + ,7~V;~/2a - b,
a and b denoting perturbations to the shear stresses. The boundary
conditions for the perturbations are
~(~d) = ~(+d) = r - O. (50)
In view of the above the second of equations (49) at once implies that ~2
is zero, and elimination of ~3 between the remaining equations yields
~,, + V(7/a + rl2)(rl - r5)cos ~bo [ ~ _ 2bd/V(rll + r/2)] - 0. (51)
2K~rl~ d
If the product (rl - T5)cos Co is positive, the above equation has a
solution subject to conditions (50) of the form
2bd cos wz Y (7/1 + r/2 )(T1 - T5) COSr
- - V(~l + ~ 2 ) ( 1 - coswd ) ' w 2 = 2K~rlld , (52)

and the boundary conditions for ~ require that the last of equations (49)
yield

712V / _ ~ Cdz - 2bd, (53)


-~J d
which in turn leads to
r/2 tan wd + rllWd - O, (54)
and which straightforwardly yields positive values for w.
However, if the product (rl - T 5 ) c o s r is negative, one finds that the
relevant solution is
2bd cosh w z Y (r/1 + r/2)(r5 - T1) cos r
Y(yl -~- Y2)(1 - coshwd ) ' w2 -
- - 2K~rlld , (55)
and equation (53) now yields
r/2 tanh wd + rll wd - O. (56)
As remarked earlier r/1 + 7/2 is necessarily positive, and consequently the
above equation has no non-trivial solutions.
As a consequence our analysis predicts an instability with respect to
the particular perturbations considered for the unstable solutions at a
critical velocity V~ given by
603

Vc - 2g~rll~2c/d(rll + r/2)(7-1 - 75)cos r (57)


where ~ is the smallest positive root of
~1 x + v/2 tan x -- 0. (58)
While other more general perturbations may lead to a lower thresh-
old, it is perhaps premature to attempt further more complex analyses
until experimental evidence becomes available, and also our knowledge of
surface anchoring at a smectic-solid interface improves. With this latter
aspect in mind, Blake and Leslie [20] discuss the above problem when a
magnetic field is applied, this allowing consideration of the two limiting
cases of strong anchoring and no anchoring of the c-director.
5. B A C K F L O W EFFECTS
Given that the original motivation for the theory presented in sec-
tion 2 is primarily to model behaviour in smectic devices, it is of more
than passing interest to examine flow induced by switching of alignment
in smectic cells, particularly so in view of experience of such effects in
nematics [4-6]. Not surprisingly perhaps we turn first to the simplest ge-
ometry in which the layers are parallel to the bounding plates, with the
alignment initially uniform due to strong anchoring at the surfaces. Ap-
plication of a magnetic field parallel to the plates, but perpendicular to
the initial alignment rotates the c director around the layer normal, the
distortion symmetric about the plane midway between the plates. Below
we present an analysis of the relaxation of this alignment when the field
is removed.
With a choice of Cartesian axes so that the z-axis is again normal
to the plates with the origin midway between them, and the x-axis now
coincident with the direction of the initial alignment dictated by the
strong anchoring, it is natural to consider solutions of the form
a - (0, 0, 1) , r - (cos r t ) , sin r t), 0 ) , (59)
- t), t), 0),
clearly consistent with the constraints (1), ( 2 ) a n d (3). With this choice
equations (15) simply reduce to

K~~z2 - 2)~5-~ + (7"1 - T5)(sin r cos COz) -- 0, (60)


plus expression for fl, 7, # and T, and equations (16) yield
604

C~U 0 [ Ou Ov
P Ot - 0---~ (rh + 7?2 cos 2 r + r/2 sin r cos r

+ ( ~ - ~-~)si~ r
(61)
Ov 0 [ Ov ~U
P-~ - ~ t(r/~ + r/2 sin 2 r O--~z+ r/2 sin r cos r Oz

+ ( n - ~) r162
and an expression for the pressure p, the viscosities 771 and r/2 again given
by equations (23). From the above equations it is at once evident that one
must include both flow components in order to avoid an over-determinate
system. The boundary and initial conditions are
~(+d) - ~(+d) - r177 - 0, (62)
u ( z , 0) - ~ ( z , 0) - 0 , r 0) - r , Izl < d,
the cell gap again 2d, and r a known even function of z.
Introducing the notation
a - (7"5 - n ) sin r , b - (7"1 - r5) cos r (63)
m-r/l+r/2cos 2r , n - 7 ? l + r / 2 s i n 2r , r-7?2sinCicosr
where r is a constant angle associated with the initial alignment, Leslie
and Blake [19] replace the above equations by
02r 0r Ou Ov
g ~ -~z ~ - 2 ~ ~ -8-i - ~ -8-;z - b -8; - O ,
cgu c92u c92v 02r (64)
P~ - m~z~ + ~ - ~ + a OzO----i '
Ov 02u 02v 02r
p-~ -- r -~z 2 q- n -~z 2 + b cgz cgt ,
whose solution presumably gives at least a good approximation to the
initial relaxation. With the removal of a strong field, it would be ap-
propriate to select r to be r / 2 . By introducing dimensionless variables
given below, one finds that the inertial terms in the last two of equations
(64) are negligible, and therefore it is reasonable to set p equal to zero in
the above. As a consequence the last two equations quickly yield
c92u = ( r b - a n ) 02r (ran - r 2) c92v - (ra - bin) 02r (65)
(mn- ~)-~ ~ OzOt ' ~ o~Ot '
605

and hence with the first


02r
~162 (66)
K~ Oz 3 = (OzOt'
where
a ( r b - a n ) + b ( r a - bin) _ 2)~5 - (75 - T1)2
- 2~ + - . (67)
m n - ?.2 rl1
By appeal to the viscous dissipation inequality one can show that the
above parameter ~ is positive.
Introducing non-dimensional variables as follows
K~ du dv z kt (68)
k-2,~5, ~-/~ , ~--~-, i=~, r-d-- 7,
the first of equations (64) and equations (65) and (66) can be rewritten
02r 0r a 0~ b 0~
~" 0~
/)('2 OT 2A5 0~ 2A5 0~
02~i 02r
(ran - r~) ~ - ~ - ( r b - an)
Or
~ - (~ - bin) ~ 1 6 2 (69)
O(Or'
03r 02r ( , 0<#<1.
O~3 = # O ( O T ' # - - 2 A S
Note that the dimensionless time r in the above is not to be confused with
the r that occurs in equations (15). The above equation.s have solutions
satisfying the boundary conditions (62) of the form
r __
r162 -q~ /" u v
U(r q~ /"
-
v-V(r
__ -- v (70) ~ --q2"r/l~

where
r -- A(cos q~ - cosq) , U(~) - Aq(an - r~-
br)i (sin q~ - ~ sin q),
#(mn J

(71)
v(r - Aq(bm ) (sin q~ - ~ sin q)
~ ( m ~ - - ~at)

A an arbitrary constant and q satisfying


(1 - , ) t ~ n q - q. (72)
Superposition of the solutions corresponding to the infinity of roots q~ of
this last equation leads to
606

oo

q~-- ~ An(cOsqn~ - cosqn)e -q~r/tt


n~l

an ~ bT" oo
q ~ A n (sin q ~ - ~ sin q ~ ) e - q ~ / " , (73)
~t = t , ( m n _ r2 ) n=l

bm - ar oo 2 ~/#
= I , ( m n - r 2) y ~ q,A~(sin q~( - ( sin q n ) e - q " 9
n=l

Finally the initial condition on the alignment (62) requires that


oo

A n ( c o s q ~ ( . - cosq~) - r , (74)
n--1

and noting that the necessary orthogonal function is cos q m ~ , it follows


that
A~ = 2qn /.1
q~ - sin q~ cos q~ ~. r cos q ~ ( d ( . (75)
The above solution resembles rather closely that given by Clark and Leslie
[6] for a nematic, despite the fact that two flow components are induced in
a smectic. Blake and Leslie [21] discuss this solution in greater detail and
show that as in a nematic the induced flow can give rise to a 'kickback'
effect, before the alignment finally relaxes to its initial configuration.
They also briefly consider the more complex bookshelf geometry, but do
not pursue this given the greater number of unknown parameters that
appear in the relevant equations.
Carlsson, Clark and Zou [22] and Zou, Clark and Carlsson [23] discuss
the effects of coupling with flow upon the reorientation of alignment in a
smectic liquid crystal due to the application of an electric field, showing
that flow can play a significant role in such transients. However, their
analyses neglect flow transverse to the smectic layers on the grounds that
it is small. Blake, Leslie and Towler [24] give similar results including this
transverse flow, while Barratt and Duffy [25] provide a linear analysis of
such effects.
6. S M A L L P E R T U R B A T I O N S OF A UNIFORM SMECTIC
First attempts at a theory for smectics were confined to describing
small perturbations of uniformly aligned samples in planar layers. The
Orsay Liquid Crystal Group [26] proposed an elastic energy for smectic C
607

liquid crystals, considering only such small perturbations, and Carlsson,


Stewart and Leslie [27] show that the general, non-linear energy given
in section 2 reduces to exactly the Orsay energy when restricted to such
small disturbances. Also, Martin, Parodi and Persham [28] developed a
corresponding linear, dynamic theory, but by its very nature incapable of
describing most of the effects discussed in earlier sections. However, this
theory presumably prompted Galerne, Martinand, Durand and Veyssie
[29] to study light-scattering by a smectic C liquid crystal, and below we
use their experimental results to obtain estimates for certain combina-
tions of parameters in the theory employed in this chapter.
Consider a smectic C liquid crystal at rest and uniformly aligned
with the directors a and c given by
a--a ~ , c-c ~ (76)

a ~ and c ~ being two constant orthonormal vectors, and examine small


periodic perturbations of the form
a- a~ , c - c ~ + E5 , v-El) , E -- E e x p i ( w t - q . x ) , (77)
where fi, 1~ and ~ are constant vectors, E is a small constant, and w and q
are a constant frequency and wave vector. Since a and c are orthonormal,
then one must have
a~ c~ - a~ + e~ - O, (78)
but the further constraints (2) and (3)lead to
qxfi--O , q.l~-O, (79)
the former implying that
f i - c~q, (80)
a an arbitrary scalar. Consistent with the above one naturally assumes
that
p- E~ , 7- E~/ , p-Eft, r-- E~- , fl- Efl, (81)
where/3, ~,/2, # and f} are constant scalars and a vector.
Leslie and Gill [30] give the general forms of the equations (15) and
(16) consistent with the above small perturbations. Here we wish to dis-
cuss only one special case, which assumes that the layers are undisturbed
so that
= 0. (82)
608

In this event, choosing Cartesian axes so that


a ~ -(0,1,0) , c ~ -(0,0,1), (83)
it follows at once from equations (78) that
- (c, 0, 0). (84)
Also, if we choose
I ~ - (b,O,O) , q - ( O , q s i n O , q c o s O ) , (85)
one finds from equations (16) that 15 is necessarily zero and that
( 2 + ~ + ~)b + 2 ~ c - o, (86)
where
r / - [(#o + #2 - 2A1 + A4)sin e 0 + (#o + #4 - 2A2 + As)cos 2 0
+ 2 ( ~ x - 7-1 - 7-2 + r s ) s i n 0 c o s 0lq 2, (87)
v - [(r5 - T1) sin 0 + (As - A2) c o s O]q.
Also, equations (15) simply yield
(2+~+~ + ~)c + i~b - 0, (88)
where
a - [g~ sin 2 0 + (K~ + g ~ ) cos 2 0 + 2g~ sin 0 cos0]q 2, (89)

plus expressions for/3 and -~, ft and #.


From equations (86) and (88) one readily obtains the equation
2 i p w + *1 2wv,
- o, (90)
iu 2i~5w + a
or equivalently
P)~5w 2 + ()~5q - u 2 + p a ) w + a 7 1 - 0 , w -- 2iw. (91)
In general the elastic coefficients occurring in the energy (6) are small
compared with viscous terms, and therefore one can assume that
p a < < )~5~- u 2. (92)
As a consequence the two roots of equation (91) are to a good approxi-
mation given by
w+ - - ' 7 a / ( ~ 5 ' 7 - v2) , ~'f - - ( ~ 5 ' 7 - u2)/p~5, (93)
609

the 's' and ' f ' subscripts denoting slow and fast modes, respectively, the
former detected by light-scattering experiments. Thus we predict a decay
time rd given by
74 = f?(q, O)/a(q, 0), (94)
where
~ / - 2(X5~/- v2)/rl. (95)
The coefficient ~5 is positive as remarked earlier, but appeal to the viscous
dissipation inequality (18) also shows that r/and f/are both positive.
Galerne and coworkers [29] find that their data for the decay time
observed in their experiments fits an expression analogous to our result
(94), their expression for the elastic factor taking the form
B(O) = (B1 cos 2 0 -b B3 sin 2 0 + 2B13 sin 0 cos O)q2, (96)
the coefficients those occurring in the Orsay energy [26]. However, as
Leslie, Stewart, Carlsson and Nakagawa [31] show, the above expression
is identical to that given by equation (89). Galerne and colleagues find
also that their data is consistent with a viscosity function given by
1 1 cos2(0 -- 0 1 )
= -- + (97)

7g, #a and Vg are viscous constants, and 01 and 02 constant angles. While
the two viscosity functions appear to differ, they are in fact equivalent,
as one finds by setting
~o -[" ~4 -- 2X2 q- )15 -- ttg c o s 2 02 -[- vg q- 7g c ~ 01,

#o + it2 - 2A1 + A4 -- #o sin2 02 + v 9 + 7g sin2 01,


~1 - rl - r2 + r5 -- ttg sin 02 cos 02 + 7g sin 01 cos 01, (98)
2)15 ---- ")'g , V/'2()15 -- )12) = -l-TgCOS01 , ~/r2(T 5 -- 71) -- + ~ ' g S i n 0 1 .
Employing the experimental values for 7g, #o and Vg and the angles 01
and 02 one obtains values for groups of viscous coefficients in the theory
discussed in this chapter.
Given that traditional viscometric measurements may prove inappro-
priate for smectics given their layered structure, it is necessary to devise
suitable experiments employing small samples in which the orientation
of the layers and alignment is well monitored. Gill and Leslie [32] pro-
vide the necessary theory for one further such experiment, involving the
610

reflection and refraction of a shear wave at a smectic-solid interface. How-


ever, there is some way to go in terms of determining the parameters in
the above theory, but this may change with its increasing involvement in
device modelling.

REFERENCES

.
J.L. Ericksen, Trans. Soc. Rheol., 5 (1961), 23.
2. F.M. Leslie, Arch. Rat. Mech. Anal., 28 (1968), 265.
3. F.M. Leslie, Adv. Liq. Cryst., 4 (1979), 1.
4. C.Z. Van Doorn, J. Appl. Phys., 46 (1975), 3738.
5. D.W. Berreman, J. Appl. Phys., 46 (1975), 3746.
6. M.G. Clark and F.M. Leslie, Proc. Roy. Soc. A, 361 (1978), 463.
7. F.M. Leslie, I.W. Stewart and M. Nakagawa, Mol. Cryst. Liq.
Cryst., 198 (1991), 443.
Q
C.V. Brown, P.E. Dunn and J.C. Jones, Europ. J. Appl. Math., 8
(1997) 281.
0
J.C. Jones and C.V. Brown, Proc. 16th Int. Displays Conference
(1996), 151.
10. P.G. de Gennes and J. Prost, The Physics of Liquid Crystals, 2nd
Ed. Oxford University Press, (1993).
11. S. Chandrasekhar, Liquid Crystals, 2nd Ed. Cambridge University
Press, (1992).
12. P.J. Collings, Liquid Crystals: Nature's Delicate Phase of Matter,
Princeton University Press, (1990).
13. C.W. Oseen, Trans. Faraday Soc., 29 (1933), 883.
14. F.M. Leslie, Contemporary Research in the Mechanics and Math-
ematics of Materials, Eds. R.C. Batra and M.F. Beatty, CIMNE,
Barcelona (1996), 226.
15. T. Carlsson, I.W. Stewart and F.M. Leslie, J. Phys. A, 25 (1992),
2371.
16. S.P.A. Gill and F.M. Leslie, Liq. Cryst., 14 (1993), 1905.
17. T. Carlsson, F.M. Leslie and N.A. Clark, Phys. Rev. E, 51 (1995),
4509.
611

18. P. Pieranski, E. Guyon and P. Keller, J. Phys. (Paris), 36 (1975),


1005.
19. F.M. Leslie and G.I. Blake, Mol. Cryst. Liq. Cryst., 262 (1995),
403.
20. G.I. Blake and F.M. Leslie, Meccanica, 31 (1996), 611.
21. G.I. Blake and F.M. Leslie, (pending).
22. T. Carlsson, N.A. Clark and Z. Zou, Liq. Cryst., 15 (1993), 461.
23. Z. Zou, N.A. Clark and T. Carlsson, Phys. Rev. E, 49 (1994), 3021.
24. G.I. Blake, F.M. Leslie and M.J. Towler, Europ. J. Appl. Math., 8
(1997), 263.
25. P.J. Barratt and B.R. Duffy, Liq. Cryst., 21 (1996), 865.
26. Orsay Group on Liquid Crystals, Solid St. Commun., 9 (1971), 653.
27. T. Carlsson, I.W. Stewart and F.M. Leslie, Liq. Cryst., 9 (1991),
661.
28. P.C. Martin, O. Parodi and P.S. Pershan, Phys. Rev. A, 6 (1972),
2401.
29. Y. Galerne, J.L. Martinand, G. Durand and M. Veyssie, Phys. Rev.
Lett., 29 (1972), 562.
30. F.M. Leslie and S.P.A. Gill, Ferroelectrics, 148 (1993), 11.
31. F.M. Leslie, I.W. Stewart, T. Carlsson and M. Nakagawa, Continuum
Mech. Thermodyn., 3 (1991), 237.
32. S.P.A. Gill and F.M. Leslie, J. Mech. Phys. Solids, 40 (1992), 1485.
613

EXTENSIONAL FLOWS
Christopher J S Petrie

Department of Engineering Mathematics,


University of Newcastle upon Tyne, Newcastle upon Tyne, NE1 7RU, UK.

1. I N T R O D U C T I O N

The simple, practical, definition of extensional flow is a flow in which the


velocity vector can be expressed in the form

u = elx , v = e2y , w = e3z (1)

in a fixed rectangular Cartesian coordinate system, (x, y, z), where (u, v, w) are
the components of the velocity vector and el, e2 and e3 are the principal rates of
strain, (which may be functions of time). For an incompressible fluid, the sum
of the principal rates of strain is zero. Uniaxial extension is obtained when

el = k ; e2 = e3 = - 8 9 (2)

For many fluids, including all Newtonian fluids, there is no more difference
between shear and extension than there is in linear elasticity. Indeed, Trouton
[1] introduced what we now call the elongational viscosity (his "coefficient of
viscous traction") because he wished to obtain the conventional viscosity (the
shear viscosity) of some highly viscous fluids. The elongational viscosity is the
ratio of stress to rate of strain in the steady uniaxial extension of a uniform
cylinder of material. For an incompressible Newtonian liquid Trouton proved
theoretically (and demonstrated experimentally) that this is three times the shear
viscosity, r/o,

r/E = 37/0, (3)

just as the Young's modulus is three times the shear modulus for an incompress-
ible Hookean solid.

We find that the behaviour of polymeric liquids and of suspensions of long


slender particles (fibre suspensions) is markedly different in shear and extension.
The ratio of elongational viscosity to shear viscosity, which we call the Trouton
614

ratio, is no longer three, or even close to three. We associate this behaviour


with the existence of some sort of structure within the fluid which can become
markedly anisotropic during flow, for example when fibres or extended polymer
molecules become aligned parallel to one another and to the flow direction.
This aspect of the behaviour of the fluids mentioned is highly significant and,
especially in the case of dilute polymer solutions, can give rise to Trouton ratios
of 100 or 1000 and hence to stresses which are two or three orders of magnitude
larger than in shear.

1.1. Definitions
The Society of Rheology, in an effort to keep some sort of order in the use
of terms by rheologists, has provided an unambiguous definition of "tensile
viscosity" [2] - the quantity we have referred to as elongational viscosity, also
known as uniaxial extensional viscosity. Tensile viscosity is defined as follows:
A material is subjected to homogeneous simple extension, i.e. to a flow which
is spatially uniform, with constant rate of strain, ~, in the x~-direction and - 89
in every direction perpendicular to the x 1-axis. The ratio of "net tensile stress",
O'E ~ fill --flEE, to rate of strain is monitored as a function of time and the
"tensile viscosity" is defined as

(4)

This definition, of course, says nothing about methods of, or even the feasibility
of, experimental realization. The notation r/T is sometimes used, instead of r/E,
in honour of Trouton, and the elongational viscosity may be called the Trouton
viscosity.

If surface tension is significant, we need to be clear that the "net tensile stress"
a E in our definition, Equation (4), is the measured stress corrected for surface
tension according to
Applied force Coefficient of surface tension
O"E = -- . (5)
Area Radius
We also need to be clear that the use of O E ~ O l l - - a 2 2 i s , in any case, valid
only for incompressible fluids. This matter is discussed further elsewhere [3,4].
For simplicity, our discussion of other extensional flows, Equations (8), (10) and
(11) below, refers to incompressible fluids in the absence of surface tension.

As well as uniaxial extension, we may consider equibiaxial extension,


el = e2 = ~ ; e3 = - 2 ~ (6)
and planar extension,
el = ~ ; e2 = 0 ; e3 = - + . (7)
615

The equibiaxial extensional viscosity [2] is defined by


r/B(~. ) = O'11 -- 0"33 (8)

and for a Newtonian fluid


~Tn = 6770. (9)
Equibiaxial extension is kinematically the reverse of uniaxial extension, but
significantly different in terms of the effect of the flow in tending to align long
molecules or fibres. The idea that the functions r/E(~) and r/B(~) may be regarded
as the same function, for positive and negative values of the rate of strain in the
direction of the axis of symmetry, is not a particularly helpful one since there is
no reason at all for supposing that the values of this function over the two ranges
of values of its argument are in any way connected (except by continuity in a
mathematical sense, which corresponds to Newtonian behaviour in the limiting
case of small rate of strain).

Planar extension is sometimes referred to as "pure shear" but it must be clearly


understood that it is qualitatively different from simple shear, being irrotational
(relative to the usual fixed axes). A particular point of interest for planar exten-
sion is that there are two extensional viscosities, the planar extensional viscosity

,p(g) =
O'll -- 0"33
,
(10)

which refers to the tensile stress required to stretch the material in the x 1-
direction and a second "cross-viscosity"
r](0)(~ -) = O"22 -- O"33 (11)

which refers to the tensile stress required to prevent deformation in the neutral
direction (the x2-direction). The theoretical relations for a Newtonian fluid, with
shear viscosity r/0, are

Tip - - 477o ; r/(~ = 2770. (12)

The notation 7/(2~ is a simple example of the general notation [2,5-7] for general
extensional flows. If el = ~ is the largest (positive) rate of strain, then we may
define m such that e2 = m~ and then
el = ~ ; e2 = m~ ; e3 = - ( 1 + m)~ (13)
for an incompressible fluid. The parameter m, which we take to be independent
of time, describes the geometry of the extensional flow, with m = 1 for equibi-
axial extension, m = 0 for planar extension, m = - 89 for uniaxial extension and
616

< m < 1 in general. The two extensional viscosities for this general flow
2 --
are

~m)(~.) = fill -- 0"33


(14)

and
Tl(m)(~.) = 0"22 - - 0"33 . (15)

These are equal to one another for equibiaxial extension and the second is zero
for uniaxial extension while for all other extensional flows we have the two
physical quantities which we may try to measure. As well as for the three
standard cases, m = - ~1, m = 0 and m = 1, experiments have been carded
out by Demarmels and Meissner [8] for m = +89 This flow has been referred
to as "ellipsoidal extension" but may best be visualized as an unequal biaxial
extension, with stretching at rates ~ and 1~ in two perpendicular directions.

2. T H E E X P E R I M E N T A L E V I D E N C E

Experiments for determining extensional viscosity are difficult to carry out,


and even where there is a consensus about results it is important to remember that
this consensus is the fruit of much painstaking experiment. For polymer melts, it
has been possible for some time to obtain reliable values for low and moderate
rates of strain; the review by Meissner [9] is a useful source of information.
Different polymers show different behaviour in the variation of Trouton ratio
with rate of strain. As has been remarked above, the value of the Trouton ratio
can be large, but for polymer melts it does not reach such extremely high values
as for polymer solutions.

The situation is much less clear for polymer solutions, as has been summa-
rized by James and Waiters [10]. The experiment which corresponds to the
definition of extensional viscosity has not been easy to perform for these mobile
(low viscosity) fluids - in fact it had been judged impractical until the work
of Matta [11] and Sridhar [12-14]. It is, however, generally accepted that the
Trouton ratio for polymer solutions is large. Even with the considerable uncer-
tainty surrounding interpretation, we believe that estimates of 100 or 1000 have
genuine physical significance, whether or not the measured quantity is truly an
extensional viscosity. Suspensions of long fibres generally behave in a similar
manner to polymer solutions.

In all cases, at least for fluids which are unoriented in their natural state (i.e.
when at rest and relaxed), the limiting values of extensional viscosity for small
rate of strain agree with the Newtonian predictions. There is no clear evidence
617

of what happens at high rates of strain, where experiments are difficult, but what
evidence there is supports the theoretical conclusion [7] that uniaxial and planar
extensional viscosities should be the same. This theoretical study shows two
classes of asymptotic behaviour at large rate of strain in planar and uniaxial
extension. When we use constitutive equations appropriate to polymer solutions
or fibre suspensions, there is an "upper Newtonian" r6gime with a constant
extensional viscosity corresponding to fully extended and fully aligned molecules
or fibres. With the constitutive equations that were developed for polymer melts
a decreasing extensional viscosity was found. The stress must grow with rate of
strain, but can do so at a rate less than linear (e.g. (rE ~ In ~ for the Phan-Thien-
Tanner model) [7]. This decrease sometimes follows a maximum in extensional
viscosity at some intermediate rate of strain.

It is extremely important to distinguish clearly between measurements which


do give the extensional viscosity functions directly, i.e. experiments in which
steady spatially homogeneous extensional flow is attained (to a reasonable ap-
proximation), and measurements which give a stress growth function or transient
extensional viscosity (or something even less well-defined). This is not to say
that the experiments in which the ideal extensional flow is realised are necessar-
ily the best experiments. If we want measurements that will allow us to predict
fluid flow and stress levels in geometrically complicated unsteady flows, we may
well find that some sort of transient extensional viscosity is more appropriate.
For fundamental understanding of fluid deformation and flow, however, the true
extensional viscosity remains an important physical quantity which we should
like to measure accurately and reliably.

2.1. Extensional viscosities of polymer melts


The most extensively studied material, historically, is low-density polyethylene
(LDPE). The most reliable experiments, of Meissner and co-workers (past and
present) [15,16], show the elongational stress growth function (commonly called
the transient extensional viscosity) for experiments at constant rate of strain. This
function, for uniaxial extension, is compared with 3 times the analogous function
in shear at low rates of strain (this is regarded as the linear viscoelastic response).
Typical behaviour of LDPE is to show the linear viscoelastic response at very
low rates of strain (such as 0.001 s -1) and a greater stress at moderate rates of
strain (typically 0.01 s -1 up to 1 s -1). This behaviour is sometimes referred to as
"strain hardening" or "deformation hardening" (although "strain rate hardening"
might be a more appropriate term). This behaviour is also reported [17] for
HDPE, which does show "strain hardening", although to a smaller extent.

For other geometrical forms of deformation, there are fewer results [5,8,18].
Most work until recently has been on polyisobutylene (PIB), because it can
be investigated at room temperature. PIB behaves in a similar way to LDPE
in uniaxial extension. In biaxial extension the response is closer to the linear
618

viscoelastic response (multiplied by the expected factor of 6) at moderate rates


of strain. In planar extension the first stress growth function (corresponding to
~Tp) lies between the uniaxial and biaxial responses (all suitably scaled), while
the second stress growth function falls below the linear response after a fairly
short time.

Improved versions of the extensional rheometers have allowed different defor-


mation histories (extension at constant stress or force as well as at constant rate
of strain) and measurement have been made on polymethylmethacrylate [19],
polystyrene [19] and polybutadiene [20]. Another recent development of con-
siderable importance is the new multiaxial elongational rheometer from Meissner
[21] which has been used for extensional flow investigations on polymer melts
and blends at elevated temperatures [22].

2.2. Extensional viscosities of polymer solutions


The history of extensional flow measurements for polymer solutions is full
of apparently contradictory results [4,10]. The major source of this is the use
of different experimental techniques, none of which, until recently, are even
claimed to reproduce steady uniform uniaxial elongation. Among the techniques
that have been used are
(a) fibre spinning and open syphon (Fano) flow,
(b) contraction flow (into an abrupt contraction or through an orifice),
(c) converging flow (through a smoothly tapering channel),
(d) stagnation-point flow (e.g. use of opposed jets),
(e) falling weight and controlled extension.
and each of these has its own advantages and drawbacks.

James and Waiters [10] have collected results for one polymer solution and
show, in their Figure 2.1, values claimed for the extensional viscosity over a
range of three decades (10-10,000 Pa.s) for rates of strain in the range 1-100
s -1, with experimental results scattered over the whole of this region. The
explanation of this is not that the experiments are all meaningless, but that,
since we are dealing with a viscoelastic material, the differing flow histories
and the fact that steady flow is not attained made considerable differences to
results. The various collaborative testing programmes reviewed in [4] (one of
which produced the results in [10] mentioned above) have made it clear that
differences in the material being tested do not explain these differences. In
short, the reported quantities are not "extensional viscosities".

While the variety of experiments and corresponding predictions may not be


diminished in the near future, recent work does suggest that reliable extensional
viscosities (in the sense that experiments are reproducible in different laborato-
ries) may be obtained for polymer solutions.
619

2.3. Other experimental measurements


The measurement of extensional viscosities and stress growth functions by
no means exhausts the possibilities for useful experiments on the extensional
flow behaviour of polymeric liquids. Meissner [9] has consistently urged that
recoverable strain should be measured as well as total strain, and this has been
done for a number of polymer melts. Recent measurements on polymer solutions
[23-25] have demonstrated interesting behaviour in stress relaxation at the end
of a period of steady uniaxial extension.

Attempts have also been made to investigate further the fundamental difference
between shear and extension by looking at flows in which the history of the
principal extension ratio is the same for a shear flow and an extensional flow
[26,27]. Experiments were carried out by Ztille et al. [26] both with ~ increasing
exponentially and with ~ constant, where ~ is the principal extension ratio. In
each case both shear and uniaxial extension were used. The results for LDPE
showed that at large strains and large rates of strain the rheological behaviour is
determined primarily by the history of the principal extension ratio, ~(t), and not
by whether the flow is an extensional flow or a shear flow. This means that they
were able to show deformation hardening and deformation thinning behaviour in
extensional flow of LDPE, by choosing in the first case elongation at constant rate
(~ = ~f = ~ exp(~t)) and in the second case, elongation with ~ constant. Similarly
in shear the apparent viscosity (measured at constant shear rate) is a decreasing
function of shear rate while the corresponding quantity for "exponential shear"
increases with shear rate. Samurkas et al. [27] point out that to conclude from
this that the material is deformation hardening in exponential shear may be
misleading. Their point is that under exponential shear a Newtonian liquid will
show a similar behaviour. Indeed Ztille et al. [26] show (in their Figure 4) that
the measured stresses are generally lower than the predictions for exponential
shear of linear viscoelasticity. Samurkas et al. [27] compare exponential shear
with planar extension, and present their result in term of the damping function
for the Wagner model (see below). Their conclusion is that the damping function
obtained from steady shear flow measurements is better than that obtained from
planar extension in predicting behaviour in exponential shear. This tends to the
opposite conclusion to that reported above [26]. However neither set of results
is conclusive, and direct comparison is difficult because of the different ways in
which the data are presented.

3. BASIC P R O P E R T I E S OF SOME C O N S T I T U T I V E EQUATIONS IN


EXTENSION

The extensional viscosity (in uniaxial extension), as well as the planar and
equibiaxial extensional viscosities have been tabulated for a number of simple
constitutive equations [3]. We reproduce some of these results below (Table 1)
620

and discuss two classes of constitutive equation which have become popular in
recent years. These are the Wagner integral model [28] (often used for polymer
melts) and the FENE dumbbell models [29] for polymer solutions. Larson [30]
gives useful discussion of constitutive equations which is both more thorough
and more up-to-date than [3].

Many of the results in [3] are for special cases of the Oldroyd eight-constant
model,

T + ,~1 '~ - / Z l (T- D + D . T) + vl tr(T- D) I + I~o tr(T)D


i" ]
= 27?0 [D + )~2 - 2#2D" D + y2 tr(D. D) I] , (16)

in which the notation T denotes the corotational derivative of the extra-stress


tensor, T, and D is the rate of strain tensor. The generalized convected Jeffreys
model

is one useful example, with A = 0 giving the generalized convected Maxwell


model. A generalized convected derivative
D V
T =T+{(T. D + D . T) = T - a ( T - D + D - T) (18)

is sometimes defined, with a = 1 - ( being used in [3] and elsewhere, and


Equation (16) could use the generalized convected derivative if a = #l/A1 =
//2/,~2- The parameter ( takes values in the range 0 <_ ( _< 2 and correspondingly
1 >_ a >_ - 1 . The upper convected derivative corresponds to ( = 0 (i.e. a - 1)
while ( - 1 (or a - 0 ) gives us the corotational derivative and ( = 2 (or a = - 1 )
gives the lower convected derivative.

In terms of molecular (or micro-structural) models, only the upper and lower
convected derivatives are obtained from so-called affine deformation models.
In these models, hypothetical micro-structural elements in the fluid, (such as,
for example, either network junctions or dumbbells) move affinely with the
fluid. The parameter ( can be interpreted in terms of a non-affine deformation
of elements in the fluid, i.e. a slip between micro-structural elements and the
surrounding fluid. This leads to some predictions in shear flow which are re-
garded as undesirable and in extreme cases are contrary to physical intuition
(e.g. negative stresses in response to a sudden large shear deformation). There
are other consequences in extensional flows, such as non-existence of solutions
to flows problems for some values of the parameter (, e.g. ( > 0.5 [3], but
the whole range of non-affine convected derivatives are by no means ruled out
621

Model tiE ~B ~p r](0)

Newtonian 3770 67/0 4r/o 27/0


3r/ 60 47/ 2q
UCM (1+)~)(1-2,~) (1+4a~)(1-2)~) (1+2,~)(1-2,~) (1+2)~)
3rl 67/ 4O _ 27/
LCM (1+2,~0(1- a0 (l+2aO(1-4aO 0+2aO(1-2aO (1-2a~)
3r/ 67/ 40 27/
GCM (l+a,~)(1-2a)~O (l+4a,~)(1-2aa~) (l+2a)~)(1-2a,~) (l+2aag)
Corotational 3r/ 67/ 4r/ 27/
1-Ag(I+2M) ~_ l+2Ag(1-4ag) 4r/(1-4aAg
2) 27/ I+2Ag
Oldroyd B 3r/(l+,~)(l_2M) o~'](1+4)i~)(1-2~) (1+2~)(1-2~) (l+2,~g)
l+Ag(1-2)~) 1-2Ag(l+4~)
6r/ (1-4~)(1+2~) 4r/(1-4~Ag
2) ,-,
z r / ~1-2Ag
Oldroyd A 37/(1-~)(1+2~) (1+2~)(1-2~)

Table 1. Extensional viscosities for some simple constitutive equations.

The abbreviations in Table 1 for rheological model names are " U C M "
for the upper convected Maxwell model, "LCM" for the lower con-
vected Maxwell model, "GCM" for the generalized convected Maxwell
model. The corotational result applies to Maxwell and Jeffreys models.
The Oldroyd fluids B and A are exactly equivalent to upper and lower
convected Jeffreys models, respectively.

by this behaviour (unless one adopts a very strict position on the mathematical
requirements for a constitutive equation [31-33]).

The material functions for the eight-constant Oldroyd model are

1 -/z2~" + (/Zlfl2 - 3T2)6"2 (19)


r/z = 37/0 1 -- #1~ + (#21 -- 371)~"2

1 + 2/z2~ + 4(/Zl/Z2 - 37-2)c2


(20)
r/a = 67/0 f + 2/Zl~ + 4(~u2 37-1)~2

1 + 47-2~2
r/p = 47/o 1 + 47-1~2 (21)

1 - 2(#1 - ~ 2 ) c + 47-2c 2 + 8(,tt271 - ~ 1 7 2 ) ~'3 (22)


r/(2~ = 2r/o 1 + 4 T1 ~-2
where we have defined parameters 7-1 and 7-2:

7"1 = (/tO/2)(2/z1 -- 3Vl) -- ~1(/Zl -- 171) , (23)


622

7-2 = (#0/2)(2#2 - 3y2) - / t l ( t t 2 -/,'2) . (24)


These are related to the parameters al = r~ + A2 and a2 = 7-2+ A1A2 which occur
in the viscometric functions
1 + cr2k2
O(k) = ~o 1 + 0"1k 2 ' (25)

~t~l(k ) ---- 2 A l r / ( k ) - 2A2r/o, (26)


~2(k) = -(A1 - / t l ) r / ( k ) + (A2-/t2)0o (27)
for shear rate k.

One interesting aspect of this is that all the parameter combinations which ap-
pear in the expressions for the extensional viscosities also appear in the expres-
sions for the viscometric functions so that, in principle at least, the extensional
flow behaviour of the eight-constant Oldroyd fluid can be predicted from com-
plete data on shear flow. In fact, data for shear flow allow the calculation of the
characteristic times A1 and A2, which do not appear in the extensional viscosity
formulae, so that in a sense, for this model, the viscometric functions contain
more information than the extensional viscosity functions. Note that the actual
calculation of the parameters from viscometric data is not a trivial task and, for
example, to obtain the parameters ttl and #2 which are essential for extensional
viscosities one must have very good data on the second normal stress function,
~2(k).

3.1. Wagner and Kaye-BKZ constitutive equations


The Wagner equation started out as a form of the Kaye-BKZ equation in
which the nonlinear and time-dependent parts of the material behaviour are
assumed to be independent in the sense that the kernel of the constitutive equation
may be factorized. It is, at present, probably the most successful compromise
between simplicity and generality in a constitutive equation for the quantitative
description of the rheological behaviour of polymer melts.

The model may be written

T = M ( t - t')h(I~, I2)Ct~(t ') dr' (28)


Oo

in which M ( t - t') is the memory function, c~-l(t ') the relative Finger deforma-
tion tensor and h(I1,12) the damping function, which introduces non-linearity of
dependence on deformation through the invariants I1 and 12 of the Finger tensor.
The product h(I1, I2)Ctl(t ') may be thought of as a non-linear measure of defor-
mation [30,34]. If h = 1 and M ( s ) = ( G / A ) e x p ( - s / A ) we recover the integral
equivalent of the upper convected Maxwell model. A sum of exponentials for
M ( s ) gives the usual discrete relaxation spectrum. Equation (28) is not, in fact,
623

a Kaye-BKZ equation unless the damping function is independent of the second


invariant, I2 - see below, Equation (35).

In a general extensional flow, Equation (13), we may define the relative


Hencky (or logarithmic) strain, e, and the relative extension ratio, g:

e = +(t - t ' ) , g = e ~ = e eCt-t'~ (29)

and then we have

ctl(t ')= 0
00 ) (30)
0 0 (-2(re+l)

with invariants
I1 = g2 + g2m + g-2(rn+l) , 12 = g-2 + ( - 2 m + ~2(m+l) . (31)

The damping function is chosen to fit data from shear or extensional viscom-
etry (or preferably both). Early attempts used functions like h = e x p ( - n ~ ) for a
shear strain of ~ = k ( t - t') (and we could write ~ = vql - 3, noting that I1 = 12
in simple shear). In order to fit data from different flows (shear and extension),
the introduction of an invariant

K = ~//311 +(1 -/3)I2 - 3 (32)

has been successful [35] and better fits to data have been obtained using a sum
of exponentials

h(I1,12) = f e - n l K + (1 - f)e -n2K . (33)

Wagner and Laun [36] found that values f = 0.57, n l = 0.310 and n 2 = 0.106
gave a good fit to shear data for the LDPE Melt known as "IUPAC A", and
/3 = 0.032 gave a good fit for uniaxial extension also. Another form of damping
function which is popular is due to Papanastasiou [37]
1
h(I1,12) = 1 + a K 2 " (34)

Larson [30] notes that neither of these functions, Equations (33) and (34),
fits data for biaxial extension well and he proposes a model which is strictly of
the Kaye-BKZ form, with a potential function, U, and the Cauchy deformation
tensor, Ct(t~),

0U(I1, I2) 0U(I1, I2)ct(~,)] d r ' . (35)


T = ft__~ M ( t - t') 2 0-]-i C t l ( t ') - 2 012
624

The potential function used here is

U(Ii,12)=~In 1+ (I-3) (36)

with both c~ and I depending on the invariants I1 and 12:

[ Cl (I2 -- I1)3 ]
a = CO+ C2 tan -1 1 +(12--I1)2 , (37)

I = (1 - 3)11 + X/1 + 2fli2 - 1 (38)


and, for IUPAC A, co = 0.20, cl - 0.05, c2 - 0.121 and 3 = 0.1. This does a
better job of fitting all the data than the Wagner equation without the Cauchy
deformation tensor, Equation (28); Equation (28) also has the defect of predicting
a zero second normal stress difference in simple shear.

Similarly, Demarmels found [8,38] that the Wagner equation does not give a
consistently good fit to data from shear and several different extensional flows
for PIB and that a better fit can be obtained with a model which introduces the
Cauchy deformation tensor, Ct(tt), as well as the Finger tensor. This gives us
an equation of the form proposed by Rivlin and Sawyers, which we write here

T = f t_ o M ( t - t ' ) [ h l ( I i , I 2 ) C t l ( t t ) + h2(Ii,I2)Ct(tt)] dt t (39)

and Wagner and Demarmels [38] propose a form for the two damping functions,
for one particular PIB melt,
1
hl(I1, I2) = (1 +/3)h, h2(II,I2) =/3h, h = 1 + a'~*lv'k~- 3)(12 - 3) (40)

with constant values of the parameters a = 0.11 and 3 = - 0 . 2 7 for the particular
PIB melt studied in [8]. The parameter/3 in this model, which gives the ratio of
second to first normal stress differences in simple shear, is important in general
extensional flows but unimportant in uniaxial extension. This choice of functions
h l and h2 is not obtainable from a potential, so Equations (39) and (40) do not
give a Kaye-BKZ equation.

A variation in use of the Kaye-BKZ model [39] seeks to express the de-
pendence of the kernel on the invariants I1 and 12 through dependence on the
principal stretches (or their squares, which are the eigenvalues of the Finger
tensor). This is motivated by the success of some strain-energy functions for
rubbers, but has not been found to be successful so far. It has not proved pos-
sible to pick a simple dependence on the principal stretches which gives a good
fit to shear and extensional data with the same values of parameters.
625

One final, well-known, defect of these models has been considered by Wagner
[40]. This is the fact that even the best of the models with a damping function
or non-linear deformation measure has difficulty in fitting data obtained in flows
where the fluid experiences flow reversal. The simple example of this is one step
strain followed by a second step strain in the opposite direction. The problem
can be explained in terms of a temporary network model in terms of irreversible
loss of entanglements. This suggests a damping functional as a replacement for
the damping function. The damping functional proposed is the smallest value
over the time interval of the conventional damping function. For a motion with
a "nondecreasing deformation", defined as a deformation for which the damping
function is a nonincreasing function of time, the damping function is correct.
If the motion involves a "decreasing deformation", a smaller damping factor is
used. This approach has some success in predicting recovery (elastic recoil)
after uniaxial extension [40].

3.2. FENE constitutive equations


The FENE (finitely extensible nonlinear elastic) dumbbell model is found to
be useful for polymer solutions, though there are many questions, both about
the theoretical foundations of the model and about quantitative agreement with
data on polymer solutions. There are a number of variants on the model, which
we shall discuss briefly. First we outline the basic features of the model, while
avoiding a formal derivation of the equations; for details see, for example,
[29,30].

The simplest FENE dumbbell model represents a polymer molecule in solution


as an isolated dumbbell (two beads connected by a spring) whose motion is
governed by a balance between a spring force tending to contract or coil the
molecule and hydrodynamic drag on the dumbbell ends due to the solvent, which
tends to stretch the molecule and align it with the flow. A Hookean spring has
the disadvantage of allowing the molecule to extend indefinitely, so a non-linear
spring law, commonly the Warner spring law,

HR
Fa = 1 - ( R 2 / L 2) ' (41)

is used. Here R is the end-to-end vector for the dumbbell (the molecule), H
is the spring constant, R = IR] is the end-to-end length and L is the maximum
permitted end-to-end length for the dumbbell, so that R < L. The hydrodynamic
drag force on a bead is given by a drag coefficient (d multiplying the velocity
of the bead relative to the solvent. This gives us the relaxation time, A =
( d / ( 4 H ) , which is a characteristic time for an individual dumbbell to come to
equilibrium under the competing action of the drag and spring forces. There
is a second characteristic time, 0 = L 2 ( d / ( 1 2 k T ) which is associated with the
balance between hydrodynamic drag and Brownian diffusion and a modulus,
626

G = n k T , (as in rubber elasticity, where n is the number of dumbbells per unit


volume). We define the "FENE parameter",

b = 30/A = n H L 2 / G (42)

and can also introduce a characteristic molecular dimension a by

a 2 = 3L2/b = 3 k T / H . (43)

This gives the equilibrium length of a dumbbell (when the solution is at rest),
possibly multiplied by a factor like b/(b + 5), depending on the precise details
of the FENE model.

These ideas are used to obtain a configuration tensor, A = (RR) and the
polymer contribution to the extra-stress tensor. The mean, ( ) , is an ensemble
average involving the distribution function, g,(R), and the trick is to obtain
equations without having to calculate if,. This is most usually done by making
the Peterlin approximation, which involves pre-averaging the end-to-end length,
so that
HR
Fd = (44)
1 -- ( ( R 2 ) / L 2)

instead of using the true ensemble average of Equation (41). This leads to the
FENE-P model, with configuration evolution equation
v 1 L2
XA + A = ~ I (45)
1 - ( R 2 / L 2) (b + 2)

and the extra-stress


T= G(b/L2) Gb
A I + 2r/~D (46)
1 -- ( R 2 / L 2) (b + 2)
in which r/~ is the solvent viscosity.

The Chilcott-Rallison (or FENE-CR) model [41,42] is


v 1 (L2/b)
AA + A- I (47)
1 - ( R 2 / L 2) 1 - ( R 2 / L 2) '

T= G(b/L2)
[A - (L2/b)I] + 2~/,D. (48)
1 - ( R 2 / L 2)
This model has qualitatively the same behaviour in extensional flows as the
FENE-P model, as far as is known from investigations to date [4]. It has the
property of a constant viscosity in shear, unlike the FENE-P and FENE models
627

which are shear-thinning [41], which is both a simplification and is desirable for
modelling the behaviour of Boger fluids.

One shortcoming of all the FENE models discussed above, as far as fitting
data on polymer solutions is concerned, is that they have only a single relax-
ation time. A recent discussion of multimode models (i.e. bead-spring chains)
by Wedgewood, Ostrov and Bird [43] points out that a straightforward general-
ization of the simple dumbbell leads to a complicated set of coupled nonlinear
differential equations. They point out that some earlier attempts to analyse this
contain serious errors and go on to propose a further approximation, the FENE-
PM model. The model consists of a set of N - 1 nonlinear springs joining
N beads and the end-to-end vector for the i-th spring is denoted by Ri. The
FENE-PM force law is taken to be
Fi = HRi (49)
1 --(iV -- I) -I ~N~'((R2)/L2)

and the M in FENE-PM stands for the mean value taken in the denominator
of Equation (49). The M is also used to denote "multimode", but it is neces-
sar3, to remember that the FENE-PM model is not just a multimode FENE-P
model, for the reasons of complexity to which we have alluded. Even so, the
model obviously remains more difficult to use than the single mode FENE-P and
FENE-CR models and the questions to be settled are whether the extra effort is
adequately rewarded and whether the approximation introduces any undesirable
side-effects.

As examples of unwanted side-effects, the fact that the FENE-P model leads
to equations for steady extension with multiple solutions may be instanced.
This is a comparatively minor matter, which can be resolved [4] by analysis
of the full equations, as is discussed below. A more interesting matter is the
demonstration by Keunings [44] that the dumbbells in the FENE-P model do
not actually behave as finitely extensible dumbbells. A simulation shows that a
noticeable fraction of the dumbbells exceed the supposed maximum length L.
This arises from the fact that the approximation leads to a distribution function
for end-to-end lengths which is Gaussian (with an infinite tail), while for a true
FENE model the distribution must have a cut-off at L and hence must be non-
Gaussian. A comparison between the FENE-P and FENE model predictions
(using a stochastic simulation) [44] shows that the rheological effect of the
Peterlin pre-averaging approximation is seen in a much more rapid increase in
the tensile stress during the start-up of an extensional flow. The stress growth
is smoother and somewhat slower for the tree FENE model. The steady state
stresses are the same.

None of the FENE models yields explicit formulae for the extensional vis-
cosities (unless one wishes to write down formally the algebraic solution to a
628

cubic equation). It is therefore less easy to make simple statements about their
properties. In the limit of small extension rate, the Newtonian ratios between
the extensional viscosities and the shear viscosity are recovered. In uniaxial
extension the viscosity curve is S-shaped and at large extension rate, an "upper
Newtonian" rrgime is obtained [45],
tiE = 3q~ + 6 G ~ = 3rl~ + 2b~Tp (50)
where r/p = AG is the polymer contribution to the viscosity. As far as the FENE-
PM model is concemed, the model can be expected to improve the quantitative
fit to data. No surprises in the qualitative behaviour have come to light so far.

4. S O M E M A T H E M A T I C A L ASPECTS OF E X T E N S I O N A L F L O W

Once we have a constitutive equation and a particular flow to analyse, whether


exactly or approximately, we can address mathematical questions of existence
and uniqueness of solutions. These are by no means trivial or unnecessary
exercises for the nonlinear systems with which we are faced, and there are
particular problems in analysing boundary-value problems, even for simple one-
dimensional systems (and ordinary differential equations). In this section we
discuss one existence problem in detail, recognizing that rheologically it is rather
simple, as an illustration of the sort of problem that may have to be faced. It is an
open question as to whether the limitations that are uncovered are seen as defects
in the rheological model used, defects in the fluid dynamical approximations or
merely as a warning that predictions obtained with the model must always be
treated with a modicum of scepticism.

After this, we shall discuss an example of lack of uniqueness of solutions


to the steady-state equations. This has been mentioned above, and is a likely
consequence of non-linearities of the sort we are introducing. The stability of
such simple solutions, which is described by the full, time-dependent, dynamical
equations, can be used to make a choice between the possibilities. In some cases,
such as the example below, we can in fact go further than this and make some
very general claims about the global behaviour of the system, which rule out
some physically unrealistic solutions even though they do, formally, satisfy the
steady-state equations.

4.1. Existence of solutions to a boundary-value problem


Steady fibre spinning of a convected Maxwell model can be posed as a
boundary-value problem which only has solutions for a limited range of values
of one boundary condition. We consider the steady axisymmetric extensional
flow given by Equation (2) with the spatially varying rate of strain k = U ' ( X ) in
which U is the velocity at distance X from the start of the fibre (as it emerges
from a spinnerette or die). The fibre take-up is at X = Ls where the velocity is
629

set to be U~, while the initial velocity is Uo at the spinnerette. If the volumet-
ric flow rate, Q, is given, we can relate the cross-sectional area, A ( X ) to the
velocity since we make the basic assumption that, to a first approximation, all
quantities are uniform across the fibre.

In the absence of gravity, inertia, surface tension and air drag, the five equa-
tions governing this flow for an isothermal incompressible upper convected
Maxwell model, Equation (17) with ( = 0 and A = 0, are mass conserva-
tion, equilibrium of forces axially and radially and axial and radial components
of the constitutive equation:
U(X)A(X) = Q, (51)
{Txx(X) - P(X)} A(X) = F, (52)
TRR(X) -- P ( X ) = O, (53)
T x x ( X ) + A {U(X)TJcx(X) - 2 U ' ( X ) T x x C u = 2r/U'(X) (54)
and
TRR(X) + A {U(X)T~RR(X)+ U'(X)TRR(X)} = - r l U ' ( X ) (55)
in which F is the constant force at any cross-section and P ( X ) is the consti-
tutively undetermined hydrostatic pressure (which is, in effect, calculated from
the radial force balance, Equation (53)). Equations (51)-(53) can be combined
to give
Txx(X)- TRR(X)= F U ( X ) / Q (56)
and so we have a second-order differential-algebraic system, Equations (54)-(56),
for the velocity and two extra-stress components, with three conditions,
Txx(O) = To, U(O) = Uo , U(L~) = U1, (57)
two of which are needed for the differential equations and the third to determine
the force, F, which has to be applied to maintain the specified take-up velocity.
The extra-stress value, To, is all that is required from the flow history, corre-
sponding to the internal structure of the fluid at X = 0, i.e. to that which has
been determined by the flow of the material upstream of the spinnerette.

This formulation shows the nature of the problem as a two-point boundary-


value problem, and the question we can address is whether there are solutions
for any draw ratio, DR = U1/U0, and how the answer to this may be affected by
the value of To, i.e. by the flow history of the material. In dimensionless form,
with

u
u , x
x=-;--, c~=
~Uo t,= ,1Q , 7-o=
ToQ (58)
Uo ~s L~ AUoF FUo
630

the system of equations may be reduced to


c~2u" = (1 + 2c~u')(1 - om')o~u'/u - 3it (cru'/u) 2 (59)
with # undetermined, so that the three conditions
1
u(0) = 1 , u'(0) = go = 3 a ( # + 70) - 2c~ ' u(1) = DR (60)
may be satisfied. It can easily be shown that, if DR > 1, it is necessary that
90>0.

What we can prove is that the value, u(1), of the solution of Equation (59) at
the take-up, x = 1, is bounded above by a function of the Deborah number, a,
whatever the values of To and #. This means that there is an upper limit to the
draw ratio, DR, for any given Deborah number, regardless of the flow history
and regardless of how great a take-up force is applied.

The proof involves considering the comparison equation


cr2v" = (1 + 2crv')(1 - a,v')crv'/v (61)
which can be integrated, with initial values v(0) = 1, v~(0) = 90 to give
cry'= (v 3 + K)/(v 3 - 2 K ) (62)
where
K = (cr9o - 1)/(2c~9o + 1) (63)
and then
x
= fv y3 _ 2K dy. (64)
ce y3+K
Equation (64) gives v implicitly as a function of x/o~ and K ; we may write
y3 2K
9~(v, K ) = flv y3 + K dy . (65)

and the solution to 9~(v, K ) = x/o: is written


v(x; a, 90) = ~ ( x / a , K ) . (66)
We note that the condition ago > 0 implies that - 1 < K < 0.5 and can establish
that dv/dK > 0 under this condition. Hence, for all admissible values of 90,
v(1; c~, go) < ~(1/a,, 0.5) (67)
and we can calculate the fight-hand side of Inequality (67) numerically. For
example, if (x = 1/19, ~,(19,0.5)= 20.6328.

Finally a comparison theorem applied to Equations (59) and (61) shows that
DR = u(1; c~,#,go) < v(1; c~,go) < g,(1/~,0.5) (68)
which gives us an upper bound on DR for any chosen c~, irrespective of the
values of # and go.
631

4.2. Multiple steady state solutions


We consider uniform uniaxial extension, Equation (2), which starts at time
t = 0. We will need to specify an initial configuration in order to obtain a
specific solution to the evolution equations for the configuration tensor (and
hence the stress). We consider the FENE-P model, Equations (45) and (46) and
define the dimensionless configuration variables
A~ - A22 tr A All + 2A22
y= L2 , z = L2 = L2 . (69)
The first of these, V, may be interpreted as describing the degree of alignment
of the dumbbells with the direction of elongation while the second, z, gives the
mean end-to-end length of the dumbbells in the flow (i.e. the extent to which
the dumbbells are fully stretched). It is obvious that
0 < z < 1 (70)

and also fairly easy to establish that


-z/2 <_ y <_ z (71)

(see [4,45] where an argument suggested by 0ttinger is set out). If y = z the


dumbbells are fully aligned with the flow, while if y = - z / 2 the dumbbells are
all aligned perpendicular to the flow. The result of substituting the flow field
given by Equation (2) in Equation (45) and expressing the result in terms of the
variables defined in Equation (69) is the pair of evolution equations

A~=sz+
(1)
s 1-z y (72)

m,

~ = 2sy +c (73)
1-z
in which s = ~k is the dimensionless rate of strain (which may be regarded
as a Deborah number) and c = 3/(b + 2) is a constant depending on the finite
extensibility parameter, b. For large values of b, c ~ a 2 / L 2, the equilibrium
mean square end-to-end length of the dumbbells expressed as a fraction of their
maximum length (and this is small).

We may examine the steady-state solutions of this pair of equations and dis-
cover that there are three, of which only one satisfies Inequalities (70) and (71).
It is relatively straightforward, if tedious, show that the unphysical steady-state
solutions are unstable. A more important question is whether, if y and z have
initial values which satisfy Inequalities (70) and (71), i.e. are physically rea-
sonable, the subsequent solution, (y(t), z(t)), remains physically reasonable. We
answer the question by considering the behaviour of solutions represented by
curves in the phase-plane, (z(t), y(t)). We shall prove that solutions which start
632

in the triangle 0 < z < 1, - z / 2 < y < z, for t = 0, remain in that triangle for
all subsequent time, t > 0.

It is easy to see, from Equation (73), that ~. < 0 for z ~ 1 - , so that z cannot
reach the line z = 1 provided that z(0) < 1. Next we consider the angle between
the solution curve as it crosses the line y = z and the normal the that line which
points into the region V < z. If we denote this normal by the vector n, we have
n = ( 1 , - 1 ) . The tangent to the solution curve (in the direction of increasing t)
is parallel to the vector v, given by v = (~,, #). If the angle between these two
vectors is between - 7 r / 2 and 7r/2, i.e. if the scalar product n . v is positive, then
the solution curves all cross the line y = z from the region y > z to the region
Y < z. When Y = z, ~, = y + c so the scalar product is
n. v = (#+c) - (#) = c > 0 (74)
and therefore solutions cannot leave the physically sensible region across the
line y = z.

The same argument may be applied to the line y = - z / 2 , with normal n =


(1/2, 1). Using this value of y gives
il = s z - s z / 2 + z/2(1 - z) (75)
iz = - s z - z / ( 1 - z) + c (76)
and hence the scalar product
n . v = ~ / 2 + fl = c / 2 > 0 (77)
which shows that solution curves cross the line V = - z / 2 from the region
y < - z / 2 into the region y > - z / 2 , as required. We may note, further, that
for s > 0, i.e. for uniaxial extension (rather than for uniaxial compression or
its equivalent, equibiaxial extension), # > 0 on y = 0 for 0 < z < 1 so that in
extension alignment parallel to the flow is favoured (in the sense that if V(0) > 0
we have y(t) > 0 for all t > 0 while if y(0) < 0 we may expect v(t) > 0 at
some subsequent time). Similarly if s < 0 alignment perpendicular to the flow
is favoured, as we would expect in uniaxial compression.

The same result may be proved for the FENE-CR model, Equations (47) and
(48). The differences from the FENE-P model are only in the coefficient of the
isotropic tensor, I, and the effect on the configuration evolution equations is that
c is replaced by c/(1 - z) in Equation (73). Equation (72) remains unaltered and
the arguments used above hold for the direction in which solution curves cross
the lines Y = z and Y = - z / 2 . The behaviour as z ~ 1 is the same, provided
that c < 1 and it is easy to see that this will be the case, since r > 1 would
require b <_ 1 and the model only makes sense for b > 3 (otherwise a 2 > L2).
Hence the result that solutions which start off in the physically sensible region
remain there for all time is established for this model too.
633

5. SOME OUTSTANDING P R O B L E M S IN EXTENSIONAL F L O W

There are many unsolved problems in extensional flow, and so plenty of work
for rheologists to do. The choice of the best constitutive equation is still an open
question- to which the safest answer is often that it depends on the purpose for
which the constitutive equation is required. Good quantitative agreement with
extensional flow data, and data on linear viscoelastic and viscometric properties,
may usually be obtained only at the price of introducing a complicated equation
which may present too great a challenge in the computational simulation of
complex flows. There is, of course, the question of whether the need to choose
a constitutive equation may be avoided by computational methods involving
direct simulation of behaviour at a molecular or micro-structural level. My view
on this is that we still have a need for constitutive equations, as an aid to our
understanding, even if not for our use of computers to solve problems.

A useful perspective on the importance of extensional flows may be gained


from the list of key challenges in polymer processing given by Kurtz [46]. Of
the five key challenges listed, two ("Blown film bubble stability" and "Draw res-
onance") quite clearly involve extensional flow. The other three ("Screw wear",
"Sharkskin melt fracture" and "Scale up problems") may involve extensional
flow as a more or less important feature. Some comments on this and other
issues have recently appeared [47].

5.1. Interpretation of experimental results


Two pieces of work may usefully be mentioned here. The first seemed at first
sight to be a good idea, but probably will not stand up to scrutiny. That is the
idea of a three-dimensional plot of transient extensional viscosity as a function
of strain and time [48-50]. While this appears promising for limited data, its
theoretical foundations are questionable [51 ] and the effort involved in preparing
three-dimensional plots, let alone using them, does not seem to be justified by
any increase in our understanding of extensional flows, or by any practical
application of the three-dimensional plots. Certainly it is true that a transient
extensional viscosity for a viscoelastic material should never be regarded as a
function of instantaneous rate of strain alone and the emphasis of that point
is a useful outcome of this discussion. It is doubtful whether, in general, one
more parameter can carry all the information about the state of the material (as
influenced by its flow history) that is needed to give a well-determined value
for the transient extensional viscosity.

The second set of results is more convincing, showing how results from the
"Rheotens" apparatus (in effect a melt spinning device) may be presented in a set
of "mastercurves" and "super-mastercurves" [52,53]. This presents a challenge
to the theorist to explain why such behaviour is observed - does it tell us
something about the family of experiments or about the appropriate constitutive
634

equation which we have not yet appreciated?

A final comment on experimental methods is perhaps worth making. While


the analysis of complex flows, such as converging flow or flow into a contraction,
does not give a reliable extensional viscosity in the sense of a fundamental ma-
terial property, such flows have a large component of extensional flow. Hence if
the Trouton ratio of a fluid is large, the stresses in the fluid will be predominantly
those associated with the extensional part of the rate of strain. A more prac-
tical point is that, if a consistent analysis can be made, measurements on such
flows should give data which can be used for reliable predictions of stresses for
similar flows in industrial processes. The test of a fluid property derived from
converging flow is then not so much "Does it give a true extensional viscosity?"
as "Does its use give reliable predictions in related practical flows?"

5.2. Stability of extensional flows


This topic has considerable relevance to polymer processing, where the predic-
tion of flow instability gives an understanding of limitations to production rates
for artefacts made from polymeric materials. There are a variety of instabili-
ties and failures in extensional flow which the author has discussed elsewhere
[54]. One recurrent theme is the need for clear distinction between different
instabilities such as "draw resonance" and the classical instability of a filament
due to capillarity. Similar clarity of thought and of description is needed when
filament rupture is considered [55]. Instabilities in extrusion ("melt fracture",
"sharkskin") are not immediately associated with extensional flow, but there are
strong links according to some of the mechanisms for these two distinct flow
defects [46,56].

The topic of approximations for nearly extensional flows [3,4,57] has several
connections both with stability and with the analysis of industrial processes like
fibre spinning and film manufacture, both by tubular film blowing and by film
casting. It is therefore not merely a topic of interest to theorists as a matter of
scientific curiosity, but a topic which could shed much light on our analysis of
a number of industrial processes.

REFERENCES

1. F.T. Trouton, Proc. Roy. Soc., A77 (1906) 426-440.


2. J.M. Dealy, J. Rheol., 38 (1994) 179-191.
3. C.J.S. Petrie, Elongational Flows, Pitman, London, 1979.
4. C.J.S. Petrie, Rheol. Acta, 34 (1995) 12-26.
5. J. Meissner, S.E. Stephenson, A. Demarmels and P. Portmann, J. Non-
Newtonian Fluid Mech., 11 (1982) 221-237.
6. C.J.S. Petrie, J. Non-Newtonian Fluid Mech., 14 (1984)189-202.
7. C.J.S. Petrie, J. Non-Newtonian Fluid Mech., 34 (1990) 37-62.
635

o A. Demarmels and J. Meissner, Coll. Polym. Sci., 264 (1986) 829-846.


9. J. Meissner, Ann. Revs. Fluid Mech., 17 (1985) 45-64.
10. D.F. James and K. Waiters, A critical appraisal of available methods for the
measurement of extensional properties of mobile systems, in Techniques of
Rheological Measurement, Ed A.A. Collyer, Elsevier, New York, pp.33-53,
1994.
11. J.E. Matta and R.P. Tytus, J. Non-Newtonian Fluid Mech., 35 (1990) 215-
229.
12. T. S ridhar, Recent progress in the measurement of extensional properties of
polymer solutions, in IUTAM Symposium on Rheology and Computation-
Abstracts, Ed J.D. Atkinson, N. Phan-Thien and R.I. Tanner, University of
Sydney, Australia, pp.29-30, 1997.
13. T. Sridhar, V. Tirtaatmadja, D.A. Nguyen DA and R.K. Gupta, J. Non-
Newtonian Fluid Mech., 40 (1991) 271-280.
14. V. Tirtaatmadja and T. Sridhar, J. Rheol., 37 (1993) 1081-1102.
15. J. Meissner, Trans. Soc. Rheol., 16 (1972) 405-420.
16. H.M. Laun and H. Mtinstedt, Rheol. Acta, 17 (1978) 415-425.
17. J.J. Linster and J. Meissner, Polymer Bulletin, 16 (1986) 187-194.
18. A. Demarmels and J. Meissner, Rheol. Acta, 24 (1985) 253-259.
19. J.J. Linster and J. Meissner, Makromol. Chem., 190 (1989) 599-611.
20. L. Berger and J. Meissner, Rheol. Acta, 31 (1992) 63-74.
21. J. Meissner and J. Hostettler, Rheol. Acta, 33 (1994) 1-21.
22. H. Gramespacher and J. Meissner, J. Rheol., 41 (1997) 27-44.
23. N.V. Orr and T. Sridhar, J. Non-Newtonian Fluid Mech., 67 (1996) 77-103.
24. S.H. Spiegelberg and G.H. McKinley, J. Non-Newtonian Fluid Mech., 67
(1996) 49-76.
25. J. van Nieuwkoop and M.M.O. Muller von Czemicki, J. Non-Newtonian
Fluid Mech., 67 (1996) 105-123.
26. B. Ziille, J.J. Linster, J. Meissner and H.P. Htirlimann, J. Rheol., 31 (1987)
583-598.
27. T. Samurkas, R.G. Larson and J.M. Dealy, J. Rheol., 33 (1989) 559-578.
28. M.H. Wagner, J. Non-Newtonian Fluid Mech., 4 (1978) 39-55.
29. R.B. Bird, C.F. Curtiss, R.C. Armstrong and O. Hassager, Dynamics of
Polymeric Liquids, Volume 2: Kinetic Theory, (Second Edition), Wiley,
New York, 1987.
30. R.G. Larson, Constitutive Equations for Polymer Melts and Solutions, But-
terworths, Boston, 1988.
31. Y. Kwon and A.I. Leonov, J. Non-Newtonian Fluid Mech., 58 (1995) 25-46.
32. M. Simhambhatla and A.I. Leonov, Rheol. Acta, 34 (1995) 259-273.
33. A.I. Leonov, Simple constitutive equations for viscoelastic liquids: formu-
lations, analysis and comparison with data, in this volume, 1998.
34. C.J.S. Petrie, J. Non-Newtonian Fluid Mech., 5 (1979) 147-176.
35. M.H. Wagner, T. Raible and J. Meissner, Rheol. Acta, 18 (1979) 427-428.
636

36. M.H. Wagner and H.M. Laun, Rheol. Acta, 17 (1978) 138-148.
37. A.C. Papanastasiou, L.E. Scriven and C.W. Macosko, J. Rheol., 27 (1983)
387-410.
38. M.H. Wagner and A. Demarmels, J. Rheol., 34 (1990) 943-958.
39. K. Feigl, H.C. 0ttinger and J. Meissner, Rheol. Acta, 32 (1993) 438-446.
40. M.H. Wagner and S.E. Stephenson, J. Rheol., 23 (1979) 489-504.
41. M.D. Chilcott and J.M. Rallison, J. Non-Newtonian Fluid Mech., 29 (1988)
381-432.
42. J.M. Rallison and E.J. Hinch, J. Non-Newtonian Fluid Mech., 29 (1988)
37-55.
43. L.E. Wedgewood, D.N. Ostrov and R.B. Bird, J. Non-Newtonian Fluid
Mech., 40 (1991) 119-139.
44. R. Keunings, J. Non-Newtonian Fluid Mech., 68 (1997) 85-100.
45. C.JoS. Petrie, J. Non-Newtonian Fluid Mech., 54 (1994) 251-267.
46. S.J. Kurtz, Some key challenges in polymer processing technology, in
Recent Advances in Non-Newtonian Flows, AMD-Vol 153/PED-Vol 14I,
ASME, New York, pp. 1-13, 1992.
47. C.J.S. Petrie, Recent ideas in extensional rheology, in Polymer Process-
ing Society Europe/Africa Regional Meeting - Extended Abstracts, Ed J.
Becker, Chalmers University of Technology, Gothenburg, Sweden, KN 4:2,
1997.
48. J. Ferguson and N.E. Hudson, European Polym. J., 29 (1993) 141-147.
49. J. Ferguson and N.E. Hudson, J. Non-Newtonian Fluid Mech., 52 (1994)
121-135.
50. J. Ferguson, N.E. Hudson and M.A. Odriozola, J. Non-Newtonian Fluid
Mech., 68 (1997) 241-257.
51. C.J.S. Petrie, J. Non-Newtonian Fluid Mech., 70 (1997) 205-218.
52. M.H. Wagner, V. Schulze and A. G6ttfert, Polym. Eng. Sci., 36 (1996)
925-935.
53. M.H. Wagner, B. Collignon and J. Verbeke, Rheol. Acta, 35 (1996) 117-126.
54. C.J.S. Petrie, Prog. Trends Rheol., II (1988) 9-14.
55. A.Ya. Malkin and C.J.S. Petrie, J. Rheol., 41 (1997) 1-25.
56. C.J.S. Petrie and M.M. Denn, AIChE J., 22 (1976) 209-236.
57. C.J.S. Petrie, Predominantly extensional flows, in 2nd Pacific Rim Confer-
ence on Rheology - Abstracts, Ed C. Tiu, P.H.T. Uhlerr, Y.L. Yeow and
R.J. Binnington, University of Melboume, Australia, pp.225-226, 1997.
637

MECHANICS OF ELECTRORHEOLOGICAL MATERIALS

K.R. Rajagopal

Department of Mechanical Engineering


Texas A &M University
College Station, Texas 77802

1. INTRODUCTION

Electrical and magnetic fields can significantly change the response characteristics
of many materials and electrorheology is the name given to the branch of mechanics
that is concerned with the flow of materials that are primarily affected by the action
of electrical fields. Usually, electrorheological materials are dielectrics or
semi-conductors in a non-conducting fluid, though recently Ferroelectrics have also
been used. Winslow's study (see [1 ]) of non aqueous silica suspensions under the
action of electrical fields seems to have been the first systematic analysis in
electrorheology, though the effect of an electrical field on the viscosity of pure liquids
was studied much earlier by Konig [2], Quinke [3] and Duff [4]. The work of
Winslow has been followed by a great deal of work in the field, and much of this
effort has been directed in fashioning such materials with a view towards producing
a better fluid in virtue of the potential applications for such materials in shock
absorbers, exercise equipment, valves, actuators and the like. However, for a variety
of reasons such applications have not met the perceived potential for such materials.
Initial attempts at manufacturing electrorheological materials were hampered by a
lack of understanding of the role of water in such suspensions. Other stumbling
blocks that have prevented the development of technological devices are the limited
operational temperature range, the abrasive properties of the suspension that erode
the devices that they flow through, the attrition of the particles, the stability of the
suspension and the enormous voltage requirements that are necessary to produce the
changes that are required. Much progress has been made to overcome these
limitations. Polymer based particles that mitigate the problem have been developed
638

(see Bloodworth [5]), and stabilizers have been found that increase the structural
stability of the suspension. Also, great strides have been made recently in decreasing
the voltage requirement.

A detailed account of the material science aspects of electrorheological fluids


is discussed in detail in the review article by Zukoski [6], and an assessment of the
technical applications of ER materials can be found in Krieger and Collins [7]. The
reader is also referred to Deinega and Vinogradov [8] for a review of
electrorheological materials. Here, we shall be primarily concerned with the
mathematical modelling of electrorheological materials.

Electrorheological fluids can be modeled starting at a microscopic level or within


the context of continuum mechanics in a homogenized sense. Here, we shall restrict
ourselves to continuum models; but even within the context of continuum mechanics
thee are several ways to modeling electrorheological fluids. One approach is to treat
the electrorheological fluid as a homogenized single constituent (see Atkin, Shi and
Bullogh [9], Rajagopal and Wineman [10], Wineman and Rajagopal [11 ]). Another
is to model it as a mixture (see Atkin and Craine [12], Bowen [13], Truesdell [14])
of a particulate medium and a fluid, each being treated as a single continuum (see
Yalamanchili, Rajagopal and Wineman [ 15]) allowing for interactions between the
two constituents. Here, we shall restrict ourselves to modeling the electrorheological
suspension as a single continuum.

Much of the modeling of the flows of electrorheological fluids have been restricted
to one-dimension, though there has been some work on three dimensional models.
Numerous three dimensional models can collapse to the same one dimensional model
and at the present moment there is not a sufficient body of experimental evidence in
general three dimensional flows which can be used to validate and select any one of
these models as the one that is best suited. In view of this, we shall restrict our
discussion to a reasonably general class of models. Experimental evidence suggests
that electrorheological fluids thicken significantly on the application of the electrical
field, respond in a Bingham like fashion with the yield depending on the applied field,
develop normal stress differences and stress relax (see Gamota and Filisko [ 16], Yen
and Achom [17], Gamota, Wineman and Filisko [18], Jordan and Shaw [ 19], Jordan,
Shaw and McLeish [20]). The response of electrorheological materials is also
significantly affected by the thermal conditions (Conrad, Sprecher, Choi and Chen
[21 ], Jordan and Shaw [22]). Here, we shall not consider thermal effects but restrict
ourselves to an isothermal analysis.
639

The fiber like structures that are formed on the application of the electrical field
suggest that such fluids ought to be modeled as anisotropic fluids. This leads to an
additional level of complexity which should be introduced after a better
understanding of such materials is achieved. However, this aspect of the modelling
is crucial and has to be reckoned with if one is to capture the behavior of
electrorheological fluids. For instance, the perceived viscoelasticity of the fluid that
is characterized by a time constant for the response could be due to the response time
associated with the alignment of the particles on the application of the field. The
electrorheological response of liquid crystals have also been studied (Carlsson and
Skarp [23], Yang and Shine [24]) where the material is anisotropic even before the
application of the electric field. We shall not consider these issues here, suffice it is
to say that the new framework that has been developed recently that allows for
variations in the synunetry of the body with various configurations that are natural to
the body can be used successfully to model such materials (see Rajagopal [25]).
Such an approach has been used to model crystal plasticity (Rajagopal and Srinivasa
[26]), twinning and solid to solid phase transition (Rajagopal and Srinivasa [27],
[28]), multi-network theory for polymers (Rajagopal and Wineman [29]) and
anisotropic viscoelastic fluids (Rajagopal and Srinivasa [30]).

2. K I N E M A T I C S AND BALANCE LAWS

Let f2 denote the reference configuration of a body B. By the motion of a body we


mean a one-to-one mapping ~ that assigns to each point Xc f] a point x belonging
to a three dimensional Euclidean space, at each instant of time t, i.e.,
x=x(X,O. (2.1)

The image of f] under X, denoted as ~"~t, is the configuration occupied by the body at
time t. We shall assume that X is sufficiently smooth to render all the following
operations meaningful. The velocity v and acceleration a are defined through

1Various properties ~ associated with a material point at different instants of time can be

defined through d~=~(X,t)=~(x,O. We denote d~ _ 0~ a~ _ O~,v~= 0__~and grad


dt a t ' a t at ox

r Also, div ~ denotes tr [ grad ~] and Div ~ denotes tr [V ~].


ax
640

O~ dx
v- - , (2.2)
Ot dt
and

a - 0 2 X - ~d 2.x (2.3)
Ot 2 dt 2

The deformation gradient F and the velocity gradient L are given respectively
through

F: OX:Vx, (2.4)
OX

and

L - d_~v: grad v. (2.5)


dx
The symmetric and skew part of L are denoted by D and W, respectively, i.e.,

D:~(L +Lr),W:I(L-L r). (2.6)

We shall keep our kinematical definitions and the documentation of the basic
equations to a minimum while ensuring that the treatment be self-contained. A
complete and proper flame-work for the study of electrorheological fluids would
require the laws of electromagnetism in addition to the usual laws of
thermomechanics. There are many ways of expressing the equations of
electromagnetics and we shall use the Minkowskian formulation. A detailed
discussion of the basic laws of field dependant materials within the context of
continuum mechanics can be found in Truesdell and Toupin [31] and Eringen and
Maugin [32]. We shall use the dipole-current loop model (see Pao [33]).
641

The conservation of mass is given by 2

0p +div (pv)=0, (2.7)


Ot
where p is the density. The balance of linear momentum takes the form

divTr+pf+f~=P dt" (2.s)

where T is the Cauchy stress 3, f is the external mechanical body force, and
fo the electromagnetic force density given by

f e : : q E + l j x B +10P xB +l div[(PxB)~v] +[grad B] r M+[grad E]P; (2.9)


c c Ot C

qo is the electric charge density, E is the electric field, J the conduction current, B the
magnetic flux, P the electric polarization, and M is defined through

M=M +lvxP (2.10)


c

where M is the magnetic polarization.

The balance of angular momentum takes the form

dv
div(x T) +xx pf+Q = x x p (2.11)
dt'
where ~o is the electromagnetic angular momentum density given through

2 A documentation of the governing equations can be found in Rajagopal and Ruzicka [34]

3While the phenomenological continuum models assume that the stress is symmetric, some
models based on particle dynamics lead to expressions for the stress that are not symmetric. We
shall assume here that the stress is symmetric. A discussion of the asymmetry of stresses of
electrorheological materials can be found in Rosensweig [35]
642

~e" =xxf e +Px~+MxB, (2.12)

where g" is the electromotive force intensity given through

g'=E+lvxB. (2.13)
c

The balance of energy takes the form

pd(e+-~lvl
1 2) +div q -div(Tv) +Of-V+pr+We, (2.14)

where e is the specific internal energy, q the heat flux vector, r the radiant heating
and w~ is the energy production density given by

where J is given through


J - J - q~v. (2.16)

Even in classical continuum thermomechanics, the specific formulation of the


second law of thermodynamics is an object of much contention. Thus, the exact form
of the second law in electromagnetics is far from settled. Here, we record the second
law in the form of the Clausius-Duhem inequality, though alternate interpretations of
the second law are possible:

p drl
dt +div (O)+p -~>_0,
r (2.17)

where 11 is the entropy and 0 the absolute temperature. The Clausius-Duhem


inequality places restrictions on the allowable forms of the constitutive expressions.

The above balance laws are the usual laws of thermomechanics modified to account
643

for the effects of the electrical and magnetic fields. We have to augment the above
equations with Maxwell's equations in the Minkowskian form. The conservation of
electric charge is

Oqe +div J=0. (2.18)


0t

Gauss' law is given by

(2.19)
div I)e=qe ,

where D~ is the electric displacement field given through


De=P+E. (2.20)

Faraday's law is given by

1 0B
curl E +-- =0. (2.21)
c Ot

The conservation of magnetic flux is given by

div B=0, (2.22)

and Ampere's law takes the form

1 ODe 1
curl H- + (2.23)
c Ot c

where H is defined through

H: =B-M. (2.24)
644

The system of equations (2.18) - (2.24) can be manipulated to obtain other


equations that can prove more amenable to use. We shall not get into this here but
refer the reader to [36] for a discussion of the same. While the equations of
thermomechanics are invariant under Galilean transformations, Maxwell's equations
are invariant under Lorentz transformations, and Galilean transformations are not
uniform approximations of Lorentz transformation (see [37]).

In general, to solve flow problems involving electrorheological fluids, it would be


daunting to use the full system of equations (2.7) - (2.24) except in the simplest of
problems. Thus, we need to simplify the system of equations.

A gross simplification is to ignore Maxwell's equations, as well as effects due to the


fields in the thermomechanical equations as field variables but treat the electric field
as a parameter. Much, though not all, of the modeling in electrorheology is in this
spirit. We shall discuss such an approach in some detail later. A less drastic
simplification is that based on the fact that the fluid is nonconducting and that we are
dealing with a dielectric, i.e.,

J=O, (2.25)
and

M=O. (2.26)

It follows that (see Rajagopal and Ruzicka [34])

d9 +9 div v=0, (2.27)


dt

dv
div T+pf+f = p ~ (2.28)
dt

de -kA0 =T.L + dP "g'+(P-g0div v, (2.29)


~ dt
645

T+O 2 1 "L+ >_0, (2.30)


0

div(E +P) =qe, (2.31)

1 aB
curl E + - - ~ = 0 (2.32)
c at

divB=0, (2.33)

curl B+lcurl(v 1a 1 (2.34)


c = c - ~ (E +P) ---qeV'c

dqe
~+qe div v=0. (2.35)
dt

In the above equations we have assumed that the heat flux is given through Fourier's
law and ~ is the specific Helmholtz potential. The above equations can be further
simplified on the basis of dimensional arguments for problems of interest in
electrorheology. A detailed treatment of the same can be found in [36].

In general we will have to solve the coupled partial differential equations (2.27)-
(2.35) which is tantamount to fourteen partial differential equations and one
constraint inequality for the appropriate variables, a most arduous task even under
idealized conditions. Since in most problems of practical relevance, the gaps are
exceedingly small, it may be reasonable to assume the electric field to be a constant
(of course, this assumption could be totally inappropriate as the electric field could
vary tremendously within the short gap). Such an assumption is often made in
electrorheology and the electric field then plays the role of a parameter in the
problem. We shall further simplify the analysis by restricting ourselves to isothermal
processes and essentially ignoring the thermal variables. Even such a simplified
situation serves to highlight certain interesting features concerning the flows of
electrorheological fluids.
646

It is possible that there are some electrorheological fluids that undergo only isochoric
motions when the electric field E is held a constant, however the density changing
with the electric field. This situation is similar to fluids which can undergo isochoric
flows in isothermal processes while their densities can change with temperature.
Such an assumption is the starting point for the celebrated Oberbeck-Boussinesq
approximations in fluid mechanics. The fact that the density can change with the
electric field, but is a constant in all processes in which the electric field is a constant
can be expressed through

det F =f(E).

A detail discussion of the consequence of the above constraint can be found in


Rajagopal and Ruzicka [36].

3. S I M P L I F I E D M O D E L S B A S E D O N T H E E L E C T R I C F I E L D AS
A PARAMETER: DIFFERENTIAL TYPE MODELS

We shall restrict our analysis to an incompressible electrorheological fluid. Further,


we shall assume that the Cauchy stress is given through

T - -pl f(D,e), (3.1)

where -pl is the indeterminate part of the stress due to the constraint of
incompressibility and D is the stretching tensor defined through (2.6). We shall
assume that the material is isotropic. The response of materials of the form (3.1) has
been studied in some detail by Rajagopal and Wineman [29].

It follows from flame-indifference and isotropy that f must satisfy

f(QDQr QE)=Qf(D,E)Q r VQr (3.2)

where O is the orthogonal group. It follows from standard representation theorems


that T can be expressed as (see Spencer [38]):
647

T= -pl +aIE~)E+a2D+a3D2+a4(DE~)E+E~)DE)
(3.3)
+as(D2E| +E|

where a~, I=l, ...,5 are scalar functions that depend on the following invariants

Ii:tr (E@E),/2 =trD2,/3 :tr(DE|


(3.4)
I4=tr(D 2E|

On the other hand if we require that (3.2) holds for proper orthogonal
transformations O +, i.e., invariance under rigid body motions, then T has the
representation
T : -pl +&IE~E +tx2D+~3D2 +t~4(DE| +E|
+6cs(D2E| +E| +&6(MD +DM r) +&7(MD2+D2M 7,) (3.5)
+&8(DMD2+D2M rD) +&9(E(~)MA+MA| +&lo(E| +MB|

where A, B and M are defined through

A=DE, B--DE and M=cE, (3.6)

where e is the alternator tensor and the &; , i=l,... 10 are functions of the invariants

I 1=tr(E | =trD 2,I3: tr(D E | trD3,


(3.7)
I s =tr(D2E@E),/6 =tr[D2M TA| rD]I 1.

Invariance under O or O are both assumptions, whichever choice is made. We note


that even in the simpler case corresponding to invariance under the full orthogonal
group, where temperature effects are ignored, and the electric field is treated as a
parameter, there are five arbitrary material functions that appear in the representation,
and in general it would be difficult to devise a reasonable experimental protocol to
648

determine these material functions. In order to illustrate some interesting interactions


between the electric field and the deformation, we shall simplify the model further.
However, we shall see even in this simplified model, the presence of normal stress
differences can be induced in simple shear flows due to the applied electric field.
Also, the fluid can thicken, i.e., the viscosity can increase considerably with the
electric field as is to be expected if the model is to describe the response of real
electrorheological fluids.

3.2 Simple Shear Flow

Here, we shall essentially outline the analysis of Wineman and Rajagopal [29] to
illustrate that the model (3.3) can describe normal stress differences, thickening and
other behavior characteristic of electrorheological fluids. Let us consider the simple
shear flow of an electrorheological fluid modeled by (3.3) with an electric field
applied transverse to the direction of flow, i.e.,

v =u(y)i, E=E2j +E3k (3.8)


A straightforward calculation yields (see Rajagopal and Wineman [29])

T11_ _p+_~ ~2

T22--p+ot2Ef+-~IX4 ~2 +_~ y2E2,2


T33= -p +o~2E2,
(3.9)
+~
T12:(-~ T 2)Y,

~5
TI3 =--~-YE2E3,

T23=o~2E2E3+~.~6]t2E2E3.

where
649

Y:u'0') (3.10)

First, suppose that E3=0. We notice that the shear stress T12 can be expressed as

T12: [itt(u 2)]y, (3.1 1)

where the generalized viscosity ~t is dependent on both the shear rate and the electric
field. Thus, for appropriate forms of ~t, the model can describe the increase in
viscosity that is observed in electrorheological fluids due to the application of an
electric field. Depending on the nature of the fluid, it can shear thin or shear thicken
and this can either enhance or ameliorate the thickening due to the electric field. It
is also worth noting that in general the normal stresses Tll , T22 and T33 would be
different and thus normal stress differences that are characteristics of non-Newtonian
fluids are also induced by the electric field. We note that

r, 1 -TY 2,

tX4- 2_~2E32, (3.12)


T11-T33=TY

T22-T33=o 2(E2-E2) y2 + Ix6

and thus in general the normal stress differences are distinct. We also notice that
there is a contribution to the normal stress differences due to the sheafing as well as
the electrical field. Moreover, we recognize that there is a coupling effect between
the mechanical and electrical fields that arises in the normal stress differences Tll-T22
and T22- T33.

Next, we shall illustrate the effect of the electrical field in a simple flow.

3.3 Flow Between Parallel Plates

Consider an electrorheological fluid modeled by (3) flowing between two infinite


650

parallel plates along the x-axis due to an applied pressure gradient with the electric
field applied along the y-axis.

It follows from the balance of linear momenttLm, in the absence of body force fields,

To,=-Cy, (3.13)

where C . . . .Op-constant > 0 . For the problem under consideration substituting


Ox
(3.13) into (3.11 ) leads to

[g(u',E2)lu'=Cy, (3.14)

and the above equation is solved subject to the boundary conditions

u(-h)=0, u(h)=0. (3.15)

It has been observed that in the presence of an electric field the fluid flows only
after a critical value is reached for the shear stress, i.e., it exhibits "Bingham type"
behavior. Of course, it is possible that the fluid flows ever so slowly even below the
"yield stress" in that its flow is imperceptible within the time frame of the
observation. To simplify the problem, we shall assume a Bingham type behavior.
Thus, we shall consider a representation for Txy of the form

Oo(E)+v(E)v,
T= 0, y=0 (3.16)

with the assumption that Oo(0)=0.


651

We could assume more complicated responses wherein la depends on both y and


E, i.e. ~t= la(y,E).

For a fixed value of E, the maximum value of the magnitude of the shear stress
occurs at y= and its value is Ch. If Ch < Oo(E), then the maximum value of the
stress is less than the yield stress and there will be no flow. Let C be such that Ch
> Oo(E). Then, there is some y* such that Cy *= Oo(E). It then follows that

- Oo(E)=~t(E)u '- -Ay, y* <_y<h


u '-0, y * <_y<_y (3.17)
Oo(E)+~(E)u '- -Ay -h<_y-<y*.

It immediately follows from (3.17) and the no-slip boundary condition (3.15) that

l C 2 2)
u(Y)= t(E)[~(h -Y -Oo(E)(h-y)] , y*_<y<h

uO') =constant -y "<y<y * (3.18)


1 c
u(Y) =g ~ [ ~ ( h 2-y)z-oo(E)(h +y)], -h_<y<-y *

We notice that in the absence of the electric field, Oo(E)=0 and thus y*(E)=0, and
we obtain the classical Poiseuille solution

u0,)_ _~(h 2_y 2),-h <_y<_h. (3.19)

Other boundary value problems within the context of this theory can be found in
[29]. It is also shown there that a Mooney-Rabinowitch type of relationship can be
established for such models.

Models of the class discussed above do not exhibit viscoelasticity, and it has been
observed that certain electrorheological fluids exhibit viscoelastic response (Xu and
Liang [39], Thurston and Gaertner [40], McLeish, Gordon and Shaw [41]
Vinogradov et.al [42]). In the next section we discuss models that can describe the
viscoelastic behavior of electrorheological fluids.
652

4. Simplified Models Based On The Electric Field As A Parameter: Integral


Models

We shall now discuss a class of models that have a fluid like response in a certain
regime and a solid like response otherwise.

We shall yet treat the electric field as a parameter and we shall use a transition
function A(T(t), E(t)) which delineates the regimes of response"

(i) A(T(t), E(t))=0 determines the transition (boundary) between solid-like and
fluid-like response;

(ii) A(T(t), E(t)) < 0 the material behaves like a solid;

(iii) A(T(t), E(t)) >0 the material behaves like a fluid.

We shall not get into a detailed discussion of such models here, the interested reader
can find a detailed treatment in [11 ]. In the solid-like regime, the stress has the form

to
(4.1)
T~olid(t)=.~'-[Fto(s),E(t)]
s=t o

where F (s) denotes the deformation gradient at time s with respect to the
configuration at time to when the transition function is met and the material starts to
exhibit solid-like behavior. Let us suppose that the material when it is solid-like is
isotropic. It follows from standard arguments the T (t) has the form

Tsolid(t)- ~ - [Bto(t),Ct(s);E(t)] ' (4.2)


s-t o

where Bto(t):Fto(t) Ftro(t), and Ct(s):Frt(s)Ft(s).


Next suppose in the time interval [tl, t2] the material exhibits fluid-like behavior.
653

Then,

Tfluid(t) =~'-[~-t(s),tl;E(t)] ' (4.3)


tI

and from standard arguments it follows that

Tfluid(t) =,~ [Ct(s),t, ;E(t)]. (4.4)


tI

It is possible to define response functionals which can have a more general structure
(see Wineman and Rajagopal [11 ]):
co

T(t)= G [ F ( t - s ) , E ( t - s ) ] , (4.5)
s---O

which in the absence of the history of the electric field reduces to the definition of a
simple material (see Noll [43], Truesdell and Noll [44]). The above models are too
general to be of use. Thus, we shall turn our attention to a specific model which can
describe the viscoelastic behavior of electrorheological materials. Consider the
following representations for the solid-like and fluid-like part of the response (see
Wineman and Rajagopal[ 11]):
654

Tsolid- -pl +[~IE| +[~213+[~3]~2+[~4D+[~5(l~E| +E|


+[~6(132Et~)E+E| [~7(DE@E +E|
t
+f Ll(t-slCt(slds
0

t t

+f f L2(t-Sl't-s2)[Ct(Sl )1~ t(S2 ) +(~ t(s2)C_,t(Sl)]dSldS2


0 0
t
+f L3(t-s)[C t(s)E| +E| t(s)Elds (4.6)
0
t t
+f f L4(t-s,,t-s2)[Ct(S1){~ t(s2)EI~E + E ~ t(S1)(~ t(s2)E]dsldS2
0 0
t

+f Ls(t-s)[f3(tl(~ ,(s)+(~ ,(s)B(tllds


0

+f L6(t-s)[B(t)C t(s)E@E +E@]3(t){~ t(s)E]ds,


0

and for the time t > tl when the response is fluid-like


655
t

Tnuid = -pl +y~E| +f Ml(t-s)Ct(s)ds


t1

t f

+: :M2(t-Sl,t-s2)[~'t(Sl)C-~t(s2)+f-,,(~2)C-,t(Sl)]dSldS2
tl t1
t (4.7)
+f M3(t-s)[Ct(s)E| +E|
t1

t t

+ff t(s2)E~)E+E~)(~t(s,)Ct(s2)]Edsids2,
t1 tI

where I~:B t (t)-I and Ct(s):Ct(s)-I . The scalars 13i,I= 1,--7, and Yi, I=1, 2 are
constants. Also, Li, i--1, - - 6 and M., i-1, --4 satisfy certain continuity and
monotonicity properties.

The model outlined above is too general to be useful and it is unlikely that an
experimental protocol can be devised where all the arbitrary material functions can
be determined. The purpose of documenting the model is to merely show general,
principles of continuum mechanics can be used to develop general enough models
can describe the response characteristics of electrorheological fluids. Of course, the
above model could be further simplified with the multiple integral terms ignored
which would then lead to models similar to those used in viscoelasticity to describe
the time dependant behavior of such materials. For instance, there has been
considerable interest in the response of electrorheological materials when subject to
oscillatory shears (see McLeish, Jordan and Shaw [42], Gamota and Filisko [16],
Otsubo, Sekine and Katayama [45]). Wineman and Rajagopal [11] use the above
model to study small simple shear and oscillatory shear in electrorheological fluids.
In the case of the oscillatory shear problem the stress oscillates with the frequency
to of the oscillatory shear, and G~ and G~ the storage and loss modulii of linear
viscoelasticity depend on the electric field and the phase angle depends on both the
frequency to and the electric field E. They are able to correlate and explain
phenomena observed in the oscillatory response of electrorheological fluids by
Gamota, Wineman and Filisko [16], and Gamota and Filisko [46], [47].
656

General three dimensional continnum models that can describe the response of
electrorheological materials such as thickening, stress-relaxation, normal stress
differences, yield etc., have been presented. The simple shear flow of such fluids
has been analysed and even in such a simple flow it is found that the models are
capable of describing the characteristics exhibited by real electrorheological
materials such as thickening due to the applied electric field, normal stress
differences and stress relaxation.
657

BIBLIOGRAPHY
1. W.M. Winslow, 1949, J. Applied Physics, 20:1137.
2. W. Konig, 1985, Ann. Phys., 25:618.
3. G. Quinke, 1897, Ann. Phys., 62: 1.
4. A.W. Duff, 1896, Phys. Rev., 4: 23.
5. R. Bloodworth, 1995, Electrorheological Fluids Based on Polyurethane
Dispersions, Bayer Corporation Report.
6. C.F. Zukoski, 1993, Material properties and the electrorheological response,
Annual Reviews in Material Science, 23: 45.
7. I. Krieger and E.A. Collins, 1992, Electrorheological fluids, A Research
Needs Assessment, Washington, DC: US Dept. Energy, Office Energy Res.
Program Analysis.
8. Y.F.Deinega and G.V. Vinogradov, 1984, Rheol. Acta, 23: 636.
9. R.J. Atkin, X. Shi and W.A. Bullogh, 1991, Journal of Rheology, 35" 1441.
10. K.R. Rajagopal, R.C. Yalamanchili and A.S. Wineman, 1994, International
Journal of Engineering Science, 32" 481.
11. A.S. Wineman and K.R. Rajagopal, 1995, On Constitutive Equations for
Electrorheological Materials, Continuum Mechanics and Thermodynamics,
7: 1.
12. R.J. Atkin and R.E. Crame, 1976, Q.J. Mech. Appl. Mathematics, 17: 209.
13. R.M. Bowen, 1976, Theory of Mixtures in Continuum Physics Vol. 3., ed.
A.C. Eringen, Academic Press.
14. C. Truesdell, 1984, Rational Thermodynamics, Springer-Verlag.
15. K.R. Rajagopal, R.C. Yalamanchili and A.S. Wineman, 1994, International
Journal of Engineering Science, 32" 481.
16. D.R. Gamota, A.S. Wineman and F.E. Filisko, 1991, Journal of Rheology,
35- 399.
17. W.S. Yen and P.J. Achorn, 1991, Journal of Rheology, 35: 1375.
18. T.C. Halsey, J.E. Martin and D. Adolf, 1992, Phys. Rev. Letters, 68" 1519.
19. T.C. Jordan and M.T. Shaw, 1991, Electrorheology, MRS Bulletin, 38.
20. T.C.Jordan, M.T. Shaw, and T.C.B. McLeish, 1992, Journal of Rheology,
36: 441.
21. H. Conrad, A.F. Srecher, Y. Choi and Y. Chen, 1991, Journal of Rheology,
35: 1393.
22. T.C. Jordan, M.T. Shaw, and T.C.B. McLeish, 1992, Journal of Rheology,
36: 441.
23. T.Carlsson, K. Skarp, 1981,Mol. Cryst. Liq. Cryst., 78: 157.
24. I. Yang and D. Shine, 1992,Journal of Rheology, 36: 1079.
658

25. K.R. Rajagopal, 1995, Constitutive Relations and Material Modeling, Univ.
Of Pittsburgh Report.
26. K. R. Rajagopal and A. Srinivasa, Mechanics of the inelastic behavior of
Materials, Parts I and II, to appear in International Journal of Plasticity.
27. K.R. Rajagopal and A. Srinivasa, 1995, International Journal of Plasticity,
Vol. II: 653.
28. K.R. Rajagopal and A. Srinivasa, 1997, International Journal of Plasticity,
Vol. 13: 1.
29. K.R. Rajagopal and A. S. Wineman, 1990, Archives of Mechanics, Vol. 42:
53.
30. K.R. Rajagopal and A. Srinivasa, Viscoelastic Anisotropic Fluids, in
preparation.
31. C. Truesdell and R. Toupin, 1960, The Classical Field Theories of Mechanics,
Handbuch der Physik, Vol. 3, Springer-Verlag, Berlin.
32. A.C. Eringen and G. Maugin, 1989, Electrodynamics of Continua, Vols. I and
II, Springer, New York.
33. Y.H. Pao, 1978, Electromagnetic Forces in Deformable Continua, Mechanics
today (New York) (S. Nemat-Nasser, ed.), Pergamon Press, 209.
34. K.R. Rajagopal and M. Ruzicka, 1996, Mechanics Research Communications,
23: 401-407.
35. E. Rosensweig, 1992, Magnetic Fluids, Cambridge University Press.
36. K.R. Rajagopal and M. Ruzicka, Mathematical Modeling of
Electrorheological Materials, submitted for publication.
37. J.M. Levy-Leblond, 1977, Rivista del Nuovo Cimento, 7:187-214.
38. A.J.M. Spencer, 1995, Theory of Invariants, Continuum Physics (New York)
(A.C. Eringen, ed.), Vol. 3, Academic Press.
39. Y.Z. Xu and R.F. Liang, 1991, Journal ofRheology, Vol. 35:1355.
40. G. B. Thurston and E.B. Gaetner, 1991, Journal of Rheology, Vol. 3: 1327.
41. T.C.B. McLeish, T.C. Jordan and M.T. Shaw, 1991, Journal of Rheology,
35: 427.
42. G. V. Vinogradov, P. Shulman, Yu. G. Yanovskii, U.V. Barancheeka, E.V.
Korobko and I.V. Bukovich, 1986, Inzh.-Fiz, Zh., 50: 605.
43. W. Noll, 1972, Arch. Rational Mechanics and Analysis, Vol. 48: 1.
44. C. Tmesdell and W. Noll, 1992, The Non-Linear Field Theories of Mechanics,
Springer-Verlag, Berlin.
45. Y. Otsubo, M. Sekine and S. Katayama, 1992, Journal of Rheology, 36: 479.
46. D.R. Gamota and F.E. Filisko, 1991,Journal of Rheology, 35: 399.
47. D.R. Gamota and F.E. Filisko, 1991, Journal of Rheology, 35: 399.
659

CONSTITUTIVE EQUATIONS FOR ELECTRORHEOLOGI-


CAL FLUIDS BASED ON MOLECULAR DYNAMICS

R. Tao

Department of Physics, Southern Illinois University at Carbondale, Car-


bondale, IL62901, USA

1. INTRODUCTION

Electrorheological (ER) fluids are a class of materials whose theolog-


ical characteristics are controllable through the application of an electric
field. A typical ER fluid consists of colloidal dispersions of dielectric
particles in a liquid of low dielectric constant. When an electric field is
applied, the effective viscosity of the ER fluid increases dramatically. If
the field exceeds a critical value, the ER fluid turns into a solid whose
shear stress continues to increase as the field is further strengthened [1-5].
The above novel properties are the result of a structure change in
ER fluids under an electric field. It is known that upon application of
an electric field the dielectric particles in ER fluids form chains spanning
between the electrodes. The chain will aggregate to form thick columns.
The body-centered tetragonal (bct) lattice was predicted theoretically as
the ground state structure for an ER fluid and has been verified exper-
imentally [4,7]. The microstructure of ER fluids is a fundamental issue.
The viscosity increase and solidification of ER fluids are all related to this
microstructure. The mechanical and physical properties of ER fluids also
strongly depend on the induced structure [8]. Recently, the ER effect has
also been used to produce new composite materials [9,10]. Therefore, in-
formation about the microstructure of ER fluids under various conditions
is very important for the growing number of applications.
In this paper, we will start from the first physics principle to derive
constitute equations for ER fluids based on molecular dynamics. Then we
will apply computer simulations to investigate the structure of ER fluids.
Our ER system is confined between two electrodes located at z = 0 and
z = L respectively, upon which a voltage may be applied. The system
consists of spherical dielectric particles of diameter a and dielectric con-
stant ep suspended in a nonconducting liquid. The liquid has a dielectric
constant of e / a n d viscosity r/. In an electric field, each particle has an in-
660

duced dipole moment of p - o~ef(a/2)3Eto~ where a - ( e p - e l ) / ( e p + 2el)


and Eto~ is the local field. The formation of stru&ure is driven by dipo-
lar interactions, viscous drag forces, and Brownian motions. Molecular
simulations on this model have shown that the ER fluid can readily form
the bct lattice under a strong electric field [11]. Several other computer
simulations on a similar model have also been reported [12-16]. In our
extensive simulations, we will examine the detail of the induced ER struc-
ture under various conditions.
Our results indicate that ER fluids under an electric field may de-
velop into five different structures under different conditions. In a weak
electric field, ER fluids can move from a liquid state to a nematic liquid
crystal state which only has ordering in the field direction, but no order-
ing in other directions [17]. This ordering indicates chain formation along
the field direction. In both liquid and liquid crystal states, the columnar
particle density remains uniform. When the electric field is strong and
the thermal fluctuation is weak or moderate, ER fluids develop into a
body-centered tetragonal (bct) lattice. In a very strong field without or
with very weak thermal fluctuations, ER fluids may develop into a poly-
crystalline structure consisting of many small bct lattice grains. If both
the electric field and thermal fluctuation are strong, ER fluids develop
into a glass-like structure in which the particles are aggregated together
to form thick columns, but the structure has no appreciable ordering. In
all of the last three structures, the columnar particle density peaks in a
small region, indicating a thick-column structure.
Our simulation also provides information about the solidification
and the chain formation time in ER fluids. The chain formation time is
much shorter than the solidification time. Both of them depend mainly
on the ratio of the Reynolds number to the Mason number, especially
in the over-damped case. Here the Reynolds number is the ratio of the
inertial force to the viscous force while the Mason number is the ratio of
the viscous force to the dipolar force. However, when the viscosity is not
too strong, the ratio of the Brownian force to the dipolar force also plays
a role in these two time scales.

2. CONSTITUTIVE EQUATIONS

We use the Langevin Equation to describe the motion of particle i,

d2ri _ dri (1)


m--~- - F i - gTrarl-~+Ri(t ).

Here Fi includes all electric forces on particle i, 37rorlvi is the Stokes'


drag force, and Ri(t) is a Brownian force. Fi is given by the following
661

expression

. _.-~.[fiJ
Fi -- +fidrep ]+fi self +fi wall (2)
jr

where fij is the force acting on particle i by particle j and all of j's
images, firfp is a short-range repulsive force to prevent particles i and
j from overlapping, f~lf is the force on particle i due to all its own
images, and f~an is a short-range repulsive force to prevent particle i
from penetrating the two electrodes.
The dipolar force exerted by particle j on particle i is given by

3/) 2
(efr~j) [e~(1-3 cos 20ij)-eo sin 20ij] (3)

where rij - r i - rj and Oij is the angle between the z direction and the
joint line of the two dipoles. When a dipole is placed inside a capacitor at
r i - (xi, yi, zi), an infinite number of images are produced at (xi, yi,-zi)
and (xi, yi, 2Lk 4-zi)for k - -4-1, +2,--.. The force that the j t h particle
and its images exert on the ith particle is given by

_ p2 oo 4s 3r 3 ST~fli j STrZ~' STrZj


fij,, efL-----
4 y~ (xi -- Xj) f(1 ( L ) cos( L .....) cos( L )
s=l Pij ,
_ p2 ~ 8371_3 sTrPij,) 8~zi~ 8~zj,)
f ij,y ef L4 ~ 4 (Yi - Yj)1(1( L "cos( cos( (4)
s=l Pij "L " L "
p2 co
ef L~ y ~ 4s3 7r3I(o( STcPij sTrzi 87FZj
fij~z ~-" " L )Sin( L,, )cos( n )
s--1

where Pii - X/(xi - Xd)2 + (Yi yj)2 and Ko and Kt are modified Bessel
functions.
The force on a particle due to its own images is in the z-direction
and given by

f self3p2[1 ~176 1 1 ] (5)


i,z 8E f -- 7Z -}- y ~ (Zi -- 8 L ) 4 - (zi q- 8 L ) 4 "
s=l

To introduce hard spheres and hard walls into the simulation, we use a
662

short-range repulsion between two particles

,~p 3p2e, e x p [ _ l O O ( r o / a _ l ) ] (6)


fij ~- e_/a 4

and the short range repulsion between a particle and the electrodes as

fwall 3p2ez
-- e - ~ { e x p [ - l O O ( z i / a - O . 5 ) ] - e x p [ - l O O ( ( L - z i l / a - 0 . 5 ) ) } . (7)

The random force Ri(t) has a white-noise distribution,

(Ri.~) - O. (Ri..(O)Ri.~(t)) - 67rkBTa~6.a6(t) (8)

where kB is Boltzmann's constant and T is the temperature. We intro-


duce a subinterval T which is shorter than the time steps used in the
integration of Eq.(1) but much longer than the molecular collision time
1 [.t+rp.
[18]. The average of Ri(t) over T, Ri,~(t, T) -- -~a, - . , , ( t ' ) d t ' has a
Gaussian distribution
1
W(R,..(t..)) - (9)

where ~t - V / 6 r k B T a ~ / ' r . For a time step 6t > T, we can divide it


into many subintervals of duration T, in which all quantities except Ri(t)
can be treated as constants. Then, for any smooth function r X , =
ft ~+6~r has a probability distribution

W(X,) - (Trq)-'/2 e x p ( - X 2 / q )

where q 12~vkBTa~ (~)d~ and is independent of T. Although the


value of T is not very uniquely defined, Eq.(10) implies that our results
do not depend on a specific choice of T.
The intrinsic time scale in Eq.(1) is to - m / ( 3 r a ~ ) . We re-scale the
variables, t - tot*, F i - FoF* where F o - 3p2/(ela4), R i - ~R*(t), and
ri - ar* in Eq.(1). The scaling produces a dimensionless constitutive
equation

ri''* + ri ' * - A(F* + BR*) (11)


where A - Foto/(3rrla ) and B - gt/Fo. It is interesting to note that
A is the ratio of the Reynolds number to the Mason number. The
663

Reynolds number R is the ratio of the inertial force m v 2 / a t o the vis-


cous force 37raCy where v is the speed of dielectric particles. Hence
R = mv/(37rrla 2) = vto/a. The Mason number M n is the ratio of the
viscous force to the dipolar force, M n = 3raTlv/Fo. Then, it is clear that
A - R / M n . In our problem, dielectric particles have negligible speed
before the electric field is applied or after the solid structure is formed.
During the process, the particles have steady or typical speed, either. As
a result, there is no typical Reynolds number or Mason number in our
problem. However, their ratio, A, is independent of the particle speed
and hence a good parameter for our simulation.
We also note that parameter B is the ratio of the Brownian force
to the dipolar force. In addition, 1 / ( A B 2) - (1.5T/to))~ where A =
( p 2 / E l a 3 ) / k u T , a crucial parameter [19] in study of ER system in equi-
librium process [17]. However, we must remember that our electric-field
induced solidification is a non-equilibrium process, during which the tem-
perature is not uniform throughout the system.

3. MOLECULAR DYNAMICS SIMULATIONS

In our simulation we use 122 particles confined in a space of dimen-


sions L~ - Ly - 5a and L~ - 14a. This corresponds to a volume fraction
of r - 0.183. There are periodic boundary conditions in the x and y
directions. We probe the structure at each step using the following three
order parameters,

N
1
pj - [ ~ ~ exp(ibj- ri)[ (12)
i=1

where the bj are the reciprocal lattice vectors of the bct lattice

bl- (27cIo)(2etzlv~-ez), b 2 - (2~lo')(2etylV/-6-ez), b3 - 47re~/a. (13)

Of the three unit vectors, ez is along the field direction and e~,ey~ are
along the intrinsic axes of the bct lattice. In the calculation of the order
parameters, we must rotate the coordinate system around the z axis to
find the intrinsic axes of the structure which maximizes pip2. The order
parameter P3 characterizes the structure along the z direction while pl
and p2 characterize structure in the x-y plane.
In the simulation, we assume that an electric field is applied at t - 0
instantaneously. Then we apply an adaptive step-size control Runge-
Kutta method to integrate the equation of motion. It is clear from
664

Eq.(11) that the final structure depends on the two parameters A and B.
We examine the dynamic process by monitoring the particles' positions,
velocities, and the structure's order parameters. In most cases, after
application of the field, the particles quickly move to form chains, then
the chains aggregate to form a thick structure. Afterwards, the particles
usually fluctuate slightly around their positions in the structure and the
order parameters of the system change very little. Hence we are able to
determine the solidification time and analyze the final structure. The
other interesting quantity is the chain formation time which is shorter
than the solidification time.
As seen from equations (6) and (7), collisions may interrupt the
structure formation if some particle's position change in the integration
is too big. Therefore, we specify a value 5r~. If the largest position
change among all the particles' motion during one time step 5t is greater
than 5r~, the next time step is reduced. Otherwise, 5t will be increased.
Our method speeds up the integration and effectively preVents possible
problems from the collisions because we control the position change by
selecting a proper 5r~. Since the time step changes at every step with
this method, we must employ Eq.(10) to handle the Brownian force.
There are some special situations which need to be discussed. In the
over-damped case, the viscous force becomes so strong that we have

i~* - -i~*+A(F*+BR*) ~ 0. (14)

The over-damping condition requires

IA(F*+BR*)I << 1. (15)

The constitutive equation (11) can then be simplified to

i~ - A(F* + BR*). (16)

Now, setting t* - (/A, we have

d r * / d ( - F*+BR* (17)

which is independent of A. Therefore, in the over-damped case the fi-


nal structure is independent of A, but the solidification time is inversely
proportional to A. From Eq.(15), A << 1 / ( F ~ , + B) is required to see
the over-damped case. The maximum force F ~ , comes when there are
collisions between the particles or collisions between the particles and the
665

electrodes. In our simulation, we have found that when A _< 10 -a, the
final structure has its three order parameters independent of A and the
solidification time is proportional to I / A , indicating that the system is
in the over-damped situation.
The effect of the random force B R* in Eq.(11) needs special atten-
tion. Since R,*. is a Gaussian deviate, we should compare the magnitude
of F/* with B. When the particles are randomly distributed, a typical
value of F/* in Eq.(11) is ,,~ n 4/a where n is the particle density. In our
simulation n - 0.3486 which gives an estimation of 0.2 for F/* at a random
distribution. When two particles get very close, their dipolar interaction
force has a typical value ,,~ 1. However, when the particles aggregate to
form a final structure, the joint forces on every particle are vanishingly
small. Therefore, it is easy to understand from E q . ( l l ) that if B << 0.2,
B R ~ has little effect on the early dynamic process, but may have some
effect on the final structure since the joint forces on each particle are
small. If B >> 1, B R ~ becomes the leading force in Eq.(11).

1.0

(a) 05 f , , , , , . , , , , . ~ ~ y- ry structure
0.8

04
0.6
BCT Lattice
13:1 07
0.4
2"
O~
x~.~~ass-like structure ,.< Glaas-like structure
0.2
O~ Nematic liquid crystal
BCT Lattice ~ ~ ----
Liquid
0.0 : 7 - / ' 7 7 " / 7 - . / 7 ./-7- 7 " 2 7 7" 7 O( , ! , l. , ! , ! ,
I . . . . . . . . I' ........ I ......... I ......... I "''''''~"

A o.o 0.2 0.4 0.6 0.8 1.0 A 0.0 0.2 0.4 0.6 03 1.0

Fig.1. (a) Different structures under various conditions. In the shaded


area, ER fluids may develop into a poly-crystalline structure. The bound-
ary between the liquid and the nematic liquid crystal has A B 2 ,,., 0.25
(~ ,,o 6.7). The boundary between the bet lattice and liquid crystal state
has A B 2 ,,~ 10 .2 (A ,,o 167). (b) The same results presented with axes
A and ~. The poly-crystalline structure and glass-like structure are non-
equilibrium products.

4. STRUCTURE OF ER FLUIDS

If the ER system can always evolve into its equilibrium state after
666

an electric field is applied, the final structure may only depend on the
parameter A - p2/(e.la3ksT ). However, in the actual dynamic process,
the ER system may not be able to reach the ground state after it is
trapped in a local energy minimum state. The viscous force enhances
such a possibility by a quick reduction of the particles' kinetic energy.
z(~) z(~)
14.00 .~(: 14.00

!
,, ,7 0 / it ~

g.s3

7 O0 7.00
(

467 ( 4.67

2 ~3 2. ]3

..~2. '1"92 ~.413


0 00" 6i O. O0
3. 4 d
,9
2.42 1.42 ~
0.9t
x(, v 0.56

Fig.2. The initial state has the Fig.3. A good body-centered


particles randomly distributed in tetragonal lattice is formed with
space. Pt, P2 ~ 0.8 and ps >__0.9.

Varying the two parameters A and B in the simulation can be re-


alized by changing the applied electric field and temperature in the ex-
periment. We have clearly seen five different structures of E R fluids in
a wide range of A and B" liquid, nematic liquid crystal, glass-like struc-
ture, poly-crystalline structure, and bct lattice. Figure 1 depicts the final
structures the system evolves over a wide range of A and B. In order to
make a comparison, we derive all these final structures in our simulation
from an initial randomly distributed state (Fig.2). In Fig.la, we present
the result with axes A and B. In Fig.lb, we present the same result
with axes A and )~. The boundary between the liquid and the nematic
667

liquid crystal and the boundary between the bct lattice structure and the
nematic liquid crystal structure seem to have A = 6.7 and A = 167 respec-
tively. Although this is expected for an equilibrium statistical physics,
we should note that the final structure in our simulation may not be an
equilibrium state. Therefore, Fig.1 may differ from a conventional phase
diagram.
4.1 BCT Lattice and Poly-crystalline Structure
As seen from Fig.l, the ordered state, a bct lattice, has been obtained
in quite a wide range of A and B. In Fig. 3, we plot a typical bct lattice
structure with order parameters p], p~ ~ 0.8 and p3 >_ 0.9. The projection
of tile three-dimensional structure onto the x-y plane shows a centered
square lattice (Fig.4), a typical characterization of the bct lattice [4]; the
marked square also has its side ~ lv/~.5.ha; and the characterization of
these chains is also correct for the bct lattice [4].

"1
.--/ .'7~
~ 3
/ " /.

~ 2
9 l

.... I ' ' ' ' 1 ' - ' ' " 1 .... I .... I

0 1 2 3 4 5
x (~)
Fig.4. Projection of the bct lattice on the x-y plane. The marked square
has its side ,-~ x/1--5a.

We also note that there is a small region inside this ordered region
where the system may likely develop into a poly-crystalline structure
(Fig.5). This small region has a very small B and a big A (___ 0.1).
Examination of this structure reveals that the system has thick columns
consisting of several bct lattice grains. However, these grains do not form
a single crystal. There is some mismatch, mainly caused by their rotation
around the z axis by slightly different angles. In Fig.6, we plot a part of
a thick column of poly-crystalline structure which clearly shows a twist
of bct lattice grains. Since these rotations do not affect P3 very much
but reduces p] and p~, P3 remains ,-, 0.9 while Pl and p2 are reduced to
,-, 0.5. The poly-crystalline structure is a produ~:t of fast solidification.
668

Because of very small B in this region, the E R system may not be able to
relax into a good crystal. It is thus easy to understand that in this region
the final structure is somehow sensitive to the initial random state. The
computer simulation confirms the conclusion, too: in this region from
some r a n d o m initial state the E R system may develop into a good bct
lattice, while from some other random initial state the system ends up in
a poly-crystalline structure.
To further understand the issue, we have paid special attention to
the situation of B = 0 which can be realized at zero temperature. If
A <__ 10 -a, the system is over-damped. The final structure has the three
order parameters independent of A: pl and p2 around 0.62, and P3 around
0.89.
ZCa)
z(,,) 14.oo
14.00

11.67
11.57

9.33
9.,]3

7.00

7.00
4.57

4.57
2.33

2.33
0.00
4.17 2.77 .9G
0.00 60 + i
4.
f~ I. 0 o ~ .L.o,
Fig.6. A thick column of poly-
Fig.5. A poly-crystalline struc- crystalline structure which clearly
ture. sllows a twist of bct lattice grains.
(:7.

If A is between 10 -3 and 0.1 with B - 0, the final structure is


improved with Pl and p2 around 0 . 8 - 0.9 and pa around 0 . 9 0 - 0.93, a
rather good bct lattice. The order parameters of the final structure in
669

this region depends on A. As A continues to increase, the three order


parameters are fluctuating in the poly-crystalline region. For example,
when A > _ 0.1 and B - 0, the final structure developed from the initial
state in Fig.1 has three order parameters around 0.5 while P3 is close to
0.6. However, the final structure at A _> 0.1 and B - 0 is now very
sensitive to the initial random state. For example, after changing the
initial state at A - 0.1 and B = 0, we have ended up with a good bct
lattice" pl and p2 are close to 0.84 and p3 is close to 0.92. This also
implies that the ER system at B = 0 can be easily trapped in a local
energy minimum. The final structures derived at a moderate B is not
sensitive to the initial random state because the ER system can get out
from a local minimum energy state and develop into the global energy
minimum state with the help of the thermal fluctuations.
z

14.00
14.0

11.67
11.6

9.3
9.J3

7.0 7.00

4.6
4.67

2.3
2.33

0.0 ,r
e
3.62
0.00 48
4. ~~
o.83

Fig.7. In a nematic liquid crys- Fig.8. In a nematic liquid crystal


tal state, the particles do not ag- state, the system has ordering in
gregate together. D(x, y) is quite the z direction but almost no or-
uniform over most of the region. dering in the x-y directions.

We have compared the final structures in the poly-crystalline region


670

when we fix A and increase B. For example, at A = 1.0 and B = 0, the


final structure derived from the initial state in Fig. 1 has Pl and p2 around
0.5-0.6 and pa abot/t 0.83. When we increase B, although we continue to
start the system from the same initial state, the final structure improves.
For example at A = 1.0 and B = 0.1, the final structure has Pl and p2
around 0.8 and P3 around 0.87. This implies that a moderate B can help
the system relax into a global energy minimum state.
z (~,)
14.00
14.00

11.67
11.67

9. ,1,1 9.,13

7.00 1.00

4.67
4.67

2. ,I,1
2.33

o.o ,1.92
2.30
0.00 .x,
! ,68
] .... z.~l 1.15
x(~)
Fig.9. In a glass-like structure, Fig.lO. In a glass-like structure
D(x, y) is peaked in a small colum- tile ER fluids form thick columns
nar region, a main difference be- but t h e t h r e e order p a r a m e t e r s ar~
tween a glass-like structure and almost the same as that of tile ne
liquid. matic liquid crystal state.
4.2 Nematic Liquid Crystal and Glass-like Structure
W h e n we increase B, equivalent to raising the t e m p e r a t u r e , we come
from the region of the bct lattice to a region of final s t r u c t u r e with sig-
nificant P3, but small pl and p2. Typically, pl and p2 are around 0.3 while
P3 >_ 0.6. This implies t h a t the system has ordering in the z direction
with weak or no ordering in the x-y directions. From the definition of a
particle density
671

N
D(r) - E 5(r-rj) (18)
j=l

where rj is the position of the j t h particle center, we define a columnar


density,

D(x, y) - / D(r)dz. (19)


0

After analyzing the columnar density of these structures, we find that


within this region, there are two slightly different structures. When A
is relatively small and B is relatively strong, D(x, y) is quite uniform,
as in Fig.7. This implies that in these structures, the particles do not
aggregate together to form thick columns though there is some ordering
in the field direction. The system remains in a liquid state, but similar
to a nematic liquid crystal structure (Fig.8). When A is relatively strong
and B is relatively small, the ER fluids form thick columns. As indicated
by D(x, y) in Fig.9, the particles are concentrated in a small region, a
main difference distinguishing this structure from a nematic liquid crystal
structure. The three order parameters of this structure are not too much
different from that of a nematic liquid crystal structure. Although there is
some ordering in the fielddirection, there is no significant lateral ordering.
Therefore, this is a glass-like structure (Fig.10). In this region, the strong
electric field forces the particles to aggregate to form thick columns, but
the thermal fluctuations prevent the system from forming a crystalline
structure.
4.3 Liquid
A further increase of B leads to a region which has the final structure
in a liquid state (Fig.l 1). All three order parameters are very small
for these structures. The particles are randomly and quite uniformly
distributed in the space, as seen from D(x,y) in Fig.12. In this region,
the random Brownian force is too strong to prevent formation of any
ordered structures.
4.4 Non-equilibrium Process and Boundaries.
Our simulation shows a dynamic process. The poly-crystalline struc-
ture is a product of non-equilibrium processes. The difference between
the glass-like structure and liquid crystal is only in the columnar den-
sity, not in the ordering: In a glass-like structure, the particles aggregate
together while they do not in a nematic liquid crystal (see Fig.7 and
672
Fig.9). Therefore, poly-crystalline and glass-like structures may not be
closely related to the equilibrium state. On the other hand, both the
boundaries between the liquid and nematic liquid crystal and the bct
lattice and liquid crystal structure seem to be related to the equilibrium
state. Although they are not exact, these two boundaries both have A B 2
roughly as a constant. Since the parameter A is proportional to 1/(AB2),
these two boundaries are roughly along the lines of a constant A [17].
The boundary between the liquid and the nematic liquid crystal has A B 2
close to 0.25, corresponding to A ,.- 6.7, The boundary between the bct
lattice and liquid crystal state has A B 2 close to 10 -2, corresponding to
,.- 167.
z (,) DCx,v)
14.00 14.00
11.57 11.67 [~
9.33
9.,33 I li~ ~
7.00 7.O0 f-~ II "
4.67 4.57 ~t II
2..13
4.20
0.0~
0.00
4. ~"~
45

.f" (~j 0.75


Fig.ll. In a liquid state, three Fig.12. In the liquid state, D(x, y)
order parameters are vanishingly is quite uniform over the whole re-
small and the particles are ran- gion.
domly distributed in the space.
5. RESPONSE TIME OF ER FLUIDS

ER fluids are marked for their fast response to an electric field. A


number of experiments established that a typical response time of ER
fluids is of the order of milliseconds. This response time is Usually defined
as the time needed for ER fluids to have a significant viscosity increase
673

immediately after an electric field is applied. In our simulation, we define


the solidification time as the time interval between the application of an
electric field and the establishment of a final structure [20]. It is clear
that our solidification time should be longer than the response time since
ER fluids deliver a significant increase of viscosity before they reach their
final structure. However, these two time scales are closely related and
our solidification time has clear physical meaning and is important for
applications as well.
The relationship between the solidification time and the parameters
A and B is in Fig 13. We note that at a fixed B, the solidification time
is getting longer as A gets smaller. In the over-damped case, Eq.(17)
indicates that the solidification time is inversely proportional to ' A. Our
simulation verifies this conclusion,

t~o,id 60/A.
- (20)

This relationship holds up to A ,,~ 10 -2. In real time, for example, at


A - 10 -3, this solidification time is of the order of a second. As the
value of A increases, the viscosity reduces. When the system is not over-
damped, the solidification time further decreases as A increases, but this
reduction is slower than 1/A. For example, as A increases from 10 -2 to
10 -1 , the solidification time only slightly reduces.
109

lOa o B=I
o B=0.1
k: " B=O.01
o'3
107
E
0
m
.+J
10e
.p..i

=
10s
o~.,4

104

103 .,,l ! t i |ll|ll i i | tJl|ll __l ! Ill|l|l I ! * |nlJll I I I |Ill|

1 0-e 1 0-s 1 0-4 1 0-3 10-2 1 0-f


A

Fig.13. The relationship between the solidification time and the param-
eters A and B.
674

At a fixed A, the solidification time increases with B. This is due


to high fluctuations of the Brownian motion. However, the effect of B is
significant only when B is large enough. For example, if B < 10 -2, the
solidification time is almost unaffected by B. If B > 10 -2, the thermal
fluctuations delay the solidification process. For example, at A = 10 -~
and B = 10 -2, the formation of a bct lattice structure takes about 1.738
105t0 while at A = 10 -3 and B = 10 -1 the solidification of a similar bct
lattice takes 2.81 x 105t0.
In our simulation, we also determine the chain formation time by
examining the order parameter p3. From Fig.14, it is clear that the chain
formation is much faster than the formation of a final structure. This
again implies that the particles in ER fluids form chains first, then chains
aggregate together to form thick columns. Typically, the chain formation
time is about one third of the solidification time or shorter. In the over-
damping case, the chain formation time is also proportional to 1/A. We
also notice that in real time, the chain formation time is of the order of
milliseconds, the same order as the response time found in engineering
applications.

'~ I%
lOe
%,.
o
o
B=I
s=o.1
I- \\ , s=o.ol
lO7 =

1 oe

"~ lOS

104

10a
1 o-e 10-S 10-4 10-a 10-2 10-I
A
Fig.14. The relationship between the chain formation time and the pa-
rameters A and B.

6. DISCUSSIONS

In this section, we want to compare our simulation results based


675

on the constitutive equations with experiments. For a real E R system,


such as dielectric particles in petroleum oil, e / ~ 2, % >> 1, ~ ~ 0.2 poise,
a ~ 10#m, and the mass density of the particle p ,.., 3 g / c m 3. We estimate
to ~ 8.33 x 10 -7 s. If we choose the subinterval 7 = 0.4t0, then as E0
varies from 0 to 4 K V / m m at T=300 K, A changes from 0 to 10 -2 and B
reduces from c~ to 10 -3. When A - 10 -2 and B - 10 -2, for example, our
simulation finds the chain formation time is about four milliseconds while
the bct lattice and the solidification time is less than one second. In the
experiment, the chain formation takes milliseconds to complete, but the
formation of bct lattice is slower than that in our computer simulation
[7].
As the particle size becomes big, the inertial time to and A in-
crease. For example, if the above ER fluid has everything the same ex-
cept a ,.~ 100#m instead of 10#m, then we have A ,,~ 1 at E - 4kV/mm.
Hence, from Fig.l, we notice that ER fluids with large particles are easy
to develop into a poly-crystalline structure in the non-equilibrium pro-
cess. This interesting result is useful in production of composite materials
by the ER effect [9,10].
Our results at B = 0 are interesting enough to warrant some ex-
perimental investigation. The fact that the final structure at B = 0 is
sensitive to the initial state indicates that the Brownian force plays an
important role in driving the ER system from a'local energy-minimum
state into a global energy-minimum state. On the other hand, if B is too
strong, the thermal fluctuations prevent the system from forming a good
bct lattice. Therefore, an experimental determination of this range of B
will be very interesting. This goal may be achieved by examination of
ER fluids at cryogenic temperatures.
Our simulation also reveals that the response time defined in ER
engineering applications is related to the chain formation time. We have
also found a relationship between the solidification time and the viscosity,
temperature, and electric field. It will be very interesting to see if this
relationship holds in experiments.

A CKN OWLED G E M ENTS

This research is supported by a grant from National Science Foun-


dation DMR-9622525.

REFERENCES

1. Electrorheological Fluids, edited by R. Tao and G.D. Roy (World


Scientific Publishing Comp., Singapore, 1994).
2. L. C. Davis, J. Appl. Phys., 72 (1992), 1334; 73 (1993), 6so.
676

3. H. Block and J.P. Kelly, US Patent No. 4,687,589 (1987).


4. F.E. Filisko and W.E. Armstrong, US Patent No. 4,744,914 (1988).
5. R. Tao and J. M. Sun, Phys. Rev. Lett., 67 (1991), 398; Phys. Rev.
A, 44 (1991), R6181.
6. J. E. Martin, J. Odinek, and T. C. Halsey, Phys. Rev. Lett., 69
(1992), 1524.
7. R. Tao, J.T. Woestman, and N.K. Jaggi, Appl. Phys. Lett., 55 (1989),
1844.
8. T. J. Chen, R. N. Zitter, and R. Tao, Phys. Rev. Lett., 68 (1992),
2555.
9. G. L. Gulley and R. Tao, Phys. Rev. E, 48 (1993), 2744.
10. X. Wu. X. Zhang, R. Tao, and R. P. Reitz, Bull. of Amer. Phys.
Soc., 41 (1996), N.1, 191.
11. C. A. Randal, C. P. Bowen, T. R. Shrout, G. L. Messing, and R. E.
Newnham, in ref. 1, p516.
12. R. Tao and Q. Jiang, Phys. Rev. Lett., 73 (1994), 205.
13. D. J. Klingenberg, F. van Swol, and C. F. Zukoski, J. Chem. Phys.
91 (1989), 7888; 94 (1991), 6170.
14. N. K. Jaggi, J. Stat. Phys. 64 (1991), 1093; W. Toor, J. of Colloid
and Interface Science 156 (1993), 335.
15. K. C. Hass, Phys. Rev. E, 47 (1993), 3362.
16. R. T. Bonnecaze and J. F. Brady, J. of Chem. Phys.,96 (1992), 2183.
17. H. X. Guo, Z. H. Mai, and H. H. Tian, Phys. Rev. E, 53 (1996),
3823.
18. R. Tao, Phys. Rev. E, 47 (1993), 423.
19. For example, see, S. Chandrasekhar, Rev. Mod. Phys. 15 (1943), 1;
R. Reif, Fundamental of Statistical and Thermal Physics (McGraw-
Hill, New York, 1965), 560-562.
20. P. M. Adraini and A. P. Gast, Phys. Fluids, 31 (1988), 2757.
21. H. See and M. Doi, J. of Phys. Soc. of Japan, 60 (1991), N.8, 2278.
677

ELECTRO-MAGNETO-HYDRODYNAMICS
AND SOLIDIFICATION

G. S. Dulikravich

Aerospace Engineering Department, The Pennsylvania State University,


University Park, Pennsylvania 16802, USA

1. INTRODUCTION

Fluid flow influenced by electric and magnetic fields has classically been
divided into two separate fields of study" electro-hydrodynamics (EHD)
studying fluid flows containing electric charges under the influence of an
electric field and no magnetic field, and magneto-hydrodynamics (MHD)
studying fluid flows containing no free electric charges under the influence of a
magnetic field and no electric field. Traditionally, this division was necessary
to reduce the extreme complexity of the coupled system of Navier-Stokes,
Maxwell's and constitutive equations describing combined electro-magneto-
hydrodynamic flows. Recent advances in numerical techniques and computing
technology, as well as fully rigorous theoretical treatments, have made analysis
of combined electro-magneto-hydrodynamic flows well within reach. A survey
of electro-magnetics and the theory describing combined electro-magneto-
hydrodynamic (EMHD) flows is presented with an emphasis on describing the
intricacies of the mathematical models and the corresponding boundary
conditions for fluid flows involving linear polarization and linear
magnetization. This survey concludes with a presentation of EHD and MHD
flow models involving solidification.

NOMENCLATURE
b - electric charge mobility coefficient, kg A s2
B_B_ = magnetic flux density vector, kg A 1 s2
678

(
d = Vv_ + Vv = a v e r a g e rate o f d e f o r m a t i o n t e n s o r , s -~

Do = e l e c t r i c c h a r g e d i f f u s i o n c o e f f i c i e n t , m 2 sl
D=eo_E+P = e l e c t r i c d i s p l a c e m e n t field v e c t o r , A s m -2
e : c~T + (v._v)/2 = total e n e r g y p e r u n i t m a s s , m 2 s ~
E = e l e c t r i c f i e l d v e c t o r , k g rn s 3 A -~, or V m ~
E=E+vxB = e l e c t r o m o t i v e i n t e n s i t y v e c t o r , k g m s 3 A -1
f = mechanical body force vector per unit mass, m s2
g = a c c e l e r a t i o n d u e to g r a v i t y , rn s -2
h = h e a t s o u r c e or s i n k p e r u n i t m a s s , m 2 s -3
a_=B_/ to-M = m a g n e t i c f i e l d i n t e n s i t y v e c t o r , A m -1
J= + Jd = electric current density vector, A m 2
J, = electric conduction current vector, A m 2

J d = V_qo = e l e c t r i c drift c u r r e n t v e c t o r , A m -2
M = total m a g n e t i z a t i o n v e c t o r p e r u n i t v o l u m e , A m ~
M=M+vxP = m a g n e t o m o t i v e i n t e n s i t y v e c t o r p e r u n i t v o l u m e , A m -1
P = p r e s s u r e , k g m -~ s -2
P = total p o l a r i z a t i o n v e c t o r p e r u n i t v o l u m e , A s m -2
qo = total or f r e e e l e c t r i c c h a r g e p e r u n i t v o l u m e , A s m -3
q = h e a t flux v e c t o r , k g s -3
S = e n t r o p y p e r u n i t m a s s , m 2 kg-' K" s -2
T = absolute temperature, K
V = fluid v e l o c i t y v e c t o r , m s -~

GREEK SYMBOLS
= v o l u m e t r i c t h e r m a l e x p a n s i o n c o e f f i c i e n t , K -1
= C h o r i n ' s ( 1 9 6 7 ) artificial c o m p r e s s i b i l i t y c o e f f i c i e n t
E = d i e l e c t r i c c o n s t a n t ( e l e c t r i c p e r m i t t i v i t y ) , k g -~ m -3 s 4 A 2
ro = 8 . 8 5 4 x 10 -12 = v a c u u m e l e c t r i c p e r m i t t i v i t y , k g ~ m 3 s4 A 2
8r = S / F , o = relative electric permittivity
K = thermal conductivity coefficient, kg m s3 K l
= e l e c t r i c c o n d u c t i v i t y c o e f f i c i e n t , k g -1 m -3 s 3 A 2
P = fluid d e n s i t y , k g m -3

_x_v = 2 g v d + g v E I ( V 9v ) - N e w t o n i a n v i s c o u s stress t e n s o r , k g m -1 s -2
EM
x = e l e c t r o m a g n e t i c stress t e n s o r , k g m -1 s -2
679

v EM
x__=x__ + x__ = stress tensor (viscous plus electromagnetic), kg m -~ s -2
~t = magnetic permeability coefficient, kg m A 2 s -2
~to = 4~ x 10 - 7 = magnetic permeability of vacuum, kg m A -2 s 2
ktr = ~t [ k t o = relative magnetic permeability
~tv = shear coefficient of viscosity, kg m ~ s -~
~tv2 = second coefficient of viscosity, kg m ~ s -~
E = er _ 1 = electric susceptibility
M = [t r _ 1 = magnetic susceptibility
= electric potential, V
= x__~'d = viscous dissipation function, kg m ~ s -3

2. BACKGROUND

The scientific field of study that analyzes the ability of electro-magnetic


fields to influence fluid flow-field and heat transfer has been investigated for
decades. The equations that are most often used to model this phenomena
consist of the system of Navier-Stokes equations for fluid motion coupled with
Maxwell's equations of electro-magnetics augmented with the material
constitutive relations. The field studying these flows is often called electro-
magneto-dynamics of fluids [ 1], electro-magneto-fluid dynamics (EMFD) [2-5],
electro-magneto-hydrodynamics [6], magneto-gas-dynamics and plasma
dynamics [7], or the electro-dynamics of continua [8-10]. The full system of
governing equations has, until recently, been far too difficult to solve because
Navier-Stokes system becomes very complex when modeling flows involving
turbulence, chemical reactions, multiple phases, non-Newtonian effects, etc.
When coupled with Maxwell's equations, the complexity of the combined
EMHD system is raised by orders of magnitude. To reduce this complexity, the
analytical modeling has traditionally been divided [11] into flows influenced
only by externally applied electric fields acting upon electrically charged
particles in the fluid, and flows influenced only by externally applied magnetic
fields without electric charges in the fluid. The former are called Electro-
Hydrodynamic (EHD) flows [12] and the latter Magneto-Hydrodynamic (MHD)
flows [13]. More recently, rigorous continuum mechanics treatments of EHD
[14] and unified EMHD flows [9,10] have been developed. These continuum
mechanics approaches are limited to non-relativistic, quasi-static or relatively
low frequency phenomenon [ 15-17].
680

This chapter should provide an introductory survey of the background theory


to allow implementation of numerical analysis of unified EMHD flows and of
classical MHD and EHD flows with addition of liquid/solid phase change. An
overview of electro-magnetic theory with concentrated effort placed on
descriptions of the electric and magnetic fields and electric charges and currents
will be made to provide a physical understanding of the field-material
interactions causing polarization and magnetization effects. The system of
equations governing the unified EMHD theory and the corresponding boundary
conditions will be presented together with its fully conservative form that is
ready for numerical discretization.

3. POLARIZATION AND GAUSS' LAW

Charge polarization is created when electric charges of opposite signs are


separated by a distance. Although many references define several sources of
polarization [ 18], there are essentially two main sources of polarization: natural
and induced [13]. Natural polarization arises from natural dipoles and charged
particles. An example of a natural dipole is a water molecule which has a
geometry such that the centers of positive charges and negative charges do not
coincide. Since the molecules are allowed to move freely and orient randomly,
water will not have polarization on a continuum level. Now consider the fluid
water as it is frozen with an applied electric field. An induced polarization will
be created by the electric field by inducing an initial charge separation in
neutral particles [19], by causing greater charge separation within the
molecules, and by causing molecular alignment with the applied electric field in
case of natural dipoles [19]. Once locked in the ice crystal structure, the water
molecules will no longer be able to change their position or orientation.
Consequently, even after the electric field is removed, the ice will still have
polarization on a continuum level since the polarization caused by the electric
field aligning the water molecules was literally frozen into the ice.
From this example it may seem that there is no reason, when dealing with
fluids, to consider natural polarization. This, however, would be an erroneous
assumption. Though the natural polarization may show no continuum effects
without the presence of an electric field, in an electric field the total
polarization, _P, combines both the induced polarization due to the electric field
and the natural polarization of the molecules which are now aligned by the
electric field [ 13, p.22].
681

If polarization is assumed to be a linear function of the steady or relatively


low frequency electric field, then it can be defined as

P__- GoXE(E__4- v x B)-- gp(E q- v x B ) = GpE (1)

The electric displacement vector then becomes [19, p.164] [9, p.178].

__D-- 8oE_E_+ P__= 8o(1 + xE)E + CoXE~ x B -- go8r___E+ 8p V X B - ~___+ 8p V X B (2)

where the material property, ZE, is the dielectric susceptibility. It is typically


obtained experimentally [20, p.86] and could be a function of frequency.
Electric charges come in two types: free and bound. Free charges arise from
electrons in the outer or free atomic shells and from ions. Bound charges are
those arising from the molecular geometry and displacement of atomic inner
electron shells [13, p.21]. Gauss's law for a linearly polarizable medium then
becomes [ 13, p.22]

V.D_=qo (3)

or

V. (ao___+ P_)= V .(~_ + epVX B_B_)= qo (4)

At this point it is important to note that qo multiplied with the charged particle
drift velocity, v d , creates the convection or drift electric current, Jd [ 13, p.67],
while polarization current, J_p, is defined as the variation of the total
polarization with respect to time [ 19, p. 121 and p. 147].

4. M A G N E T I Z A T I O N AND A M P E R E - M A X W E L L ' S L A W

If the material in question may be considered linear, that is, if the


magnetization is a function of one material property and the strength and
direction of the applied magnetic field, then the magnetization is defined as [9,
p.178] [19, p.164] [20, p.92-96] [21, p.371-377]

~M

Po ~1+
~M)
682

In addition to the electric currents arising from magnetization and direct charge
motion, other phenomenological currents have been observed and must be taken
into account when defining the total current, J [9, p.162-163]. Introducing the
effects of magnetization and polarization and rearranging constants, the
Ampere-Maxwell's law of electrodynamics may be rewritten as [ 13, p.30]

VxB__--po VxM-k-Jd +~p -k- (6)

Magnetization and magnetic field vectors are often combined to form the
magnetic field strength vector, H__,defined as

H - B _ M (7)
~to

The total current, J, is defined as the sum of the apparent magnetization current,
V x M, charge drift current, Jd, and phenomenological polarization currents,
J, [13, p.26] since the contribution to the magnetization current by intrinsic
magnetization is zero. The Ampere-Maxwell's law for polarizable, magetizable
media can therefore be written as [ 19, p.132]

0D
- VxI-I=-j (8)
0t
Detailed descriptions of these equations can be found in any number of texts
[19,20,21].

5. A MODEL OF UNIFIED ELECTRO-MAGNETO-GASDYNAMICS


(EMGD)

The full system of equations governing unified EMGD flows consists of the
Maxwell's equations governing electro-magnetism, the Navier-Stokes equations
governing compressible fluid flow, and constitutive equations describing
material behavior. Assuming a single-phase fluid and only one type of charged
particles in the fluid, this set has a minimum of 12 partial differential equations
that contains 13 unknowns: p, qo, T, p, and the three vector components of _v, E__,
and B, respectively. The thirteenth equation is the equation of state for the
683

fluid. The foundations of the electro-magneto-gasdynamic (EMGD) theory


were formulated by Eringen and Maugin [9,10] and are based on continuum
mechanics [22-25]. The rigor with which the constitutive, force, and energy
terms were derived leads to a model more complete and robust than any of those
found in classical literature [8,7,1,11-13,18-21 ].
Dulikravich and Jing [6,26] have shown that a compact vector form of the
unified EMGD system can be written as a combination of the Maxwell's
electro-magnetic subsystem and the Navier-Stokes fluid flow subsystem.
The Maxwell's subsystem (consisting of seven PDE's) is composed of
Ampere-Maxwell's law for polarizable and magnetizable medium

8D
- - VxH=-J (9)
Ot

that can also be written as

OE Vx--=--- l+ - (10)
8t eo eo

Faraday's law

0B
- ~VxE=0 (11)
8t

and conservation of electric charges

8qo
~+ V._J_ = 0 (12)
8t

that is a combination of Gauss' law

V.DD_ = qo (13)

and the Ampere-Maxwell's law. Conservation of magnetic flux

V.B =0 (14)

is also a part of the Maxwell's subsystem, but is not solved for explicitly.
684

The second part of the unified EMGD is the viscous, compressible flow
Navier-Stokes subsystem consisting of five PDE's and an equation of state of a
perfect gas. It is composed of conservation of mass equation

OP + V.(p_v)- 0 (15)
0t

and a conservation of linear momentum (including electromagnetic effects)

0p_v
+V (v_gv+pI-~)
9 _ --
V (v(P
9 __ X
B/) --
V ((B_B_M II+(___ P)I)
9 9 = " -"
P S V
(16)
0t

Here, I is the identity (unity) tensor and S v is a vector of source terms. The
following dyadic identifies were used in equation (16)

(va). M_= v. ((B. M _ ) I ) - ( ) . B_ (17)

(vE). e = v. ((E. e)i)-(ve)._E (18)

Conservation of energy equation is also a part of the Navier-Stokes subsystem

0(Pe----~)-t- V- (pev + (pI-__x)-v)+ V. ~ 1 - p h - 9 E . ~ + Iql. D B - J c .1:: = 0 (19)


Ot - - - Dt -- Dt -

It can be replaced by the entropy generation equation [9,2,4]

9 -+&.E
Ds _ oh +_______~@V.
_ -T5 J- Dt -- Dt (20)
9Dt- T T

The viscous stress tensor for a non-linear fluid is given as

v
1; - 21avd + ).tv2I(V. _v) + oqd 2 (21)
685

In the case of a media with non-linear physical properties, the unified EMGD
formulation for the electric conduction current and the heat flux can be
expressed as [9, p. 161-162].

Ic -- O'1~q- G2~'--~ -t- ~3 d2 "E + G4VT 4- ~5 d" VT + 0"6d2" VT + o7E x B


+ cY8(tl-(E x B ) - (d- E) x B__)+ cY9VT xB (22)
+ o'lo(d_. (VW x B ) - ( d " VT) x BB)+ (Yll(B 9E)B__+ o'12(B_-VT)B__

q = K1E -I'-K2d" ~ q- K3d2 "E q- K4VT + tcsd-VT + K6d2. VT + K:vE_x B


+ ~a(a=. (g B_)- (a_. E) B_)+,,:9VT B_ (23)
+ Klo(d" (VT x B ) - ( d - V T ) x B)+ Kll(B'E)B + KI2(B 9VT)B

The electro-magneto-thermal stress tensor for a non-linear fluid can be


expressed as [9, p. 177-178]

EM
I; = ~2~ (~)E + ot3B (~)B + o~4VT (~)VT + or5(_E| d. E)s
+ 0~,6 (_E~)d 2 -~__.1+o~7(VT | d-VT)s+ ot8@T | 2- VT~

-ko~9(_d.'W-W-d)q-O,,loW.d-W -+-Otll(d=2 . W - W . d 2)
(24)
- w 2 . . w),- 0,3( | T/s + E | E-W/s
~ W" F)S -t- O~16~" ~(~)W2 9E_~ + oqv~W-(E| VT- VT |
+ O~15(___.~) s
+ o~,18d-(E (~ V T - VT | 0~,18(E 1~)VT - VT |

where ~____W=Wij=gijkBk, while the subscript s indicates symmetrization.


Expressions for total polarization, P, and magnetization, M, of non-linear media
can be modeled with expressions of similar complexity [9, p.175].
In these formulas, c~i, cE and Ki are the physical properties of the media. Most
of these coefficients are still unknown although their exploitation can offer
potentially significant benefits in applications involving interacting electric,
magnetic, thermal, and stress fields. This theory is valid for the frequencies of
the electric and the magnetic fields that are less than approximately 1 kHz and
for fluid speeds considerably less than the speed of light [14-17]. For higher
frequencies, certain physical properties become functions of the frequencies.
For higher speeds, relativistic effects will have to be taken into account.
686

6. CONSERVATIVE FORMS OF E L E C T R O - M A G N E T O -
HYDRODYNAMIC (EMHD) SYSTEM

A necessary condition that an iterative numerical solution of the EMGD


system will converge to the exact solution of the analytical EMGD system as
the computational grid is infinitely refined, requires that the EMGD system
must be rewritten in a fully conservative (divergence-free) form. This is
especially needed if strong gradients of dependent variables are expected to
exist in the solution domain. The fully conservative forms can then be used
directly in the finite difference, finite volume, or finite element discretization of
the EMGD system and its iterative integration process.
In the following derivations, it will be assumed that the fluid is
incompressible, homocompositional, that it has linear polarization and linear
magnetization properties, and that the frequencies of the applied electric and
magnetic fields are less than approximately 1000 Hz for this mathematical
model to be realistic. These are the only assumptions to be used in this model
which will be referred to as a unified electro-magneto-hydrodynamics (EMHD).
A fully conservative EMHD system in a vector operator form is given as [6]

OE_E_ Vx -H- = S E (25)


Ot 8o -

8B
+ VxE = 0 (26)
8t

Oqo
+v.J_ = 0 (27)
ot

V.v = 0 (28)

Ov
- + v . l v v + - (1p i - )) - 1 v . I v ( l , B)+(B.M)I+(_EE.p)!]=S v (29)
8t P P . . . . . -= -

-8e
- + - V1 .(pe_v+ ( p I - x=)-_v + _q ) = S e (30)
& p

For simplicity of notation we can define the following terms as [6,26]


687

1 1
F=/.to (r M) =--~t (31)

1 1
= =- (32)
Co(1+ Z E)

ep =8oZ E --e--e o (33)

A= (1
p +
zEI+Pe p B . B (34)

8 = pf_ + qo_E + J_ a + (v_). P + (va). M + v . (v(_P ~_))- v . (~_v + p i - a)(35)

_Dt=V IB /- -M_M_ - J _ = V x H - J _
go
(36)

Pt = A D t +A(VxE_)xx
80
(37)
+ p(1 + Z E Dt + x A[Rt
9 - P_P_x (V x E__)]x B

If we now assume that the fluid which is subjected to applied electric and
magnetic fields is of Newtonian type and if we allow only for linear
polarization (equation 1) and linear magnetization (equation 5), the constitutive
relations for the electric conduction current and for the heat flux vector become
[9, p.173-174]

lie ~ ~I(E~ -b X x B)-4- t34VT -t- o'7(]E -+-v x B__)x B


(38)
-!-(Y9VT B q- 0"1,(B 9(~E + Ir B__))B-b o12(B- VT)B
688

_q- K1(_E + v x B_)+ ~:4VT + K:7(_E+ _v x B_) x B


(39)
+ K:9VT x B_B+ KI, (B_-(__~+ v x B))B + KI2(B- VT)B_

Then, the EMHD source terms can be given in a compact vector form [6,26] as

S_E = 1 (j_.+ Pt ) (40)


~o

s_V _ [ + _ h o E + l (V x _ ) x ___-(VM). B_B- (VP__)-E+ (J + P--t)x B_B_] (41)


9

S e --h +!(EwvxB__.). [(v-v)e+ J_c + ~t]--!~Z M B. ((v-V)B-V x F~) (42)


9 9

Notice that these source terms have been formulated in such a way as not to
contain explicit time derivatives [6,26].

6.1 Fully conservative Cartesian form of the EMHD system


The EMHD system of equations (equations 25-30) can now be written in a general
conservative form in terms of (x,y,z) orthogonal coordinate system as

cqQ c3E0F 3G ~
~+ +~+~=S (43)
& 0x 0y 0z

Here, the solution vector of unknown quantities is given as

{
Q= E x, Ey, E z, B x, By, B z, qo, p Vx Vy v z e
}" (44)

where the asterisk symbol designates transpose of a vector. The vector of


source terms (those terms that do not contain divergence operator) is given as

- { x S , Sy, S z, 0, 0, 0, 0, 0,Sx,v Sy,v Sz,vse}


S= (45)
689

In equation (44), Chorin's [27] artificial compressibility coefficient, 13, was used
to create the unsteady term in the mass conservation since physical unsteady
term does not exist in the mass conservation for incompressible fluids.
By combining equations (5), (7), (1), and (31), the Cartesian components of
the magnetic field intensity vector can be defined [6,26] as

H x = gB x + t~pVy(E z + vxBy - VyB x ) - 8pV z (Ey + vzB x - vxB z)


Hy = ~By + t3pVz (E x + vyB z - VzBy) - 13pVx (E z + vxBy - vyB x ) (46)

Hz = g B z + gpvx(Ey + vzB x - v x B z ) - gpvy(Ex + vyB z - vzBy)

The flux vectors in equation (43) can then be defined as

Hz/eo
-Hy/eo /~o
0
-Ez
Ey -Ex
Jx
Jy
E= Vx
~=Vy

Vx
2 l EP Evvx-txy+
VxVy--? [V~y--}-~~)--'lTyy-- N~BP-VyN~yB)

Ev,vzt z+vz y t
(47) (48)
690

Hy/eo
-nx/~o
0

0
Jz
G=,Vz
1
VzVx-~ (~xz+ Vx~ ~)

~z +!(p-~zz-~.~-Vz~z
p- ~)
(49)
eVz
Here, we have written components of (__Px B_B_) as

NxPB= PyBz _ PzBy


PB
Ny =PzBx - P x B z (50)
N PB = PxBy - PyBx

In addition, we have defined the terms

NEPBM-- EX Px + EyPy + EzP z + B x ( Bx - H x ) + By ( B y _ Hy ) + B z ( Bz - H z) (51)


go go go

Nap = BxP x + ByPy + BzP z (52)

0T 0T 0T (53)
N ~ - ax ~ + ar ~ + az--~z
691

NV~ _ v x ( - p + Xxx)+ Vy'l~xy+ Vz'ISxz


v,~
Ny - v x'l;xy + Vy ( - p + "l;yy) + Vz'l;yz (54)

NV~ - VxXxz + Vy'l;yz + Vz ( - P + Zzz)

Components of the electric current vector, J, were defined as

(~1 ~ 0"7 PB ~Z 0T
Jx=vxqo+~Px+~4 +~N x +cy9(- B z- B )
~p ~XX {~p ~- -~Z y
O'11
+ NBpB x + O'lzNBTBx
~;p

~1 OT
~Z B x - ~ - B z )
= Vyqo + - - e y +0" 4 +--N +(I 9
ep c~ Sp Y
{3"11
+ - NBpBy + o'12NBTBy
E;p

O"1 OT G7 PB OT OT
Jz - Vzqo + ~ P z + 0"4 + Nz + (I9 By - B x)
{~p ~ZZ ~p (~- ~-

O'11.NBpB z + (I12NBTBz (55)


8p

and heat flux vector c o m p o n e n t s were defined as

qx = ~~(~px + K4 ~~ 1(7 PB + K9( ~ ,Bz - o~F B ) + ~KI~


+~Nx N B p B x + K:12NBTBx
8p -~- C~Z Y 8p

_ KI ~ 1(7 PB 0T 0T
{ly -8pPy
- 1" K4 ~0y' 1 " ~8pN y + K9(-~-ZBx - ~ - Bz)+ 1(1!
8p
NBpBy + K12NBTBy (56)

0T K:7 Pa ~ aT Kll
Clz - }c--LPz+ K:4 +~Nz + 1(9( By - - ~ - B x ) + - N B p B z + K12NBTB z
~p ~Z 8p 8p
692

7. CHARACTERISTIC-BASED I N F L O W AND O U T F L O W
BOUNDARY CONDITIONS

For most boundary value problems of electro-magneto dynamics, jump


conditions are exclusively used [9,28] to formulate solid wall boundary
conditions where a discontinuity occurs. At the inflow and outflow boundaries
where no surface or line discontinuities exist, an alternative approach based on
conservation law for continuous surfaces or lines become necessary.
Characteristic boundary condition formulation [29,30], which starts from a
characteristic form of the EMHD system, will be sketched here since it leads to
non-reflecting boundary condition formulation [31-36,26]. To find the
characteristic boundary conditions, it is first necessary to determine analytical
expressions for all eigenvalues of the characteristic system. The most common
approach is to use one of the symbolic programming languages software (LISP,
MACSIMA) in order to determine analytical expressions for each eigenvalue.
Since these software packages cannot be used for systems that have more than
five coupled partial differential equations, in the case of a complete EMHD
system which has twelve coupled partial differential equations, it is impossible
to find the eigenvalues using available symbolic programming software.
Consequently, we will use an alternative approach in which we will divide
the unified EMHD system into a Maxwell's subsystem and the Navier-Stokes
subsystem [33]. Each of these two subsystems will then be analyzed separately
by finding the analytical expressions for its eigenvalues by hand.

7.1 C h a r a c t e r i s t i c - b a s e d b o u n d a r y conditions for M a x w e l l ' s s u b s y s t e m


For example, characteristic treatment of the Maxwell's subsystem can be
formulated by rewriting the fully conservative Maxwell's subsystem

+~ + +~ SEM (57)
at Ox Oy az

in a non-conservative (characteristic) form as

a0EM+ AEM 0EM + BEM + CEM


0EM -- SEM (58)
0t - 0x 0y -- 0z
693

In order to perform characteristic analysis for Maxwell's subsystem, care


must be exercised to ensure that all the terms appearing in the fluxes
EEM, FEM,GEM are expressed as functions of the primitive variables

-
QEM--Ex,
{
Ey, E z, B x, By, B z, qo
}* (59)

For illustration, the flux vector EEM can be extracted from equation (47) as

0
Hz/eo
-Hy/8 o
EEM -- 0 (60)
-E z
Ey
Jx

For fluids with linear polarization and magnetization, Hz and Hy are the same as
in equations (46), while Jx is given in equation (55). The flux vector Jacobian
matrix A___EMis obtained as

0 0 0 0 0 0 0
azl a22 0 a24 a25 a26 0
a31 0 a33 a34 a35 a36 0
A EM -- 0EEM = 0 0 0 0 0 0 0 (61)
-- OQEM
0 0 -1 0 0 0 0
0 1 0 0 0 0 0
_a71 a72 a73 a74 a75 a76 Vx_

where the coefficients are

a21 = --~EVy a31 =--~Ev z (62)

a22 = zEv x a33 = ~Ev x (63)


694

a24 = ~Ev x Vz a34 = --~Ev x Vy (64)

= -Z E
a25 - ~EVyVz a36 VyVz (65)

a26- - +Vy a35 = - ~ + Z +Vz (66)


ktgo ~go

a71 = 0-1 + ~llB2x a72 = cY7Bz + o l l B x B y (67)

a73 = -(YTBy + O'llBxB z (68)

NBp ) /9"I'
a74 = t37(VyBy + VzBz) + Oll(ExB x + + O'I2(NBT +B x (69)

a75 = - 0 " 1v z - O"7(Pz + epvxB )--0"9 0T +~11Ey B + Bx 0I' (70)

0"7 0T 0T
a76 =O,Vy + - s - ( P y - g p v x B z ) + 0 9 ~ + Ol,EzBx +Ol2B x (71)
~p vy 0z

Matrices BEM and CEM may be obtained in the same fashion as equation (61).
After tedious algebraic manipulations [26], the vector of eigenvalues of the
flux vector Jacobian matrix A s s is found as

- {
k~M = 0, Zk, k~, 0, Z~,, k~,, Vx
}" (72)

This means that the eigenvalues ~1 "- ~4 - - 0 , while ~'7 -- Vx" The remaining
four eigenvalues can be obtained from the fourth order algebraic equation

~4 "t- (/,EM ~3 + VEM~,2 + ~EM ~ + 6EM -" 0 (73)

where the coefficients in the fourth order characteristic polynomial are

~EM ------a22 -- a33 (74)


695

VEM "- a22a33 - a26 + a35 (75)

~EM -- a26a33 - a22a35 (76)

~EM -- a25a36 -- a35a26 (77)

The four eigenvalues are the analytical roots given as

+ 10 fil 1 (78)
~E=--4 " EMI+-vi-602MI---2 (II/EM-I-nEM~

19 ~/'1 1 (79)
~'E=--~ EMI-- "i-~O2EMI--~(II/EM'F~-)EMo)

+ 1 ~1 1
~:~ - -~-r + igc, G - -~(v~,,, - a~.,,,o) (80)

i
~'B -- -- ~ (I)EM2 --
jm~ 0 2 M 2 -- -~1 (ll/EM -- ~"~EMo) (81)

Here, different terms are defined as

(I)EMI -- (gEM + ~/ (I'EM


2 _ 4 V E M + 4~EM = O~EM + (I)EMo (82)

2
(I) EM2 = ~ EM -- (g EM - 4 v EM + 4~4/EM = (X,EM --OEM o (83)

(I)EMo = 4Ot,2M --4VEM + 4q/EM (84)

~"~EMo = ((~EMVEM -- 2~EM)/OEMo (85)

(86)
3

3(aEMYEM_ 48EM) _ VEM


2 (87)
ZEM =
696

[
3
YEM = VEM (4~EM -- (XEM~tEM) I VEM xl4VEM6EM-- ~tEM
2 -- (XEM~EM)
2 (88)
6 27 2

For illustrative purposes, the following are the eigenvalues in the case of one-
dimensional EMHD flow where Vy - v , - 0 and a22 = a33 and a25 = 0. Hence

~-~-
;[E IE2 / 1
x Vx+ x Vx+4 ~;o~to(1 + X M ) -- ~EVx
2)] (89)

~E--~B--~ " X Vx-- xE2Vx +4 --Z Vx (90)


eo~to(1 + )

1
Since /~o~-~o equals the speed of light in vacuum, it seems that for most
practical applications the incoming and the outgoing electromagnetic waves will
not be influenced by the fluid except in the situations where the fluid is very
highly ionized or when the fluid moves with a speed comparable to the speed of
light. In the case of a pure electro-magnetics without any fluid motion,
polarization, magnetization, or electric charges (v = __P= M =qo =0), these
eigenvalues reduce to the eigenvalues of Maxwell's equations for electro-
magnetic fields in vacuum [35]

1 1 1 1 }*
(91)
- 0, 4 ' - S' 0, 4 ' -

After introducing the similarity transformation matrix S__EMof the flux vector

Jacobian matrix ~ r M ' the eigenmatrix ~EM corresponding to ~EM becomes

~" +
~EM diag[ 0, ~E, LE,0, L+B,L-B, v x ]
- -
(92)

where )~E, )~E, )~a, )~a are given by equations (78-81).


For locally one-dimensional problems, wave propagation direction is well
defined. For multi-dimensional problems, there is no unique direction of
697

propagation, because the flux vector Jacobian matrices AEM , BEM , CEM cannot
be simultaneously diagonalized. Therefore, characteristic boundary condition
analysis allows that only one of these matrices (relating to only one coordinate
direction) can be diagonalized at a time.
In the case that the x-coordinate is in the main flow direction, premultiplying
the equation (58) with the inverse of the similarity matrix, SEI~, gives

SE1M~QEM '~ -1 ~QEM -I ~'


~-t- ~EMSEM ~-t- SEMHEM -- 0 (93)
Ot = c~x

Here, vector HEM is given as

H r u = BEta 0QEra + CwM aQEM SEM (94)


--- ~ = aZ

For the hyperbolic system, time dependent boundary conditions could be


derived based on the principle that outgoing waves are described by
characteristic equations, while the incoming waves may often be specified by a
non-reflecting boundary condition [31,32,36]. Following this approach, the
characteristic and non-reflecting boundary conditions at the inlet boundary x = a
and at the outlet boundary x = b can be given by the i-th equation of the system
-1
(93). Here, the left eigenvector Si,EMis the i-th row of S ~

I=i,EM -O~ + Li,EM + Si,EMHEM) x=a,b - 0


S-1 C3QEM (95)

where L~,EM- 0 for incoming waves, while for outgoing waves

(96)
, 63x
698

7.2 Characteristic-based boundary conditions for Navier-Stokes subsystem


Similar derivations can be used to determine analytical expressions for the
eigenvalues and the non-reflecting boundary conditions of the Navier-Stokes
subsystem of the unified EMHD as shown by Dulikravich and Jing [26].
Characteristic treatment of the Navier-Stokes subsystem of the unified
EMHD system can be performed by converting its conservative form

63QNs + aENs
- -b aFN S -1 63GNs = SNS (97)
at ax as az
into its non-conservative (characteristic) form

g31~NS t- ANS 691~NS I- BNS C3QN----------~s


+ CNS g31~N---------~s
= SNS (98)
& 0x 0y -- 0z

where the solution vector of unknowns is given as

-
QNS= P/~,
{
Vx, Vy, Vz, e
}* (99)

From equation (47) it can be seen that flux vector ENs becomes

Vx ~,~1
2 p % __I'~BM Vx~x /
Vx + /

P P P PB p |
VxVy
'l~xy VyNx / (100)
ENS=
~xz Vz x
VxVz !|

ev_~ N_~
'~
P P J

Terms related to d,d_ 2 and VT will not be considered in the evaluation of


coefficients of the flux vector Jacobian matrix ANS since they are associated
with first derivatives of velocity, v, or temperature, T. The flux vector Jacobian
matrix ANS = c3ENs/0QNs then becomes
699

0 1 0 0 0
~[p a22 a23 a24 0
ANS - 0 a32 a33 a34 0 (101)
0 a42 a43 a44 0
_Vx[5/9 a52 a53 a54 Vx_

The coefficients in this matrix are given in detail by Dulikravich and Jing [26].
Eigenvalue vector of the flux vector Jacobian matrix ANs is

- {
~,NS = Vx, Uu, Z+v, Uw, Ue
}" (102)

which can be written as a diagonal eigenvalue matrix

~'Ns== diag[Vx' Z+~' Uv' Uw' U~] (103)

The eigenvalues Uu, X~, Uw, Ue are obtained analytically by solving a fourth
order characteristic polynomial (similar to equation 73) where

~ = - ( a ~ +a3~ +a,4) (104)

VNS - - a22a33 + a22a44 + a33a44 - a34a43 -- a24a42 -- a23a32 (105)

)'NS -- a34a43a22 - a22a33a44 - a24a32a43 + a24a33a42 + a23a32a44


(106)
- a23aa4a42 + (a33 + a44)13
P

~NS -- (a34a43
_
a33a44
)_~ (107)
9

so that the four eigenvalues are

~+-" 4 NS1 -I- ONS1 ---2 "(ll/NS q- ~'~NS~) (108)


700

--'4"(I)Ns1 -- NSI -- ~/NS q- ~'-~NSo) (109)

1 ~/1 2 1
~+w -- -- 4 (I) Ns2 q" i--~(I)Ns2 -- 2 (V NS -- ~'~NSo) (110)

_lt9 ~1 2
)~+e- 4 NS2-- ~ ~Ns2 --2(~I/NS- ~'-~NSo) (111)

with the coefficients given by equations of the type similar to equations (82-88).
Characteristic waves defined by the Navier-Stokes equations in the EMHD
system have a great dependency on both fluid dynamics and electro-magneto-
dynamics, in particular, the electro-magnetic properties of the media and
electro-magnetic field quantities. When electric and magnetic fields are absent,
these eigenvalues reduce to the well-known eigenvalues of a classical Navier-
Stokes system for Newtonian, incompressible flows. These eigenvalues are
{Vx,V~VxV x + c, Vx - c }. Here, the equivalent local speed of sound is defined
as C - - 4 V 2 + (~/p)
X "

Following Thompson's approach [30,31], non-reflecting boundary conditions


for the Navier-Stokes subsystem are hence formulated as follows. The
characteristic form of Navier-Stokes subsystem influenced by the electro-
magnetic effects is possible to write as

~ aQNs -1 ~
SN s r + ~NsSNIs ~ + S__NsHNs=0 (112)
& = 0x

where the i-th equation is

=,,NS & + NSSN s OQNs


S_ 1 ~QN_______SS & + Si,NsHNs _ 0 (1 13)

and the new source vector is

HNs - BNS aQN-----~s+ CNS 0QN------As- SNs (114)


Oy = Oz
701

Here, the left eigenvector S--i,NS


"1 is the i-th row of S~ls

Ot --1--/I
Si.NS O0NS + Li,NS + Si,NsHNs
x=a,b
-- 0 (115)

where Li,NS --0 for incoming waves, while for outgoing waves

,NS 0X (116)

Practical implementation of Thompson-type [31-33,36,26] non-reflecting


boundary conditions deserves further comments. The essence of his approach is
that one-dimensional characteristic analysis can be performed by considering
the transverse terms as a constant source term. In order to provide well-posed
non-reflecting boundary conditions in multi-dimensional cases, substantial
modifications may be required to take into account the transverse terms at the
boundaries [37,38]. It should be emphasized that physically there are cases
where flow information propagates back from the outside of the domain into the
inside through the boundaries by the incoming waves [39]. This fact makes it
possible that building a perfectly non-reflecting (absorbing) boundary condition
[40] might lead to an ill-posed problem. Under these circumstances, corrections
may be needed to make them partially non-reflecting.

7.3 Numerical integration of EMHD system


It is often highly desirable to have a time-accurate unsteady solution to the
governing EMHD equations. One numerical integration algorithm that could be
used is an advanced form of the dual time-stepping technique, also called an
iterative-implicit technique, originally developed by Jameson [41 ].
To create an instantaneous picture of the solution of the entire EMHD
system at a given physical time, equation (43) must be driven to zero in its
entirety, not, as is commonly done in time-marching techniques by driving only
the physical time-dependent term to zero. To this end, a pseudo-time derivative
is added to the EMHD system (equation 43) which can be rewritten as

c3Q 0E 0F 0G --
00 + ~ + ~ + ~ + ~ _ S (117)
c3z 0 t 0x 0y 0z
702

or as

aQ= aQ
_ & (118)

where ~ is a composite of the spatial and source terms and is called the
residual. Thus, given a physical time step the governing equations are time
marched in pseudo time, x. Upon convergence, the fight-hand side of equation
(118) becomes zero and the solution at the desired physical time level, t, is
obtained. Note that the pseudo-time dependent variable vector, 0 , does not
have to be the same as the physical time dependent variable vector, r
An additional concern of great importance is that the system of equations
develops zero terms in the pseudo-time dependent variable vector, 0 , for
incompressible fluids, fluids without electric charges, or systems in which the
electric and magnetic fields are non-interacting. This poses significant
problems for time-marching numerical solutions. This problem may be
alleviated, however, by proper selection of pseudo-time dependent variable
vector, 0 , and through the use of matrix preconditioning.
By premultiplying I~ with a properly selected matrix, it is possible to
directly control the system eigenvalues. This prevents development of zeros in
the pseudo-time dependent variable vector, 0 , and vastly improves iterative
convergence rates over a wide variety of flow regimes (low and high Mach and
Reynolds number combinations). The preconditioning matrix, F'(0), for the
EMHD system could be based on one developed by Merkle and Choi [42] for
the Navier-Stokes system. The preconditioned EMHD system may be written as

r, aQ_
= ax -- aQ
& (119)

Equation (119) can be transformed to a body-conforming non-orthogonal


curvilinear time-dependent (~, 1-1, ~; t) coordinate system. A high order of
accuracy is desired to properly resolve unsteady motions. A finite difference
scheme using fourth order accurate spatial differencing and second order
accurate physical time differencing could be used while the solution is advanced
in pseudo-time using a four-stage Runge-Kutta scheme which is second order
accurate for non-linear problems. Fourth order accuracy should be selected for
703

the spatial derivatives based on extensive research completed by Carpenter et al.


[30] which found that a Runge-Kutta advanced fourth order accurate scheme
provided the best convergence and stability of higher order schemes at
reasonable computational cost. Second order accurate differencing in physical
time could be selected based on stability and convergence studies performed by
Melson et al. [43] who found that for a Runge-Kutta advanced dual time-
stepping scheme second order backward differencing provided the most stable
physical time discretization while providing excellent resolution. The new
physical time step could be treated implicitly in pseudo-time, while all old
physical time steps and spatial derivatives could be treated explicitly. This is
unlike Jameson's early method [41] that treats both the physical time and the
spatial derivative explicitly and causes a restriction on the maximum physical
time step allowed. The discretized preconditioned system may be written as

0 ~ =(~" (120)

A'I~ "~/~ i
F'F-I+ a i -~ _
- ~ z- ~- r i m + l , i-1

_4,
m2,,t i=O (121)

(~ n-,-1= 0 4 (122)

where m=1,2,3,.., represents the physical time step, n=1,2,3,.., represents the
pseudo-time step, and i=1,2,3,4 is the Runge-Kutta stage number. Also,
F = 0 0 / / ~ and c~i are the Runge-Kutta coefficients. Note that the physical
time-dependent term on the right hand side of equation (121) is held constant
for all four Runge-Kutta stages.

8. S U B M O D E L S OF EMHD

Until now, the numerical solutions of the unsteady three-dimensional EMHD


flows that have been reported in the open literature [34-36] did not account for
polarization or magnetization effects and did not involve charge density
transport equation. The reason is that the complete unified EMHD system is
very large having extremely complicated source terms and two extremely
704

different time scales for the electro-magnetic fields and the flow-field.
Consequently, a number of simplified versions of the EMHD system have been
traditionally used in practical applications. These submodels can be grouped in
two general categories: EHD models and MHD models [ 11-13,44].
From the unified EMHD model, it can be seen that the electromagnetic field
is not the only cause of electric current and that the temperature gradient is not
the only source of heat conduction as is commonly assumed. The electric field,
magnetic field, heat conduction, and deformation (strain) may couple to produce
charge motion and heat transfer. These couplings are called phenomenological
cross effects and may be placed in four general categories: 1) thermoelectric, 2)
galvanomagnetic, 3) thermomagnetic, and 4) second order effects [9, p.161-
163]. These categories are based on the source of the effect and each will be
described in turn, as will be a comparison between classical EHD and MHD
models and the unified EMHD theory. The comparison concentrates on
similarities and differences between electro-magnetic force and electric current
and heat conduction terms in the EHD, MHD, and EMHD models. The
inadequacies of simple superpositioning of classical simplified models to fully
describe the unified EMHD flows are also noted.
Couplings between the temperature gradient and the electric field cause
thermoelectric effects so that a temperature gradient in the material produces an
electric current (Thompson effect), while applied electric field produces heat
conduction in the material (Peltier effect). These two effects together are
known as the Seebeck effect and form the basis for thermocouples. Also note
that the or, term in the electric conduction current (equation 22) and the K4 term
in the heat conduction (equation 23) are the ohmic charge conduction and
Fourier heat transfer, respectively.
When the electric and magnetic fields are simultaneously applied but are not
parallel, electric current (Hall effect) and heat conduction (Ettingshausen effect)
perpendicular to the plane containing the electric and the magnetic fields are
induced in the media. These effects are termed galvanomagnetic [9, p. 161-163].
When the temperature gradient and the magnetic field are simultaneously
applied but are not parallel, electric current (Nernst effect) and heat conduction
(Righi-LeDuc effect) perpendicular to the plane containing the temperature
gradient and the magnetic field are induced in the material. These effects are
termed thermomagnetic.
It should be noticed (equation 22) that the interaction of the average rate of
deformation tensor and the electric field can also create the electric current,
while the interaction of the material deformation tensor and the electric field can
create the temperature gradient (equation 23). These piezo-electric and piezo-
magnetic effects can further be enhanced if the material is non-isotropic.
705

8.1 Classical e l e c t r o - h y d r o d y n a m i c s ( E H D )
As mentioned previously, EHD flows are those in which magnetic effects
may be neglected and charged particles are present, while only a quasi-static
electric field is applied so that the magnetic field, both applied and induced,
may be neglected [ 11 ]. One of the implied assumptions is that the flows are at
non-relativistic speeds, although in astrophysical flows this assumption cannot
be made [1]. Atten and Moreau [44] present a detailed coverage of classical
EHD modeling and discuss the relative importance of terms in the force and
electric current through stability analysis. With these assumptions, the
Maxwell' s system reduces to [ 11 ]

V . D - V.(~E_)- qo (123)

c3q~ ~ - V . , J - 0 (124)
&

With classical EHD assumptions, the electro-magnetic force in the unified


EMHD theory reduces to"

f_.v_,M_ q~E + (V_E)-P = qo_E+ (VE). e,pE (125)

This is not the form of the electro-magnetic force usually seen in classical EHD
formulations [11]. Through the use of thermodynamics and the material
constitutive equation of state, the electric force per unit volume in EHD is most
often used in the following equivalent forms [10, p.505-507][8, p.59-63]

f EM : qo_E- E . E Ve + - V E-E p (126)


- --2- 2 - T=const

- 2 p=const constl (127)

The three terms in the equation are the electrophoretic, dielectrophoretic and
electrostrictive terms, respectively.
The electrophoretic force or Coulomb force is caused by the electric field
acting on free charges in the fluid. It is an irrotational force except when charge
gradients are present [45].
706

The dielectrophoretic force is also a translational force, but is caused by


polarization of the fluid and particles in the fluid. The dielectrophoretic force
will occur where high gradients of electric permittivity are present. This
condition will be true in high temperature gradient flows, multi-constituent
flows, particulate flows [ 18] or any time the electric field must pass through two
contacting media of different permittivities [46]. Grassi and DiMarco [47] treat
the dielectrophoretic force as it applies to bubbly flows and heat transfer.
Poulter and Allen [45] note that the dielectrophoretic force produces greatest
circulation when the dielectric permittivity is inhomogeneous and non-parallel
with the applied electric field.
The last force, the electrostrictive force, is a distortive force (as opposed to
the previous translational forces) associated with fluid compression and shear.
The electrostrictive force is usually smaller than the -phoretic forces. It is
present in high pressure gradient flows, compressible flows, and flows with a
non-uniform applied electric field. Pohl [18] describes this phenomenon in
greater detail.
Classical EHD modeling derives directly from the unified EMHD theory.
Thus, the electric current density, using EHD assumptions, reduces to

J = qo_V+ OlE + o-47T (128)

However, this is not the form seen in classical EHD models [ 11 ] which typically
define the conduction electric current as only the first term of equation (22).
However, more advanced classical EHD models define the current as [9, p.562]

J- = qo_v + J_c = qo_v + qob E_- DoVqo (129)

The last two equations imply that the temperature gradient is directly related to
the electric charge gradient. This may be shown to be true based on the
Einstein-Fokker relationships, derived from studies of Brownian motion [25,
p.264-273], which relate any concentration gradient to a charge mobility and a
diffusion. Newman [48] also provides a detailed discussion of the concepts of
diffusion and mobility. The electric charge diffusion term is often neglected
where only limited amount of free charges are available [49].
By introducing classical EHD assumptions in the unified EMHD theory, the
equation (23) for heat flux reduces to

CI = KI__E+ K4VT (130)


707

The classical EHD models neglect the contribution to heat transfer from the
electric field so that equation (130) reduces to Fourier's law of heat conduction.

Cl = -~:VT (131)

Although classical EHD modeling seems to neglects heat transfer induced by


the electric field and electric current, Joule heating effect ( - I , . _ E term from
EMHD equation 19) is usually included in the EHD computations [50,51 ].

8.2 Classical m a g n e t o - h y d r o d y n a m i c s (MHD)


The classical modeling of MHD assumes non-relativistic and quasi-
magnetostatic conditions. It implies that electric current comes primarily from
conductive means and that there are no free electric charges in the fluid [11 ].
With these assumptions Maxwell's system becomes

V.B_.= 0 (132)

0B
VxE= -- (133)
- 0t

Vxl-l= / (134)

V-J -0 (135)

The modifications to the Navier-Stokes relations come from the electro-


magnetic force on the fluid from which all induced electric field terms have
been neglected. Using the MHD assumptions, the electro-magnetic force per
unit volume in the unified EMHD theory becomes [ 11 ]

[EM = j x B_B+_ ( V B ) . M___ (136)

The second term, source of dimagnetophoretic and magnetostrictive forces, is


typically neglected in classical MHD [10, p.508]. Thus, the electro-magnetic
force per unit volume in the classical MHD is modeled as [ 11 ]

frM = j x B (137)
708

By making MHD assumptions, the conduction current in the EMHD can be


expressed with equation (38). However, classical MHD theory usually defines
the electric conduction current as [ 10, p.510]

Jc = cYlE + o4VT = cYlE + O'l(V x B) + cY4VT (138)

Here, (5'4 is the Seebeck coefficient [9, p.174] which in some classical MHD
formulations is not used [11 ]. Clearly, the classical MHD formulations neglect
a significant number of physical effects [52,53].
Similarly, in classical MHD modeling, Joule heating is often included in the
energy relation, but the heat transfer constitutive relation remains the same as in
equation (131). In comparison, the unified EMHD model for the heat flux with
classical MHD assumptions can be expressed with equation (39).
It could be concluded that classical EHD models include many important
effects and correspond to the unified EMHD theory well, while classical MHD
formulations need improvements in the force, current, and heat transfer terms.
As in classical EHD modeling, it is important to be aware of the fact that
many force, current and heat transfer terms can be written in several different
forms, each of which is equivalent. It is, therefore, important to recognize the
potential danger of simply adding terms from different EHD and MHD models.

9. S O L I D I F I C A T I O N W I T H E L E C T R O - M A G N E T I C FIELDS

During solidification from a melt, if the control of melt motion is performed


exclusively via an externally applied variable temperature field, it will take
quite a long time for the thermal front to propagate throughout the melt thus
eventually causing local melt density variations and altering the thermal
buoyancy forces. It has been well known that an externally applied steady
magnetic or electric field can, practically instantaneously, influence the flow-
field vorticity and change the flow pattern in an electrically conducting fluid
[51-59,33]. Similarly, it is well-known that applying an electric potential
difference to a flow-field of a homogeneous mixture will cause fractionation or
separation of the homogeneous mixture into regions having high concentration
of the constituents. This phenomena, known as flee-flow electrophoresis, has
been extensively studied experimentally and, to a lesser extent, numerically [50]
using classical EHD modeling. Nevertheless, there are no publications yet on
actual algorithms for determining the proper variation of intensity and
orientation of the externally applied magnetic and electric fields. This is not a
709

trivial problem because we are dealing with a moving electrically conducting


fluid within which an electric current is induced as the fluid cuts through the
externally applied magnetic field lines [11]. This induced electric current
generates heat (Joule effect) as it passes through the fluid that has a finite
electrical resistivity. In the case of solidification, the amount of heat generated
through the Joule effect due to the externally applied magnetic field is often
neglected compared to the latent heat of solidification and the amount of heat
transferred in the melt by thermal conduction.
The latent heat released or absorbed per unit mass of mushy region (where
Tliquidus > T > Tsolidu s ) is proportional to the local volumetric liquid/(liquid +
solid) ratio often modeled [59] as

f = Ve =
( T - Tsolidus
/n = 0n (139)
v,+v, Wliquidus -- Wsolidus

Here, 0 is the non-dimensional temperature, the exponent n is typically 0.2 < n


< 5, subscripts g and s designate liquid and solid phases, respectively, while f -
1 for T > Tliquidusand f - 0 for T <Tsolidus. Physical properties (density,
viscosity, heat conductivity, heat capacity, etc.) are often significantly different
in the melt as compared to the solid phase. We may assume linear variation of
density as a function of the non-dimensional temperature, 0, in the liquid

p e - p ~ 1+ . . . . (140)
00

with a similar expression for the solid phase where the reference values are
designated with the subscript "r". In this work, we assumed that electric
conductivity and magnetic permeability do not vary with temperature.
The EHD and the MHD systems of equations including solidification can be
non-dimensionalized in a number of ways. The typical non-dimensional
numbers are [33,60]:

Reynolds hydrodynamic Froude Eckert


_ ~ O
V__~r
2
R ~ - PrVrg r FR2= Vr Ec= (141)
~vr grgr CrATr
710

Prandtl hydrodynamic Stefan Grashof


PR = ~tv~Cr STE = CraT~ a R = P~ff'rgrATrg3r (142)
g2
Kr Lr vr

Hartmann Prandtl magnetic Prandtl electric

HT = grgrHr Or / )
\ l'tvr
Pm - gvrO'r~'l'r
Pr
~tvr
PrbrA~)r
(143)

Coulomb Electric field Charge diffusivity


SE = qorA~)r q org2r ~vr (144)
PrV2 N~ - 8rAter D E = PrDor

where ~tv~,c~,A,l,r,~:~,~q,Lr,g ~ are the reference values of viscosity, specific


heat, electric potential difference, heat conductivity, magnetic permeability,
latent heat of liquid-solid phase change, and length, respectively. Also, mixture
density and modified heat capacity can be defined as

Pmix = f Pt + (1- f)Ps (145)

~(Cg 0g) ~(Ceq 0s)


Cmix = fPe ~ + (1 -- f)p~ (146)
00 30

An enthalpy method [58,59] can be used to formulate the equivalent specific


heat coefficient in the solid phase defined as

1 0L
c~q = c~ (147)
STZ 00

so that latent heat is released in the mushy region according to equation (139).

9.1 EttD and solidification


EHD equations for phase-changing liquid-solid mixtures, where the solid
phase is treated as the second liquid with extremely high viscosity, can be
derived using Boussinesq approximation for thermal buoyancy [61]. We can
also define mixture electric charge mobility
711

bmi x = f b e + (1 - f)b s (148)

and combined hydrodynamic and hydrostatic pressures in liquid and solid

Pe-P+q~ and Ps=P+q9 (149)


o,~ F~ Os F~

where q) is the non-dimensional gravity potential defined as g =-Vq~.


Assuming equal velocities for both phases, the mass conservation is

V-v=O (150)

Linear momentum conservation for two-phase EHD flows with thermal


buoyancy and Coulomb force

Pmix--~ + fpgV-(VV -I-ps (1- f)ps v" (VV q- Ps!)

= f V I_~e -- -- +-~-e2PeoteOg_ (151)

{
+ ( l - f ) V Fgvs(Vv+(Vv) * +
[Re - - ~
>]~ } PsO,.s0
g-
+SEqo.

Energy conservation for incompressible two-phase EHD flows including Joule


heating can be written as [60]

O0 [ \ [_ \
Cmix
_ 1
- .Re----PT~If V" (KtVO)+ (1- f) V 9(KsVO)] (152)

+Sc(qov+qo,mix_
Electric charge conservation equation including migration and diffusion is
712

Ot
E( bmix)] 1
~V. qo v+
- ReP E
_E - ~
ReD E
V 9( b m i x V q o ) (153)

Since _E_E--V~, the electric potential equation resulting from equation (13)

V-[(fe e + (1- f)~Zs)V~] = -NEq o (154)

must be solved simultaneously with the equations (150-153).

9.2 M H D and solidification


MHD two-phase solid-liquid flows can be modeled using a similar approach.
The non-dimensional Navier-Stokes equations for phase-changing mixtures of
two liquids (solid phase is treated as the second liquid with extremely high
viscosity), can be formulated [33] so that the mixture mass conservation is

V-_v=O (155)

Linear momentum conservation for two-phase MHD flows with thermal


buoyancy and magnetic force

63v
Pmix-~+ f P e V " (---VX+ PeI)+ (1- f)P~V "(XX+ PsI)

=f V- -~(Vv+(Vv)* + eae0_+ (156)


Ke -- -- -~e2P g PmR~' e

+(,-o {-v [.vs(vv+(vv; tl+ ~


LR e - _ Re
o ~+~.,,(~215
- PmRe
)
The non-dimensional hydrodynamic, hydrostatic, and magnetic pressures were
combined to give

pc-P +~-+ {.teH.H and ~ P +~-y+ g~H-H (157)


Pe FR PmR2e Ps FR PmRo
2
713

where tp is the non-dimensional gravity potential defined as g - - V q ~ . Then,


the energy conservation for incompressible two-phase MHD flows including
Joule heating can be written as [33]

00 [ \
Cmix + fpt?V" tcf0__v)+ ( 1 - f)ps v . [csq0v)

1 1 HTE c (V x H)-(V x H
=f RePR V 9(Ks + 2 3 (158)
O"e PmRe

1 1 HTE c tV
+ ( l - f ) RePR V'(K:sV0)q--- 2 3 I-I).(VxI-I
o s VmRe

The magnetic field transport equation for the two-phase MHD flow in its non-
dimensional form becomes [ 1, p. 150]

-pm~V x +~ VxH (159)


e O's~s

If electric conductivity and magnetic permeability are assumed constant, then

_ v (_v H ) - (1- f)/(~sgs)V2H (160)


tgt PmRe --

needs to be solved either intermittently [33] with the equations (155-158).

ACKNOWLEDGMENTS

The author would like to thank Dr. Yimin Ruan and Dr. Owen Richmond of
the ALCOA Technical Center for the ALCOA Foundation Grant, Dr. Martin
Volz of Microgravity Program at NASA Marshall Space Flight Center for
partially supporting a student assistant, Professor Akhlesh Lakhtakia of the
Pennsylvania State University for stimulating technical discussions, and Mrs.
Sheila Corl and Professor Hyung-Jong Ko from Kumoh National University of
Technology, Korea for proofreading this chapter.
714

REFERENCES

1. W.F. Hughes and F.J. Young, The Electromagnetodynamics of Fluids,


John Wiley and Sons, New York (1966).
2. G.S. Dulikravich and S.R. Lynn, ASME FED-Vol. 235, MD-Vol. 71
(1995) p.49.
3. G.S. Dulikravich and S.R. Lynn, ASME FED-Vol. 235, MD-Vol. 71
(1995) p.59.
4. G.S. Dulikravich and S.R. Lylm, International Journal of Non-linear
Mechanics, Vol. 32, No. 5 (September 1997) p.913.
5. G.S. Dufikravich and S.R. Lynn, International Journal of Non-linear
Mechanics, Vol. 32, No. 5 (September 1997) p.923.
6. G.S. Dulikravich and Y.-H. Jing, ASME AMD- Vol. 217 (1996) p.309.
7. S.-I. Pai, Magnetogasdynamics and Plasma Dynamics, Springer-Verlag,
ViennaJPrentice Hall, Inc., Englewood Cliffs, N.J. (1963).
8. L.D. Landau and E.M. Lifshitz, Electrodynamics of Continuous Media,
Pergamon Press, New York (1960).
9. A.C. Eringen and G.A. Maugin, Electrodynamics of Continua I;
Foundations and Solid Media, Springer-Verlag, New York (1990).
10. A.C. Eringen and G.A. Maugin, Electrodynamics of Continua II; Fluids
and Complex Media, Springer-Verlag, New York, (1990).
11. O.M. Stuetzer, Physics of Fluids, Vol. 5, No. 5 (1962) p.534.
12. J.R. Melcher, Continuum Electromechanics, MIT Press, Cambridge, MA
(1981).
13. G.W. Sutton and A. Sherman, Engineering Magnetohydrodynamics,
McGraw Hill, New York (1965).
14. A.S. Wineman and K.R. Rajagopal, Continuum Mech. Thermodyn., Vol. 7
(1995) p.1.
15. P.G. Bergman, The Special Theory of Relativity, Handbuch der Physik,
Bd. IV, Springer-Verlag, Berlin (1962).
16. M.J. Marcinkowski, Acta Physica Polonica, A.81 (1992) p.543.
17. A. Lakhtakia, J. of Advances in Chemical Physics, Vol. 85 (1993) p.311.
18. H.A. Pohl, Dielectrophoresis; The Behavior of Neutral Matter in
Nonuniform Electric Fields, Cambridge University Press, Cambridge, U.K.
(1978).
19. C.T.A. Johnk, Engineering Electromagnetic Fields and Waves, John Wiley
and Sons, New York (1988).
20. W.N. Cottingham and D.A. Greenwood, Electricity and Magnetism,
Cambridge University Press, Cambridge (1991).
715

21. H.A. Haus and J.R. Melcher, Electromagnetic Fields and Energy, Prentice
Hall, New Jersey (1989).
22. A.C. Eringen, Mechanics of Continua, 2nd Ed., Robert E. Krieger
Publishing Co., Malabar, FL (1967).
23. A.C. Eringen, Ed., Continuum Physics-Volume II; Continuum Mechanics
of Single-Substance Bodies, Academic Press Inc., New York (1975).
24. R.M. Bowen, Introduction to Continuum Mechanics for Engineers,
Plenum Press, New York (1989).
25. S.R. de Groot and P. Mazur, Non-Equilibrium Thermodynamics, North
Holland Publishing Company, Amsterdam (1962).
26. G.S. Dulikravich and Y.-H. Jing, ASME IMECE, Dallas, TX (Nov. 1997).
27. A.J. Chorin, Journal of Computational Physics, Vol. 2 (1967) p. 12.
28. T.B.A. Senior and J.L. Volakis, Aproximate Boundary Conditions in
Electromagnetics, lEE, London, UK (1995).
29. C. Hirsch, Numerical Computation of Internal and External Flows,
Volume 1; Fundamentals of Numerical Discretization, Wiley-Interscience,
New York (1988).
30. M.H. Carpenter, D. Gottlieb and S. Abarbanel, J. Comp. Phys., Vol. 108
(1993) p.272.
31. K.W. Thompson, Journal of Computational Physics, Vol. 68 (1987) p. 1.
32. K.W. Thompson, Journal of Computational Physics, Vol. 89 (1990) p.439.
33. G.S. Dulikravich, V. Ahuja and S. Lee, International Journal of Heat and
Mass Transfer, Vol. 37, No. 5 (1994) p. 837.
34. V.J. Shankar, W.F. Hall and H.M. Alireza, Proc. IEEE, Vol. 77, No. 5
(May 1989) p.709.
35. J.S. Shang, AIAA paper 91-0606, Aerospace Sciences Meeting, Reno, NV
(January 1991).
36. M.S. Tun, S.T.Wu and M. Dryer, J. Comp. Phys., Vol. 116 (1995) p.330.
37. T.J. Poinsot and S.K. Lele, J. of Comp. Physics, Vol. 101 (1992) p. 104.
38. R. Hixon and S.-H. Shih, AIAA paper, 95-0160, Reno, NV (1995).
39. T. Hagstrom and S.I. Hariharan, Math. Comput., Vol. 20, No. 10 (1994)
p.155.
40. M.E. Hayder, F.Q. Hu and M.Y. Hussaini, ICASE Report No. 97-25 (also
NASA CR 201689) (May 1997)
41. A. Jameson, AIAA Paper 91-1596, Reno, NV (January 1991).
42. C.L. Merkle and Y. Choi, Intl. J. Num. Meth. Eng., Vol. 25 (1988) p.293.
43. N.D. Melson, M.D. Sanetrik, and H.L. Atkins, 6th Copper Mountain Conf.
on Multigrid Methods, Copper Mountain, CO (April 4-9, 1993).
716

44. P. Atten and R. Moreau, Journal de Mecanique, Vol. 11, No. 3, September
(1972) p.471.
45. R. Poulter and P.H.G. Allen, 8th Intl. Heat Transfer Conf., San Francisco,
CA, Vol. 6 (1986) p.2963.
46. M. Aoyama, T. Oda, M. Ogihara, Y. Ikegami and S. Mashuda, Journal of
Electrostatics, Vol. 30 (1993) p.247.
47. W. Grassi and P. Di Marco, VII European Symp. on Material and Fluid
Science in Microgravity, Belgium (1991).
48. J.S. Newman, Electrochemical systems, Prentice Hall, NJ (1991).
49. R.B. Schilling and H. Schaechter, Journal of Applied Physics, Vol. 38
(1967) p.841.
50. S. Lee, G.S. Dulikravich and B. Kosovic, AIAA paper 91-1469, AIAA
Fluid, Plasma Dynamics and Lasers Conf., Honolulu, HI (June 1991).
51. S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, Dover
Publication, Inc., New York (1961).
52. M. Salcudean and P. Sabhapathy, ASME MD-Vol. 20, ASME Book No.
G00552 (1990) p.115.
53. H. Ozoe and K. Okada, International Journal of Heat and Mass Transfer,
Vol. 32, No. 2 (1989) p.1939.
54. S. Lee and G.S. Dulikravich, International Journal for Numerical Methods
in Fluids, Vol. 13, No. 8 (October 1991) p. 917.
55. G.S. Dulikravich, V. Ahuja and S. Lee, Journal of Enhanced Heat
Transfer, Vol. 1, No. 1 (August 1993) p. 115.
56. G.S. Dulikravich, K.-Y. Choi and S. Lee, ASME FED-Vol. 205/AMD-
Vol. 190 (1994) p. 125.
57. H. Hatta and S. Yamashita, Journal of Composite Materials, Vol. 22 (May
1988) p.484.
58. G.S. Dulikravich, B. Kosovic and S. Lee, ASME Journal of Heat Transfer,
Vol. 115 (February 1993) p.255.
59. G.S. Dulikravich, V. Ahuja and S. Lee, Numerical Heat Transfer:
Fundamentals, Part B, Vol. 25, No. 3 (1994) p.357.
60. V.R. Voller and C.R. Swaminathan, Numerical Heat Transfer, Part B,
Vol.19 (1991) p.175.
61. D.D. Gray and A. Giorgini, International Journal of Heat and Mass
Transfer, Vol. 19 (1976) p.545.
717

C O N D U C T I O N AND DIELECTRIC E F F E C T S
IN E L E C T R O R H E O L O G Y

C.W. Wu* and H a n s C o n r a d

Department of Materials Science and Engineering


North Carolina State University
Raleigh, NC 27695

1. I N T R O D U C T I O N

Since Winslow's discovery [1,2] of electrorheology (ER) the subject


has developed into a multidisciplinary field that has captured the
imagination and attention of scientists and engineers worldwide. Early
studies [3,4] on the interaction force between two adjacent particles in
an ER suspension employed the point-dipole approximation, assuming
no electric current flows through the suspension, i.e., both the particles
and host liquid are non-conducting. This gives for the interaction force

f =24na6aoKf(~dEo)2/R4 , (1)

where a is the radius of the particles, ~o the permittivity of vacuum


space, Kf=~f&o the real component of the relative permittivity of the
host fluid, R the distance between the particle centers, E o the applied
field and ~d the polarizability induced by the particle/liquid dielectric
mismatch, which is given by

(2)

* Visiting professor, on leave from Research Institute of Engineering Mechanics,


Dalian University of Technology, Dalian 116024, People's Republic of China.
718
where Kp=E/% is the real component of the relative permittivity of the
particles. 2~he point-dipole approximation however gives an order of
magnitude lower estimate of the interaction force than actually
observed [5,6]. Solution of the Laplace equation [7,8] gave a force more
nearly that measured. Klingenberg and Zukoski [7] calculated the shear
force of a single chain to be

fs - 12So~a2I~d2Eo2(aIR)4[(2fH+fr) c~ " f~sin30 ] (3)

where f//is the force component induced by the field component parallel
to the axial line of the chain, fr the component induced by the torque
attempting to align the line-of-centers with the field leading to the
formation of the chain, f~ the repulsive component induced by the field
component perpendicular to the axial line of the chain and 0 the angle
between the applied field and the axial line of the chain. Chen, Sprecher
and Conrad [8] gave for the axial attractive force between two adjacent
particles (R/a~2.05)

fa - 37"44Eo~a2t~d2Eo2exp{[ 14.84-6.16(Wa)]~2}. (4)

Finite element calculations [9,10] gave the same magnitude for the
interaction force, with the following empirical equation [10]"

61Ua - 4
fa = ~oa2K~d2Eo2 R/a -~d ~l-~d)/2(~)4 (5)

All of the above studies considered only the dielectric polarization in


the ER suspension. They represent the so-called dielectric polarization
model for ER response induced by the particle/liquid dielectric mismatch
and predict that the electric field dependence of the attractive force and
the resulting shear yield stress is quadratic. However, both the particles
and the host liquids have in reality a fimte conductivity and during recent
years it has become increasingly evident that the conductivity of the
particles and host oil play an important role in the electrorheological
response of suspensions, especially with dc or low frequency ac electric
fields. Anderson [11], Davis [12], Chen and Conrad [13] and Conrad and
Chen [14] suggested that the conductivities of the particles and host
liquids could be included in the framework of the dielectric polarization
model by using the complex polarizability
719

(6)

where the relative complex permittivity K*=K-iK". K (usually designated


K', but here for simplicity we omit the superscript ') is the real
component, K"=o/c0% the loss component, o the conductivity which is
assumed to be independent of the electric field, co the angular frequency
of the applied ac field and i---~-l. The subscripts p and f refer to the
particles and host liquid respectively. Within the complex polarization
framework, Kim and Klingenberg [15] and Webber [16] predicted that
the interaction force between particles should be proportional to
1~*12Eo2f(fl*), where f ( f ) is a complex function of 13".We will term this
model for the ER effect the complex polarization model. When the
electric field frequency co is very high, Equation (6) reduces to Equation
(2), i.e., ~*=~d, and the dielectric effect dominates the ER response. In
this case the complex polarization model reduces to the dielectric
polarization model. When the field frequency co is very low (co-*0) The
complex polarizability ~* given by Equation (6) reduces to

)/(%+2of), (7)

where ~c is called conductivity polarizability or the conductivity-


mismatch parameter. When (~p>>Ot, Equation (7) gives ~c=l, in which
case the particle/liquid conductivity mismatch dominates the ER
response.
The conductivities of the particles and the host liquid are assumed to
be ohmic in the complex polarization models presented in [ 11-17]. These
models still predict that the attractive force between particles and the
resulting shear yield stress are proportional to the square of the applied
field. Although the models include the conductivity in the ER response,
they can only give an estimate of the shear yield stress for two limiting
cases: (a) dc or low frequency ac field and (b) very high frequency ac. For
example, Davis [17] provided a simplified analysis for a suspension of
oxidized metal-particles in a weakly conducting oil. His prediction of the
yield stress with ac electric field is in good agreement with the
experimental results by Inoue [18], but one order of magnitude lower
than those measured with dc electric field.
Although one frequently obtains reasonable agreement between the
measured rheology of ER suspensions and that predicted by the complex
polarization model using the complex permittivities [11-15,17,19-20],
720

the model does not account for the fact that the electric field dependence
of the flow stress is frequently less than quadratic, as observed in some
experiments at dc field or low frequency ac field [ 14,18,21-24]. Also the
model does not explicitly explain that the flow stress generally correlates
with the associated current density [14,25]. These discrepancies
pertaining to the polarization models were first addressed by Foulc et al
[26,27] for the case when the conductivity of the particles is much
greater than that of the host oil. A new concept in their study was that
the host liquid exhibits non-ohmic (non-linear) behavior. They proposed
that the effect of the field on the conductivity of such oils was given by a
simplified expression of Onsager's theory [28]

of(E)- j/E = of(0) [ (l-A) + Aexp~E/Ec], (8)

where of(0) is the conductivity at low electric fields and A and Ec are
constants which depend on the oil. In their model the conductivity
mismatch F=Op/Of(0) between the particles and the host oil is the
important electrical parameter. Further, to make the necessary
calculations they defined two regions in the contact zone between
spherical particles: (a) for a distance from the center line x>6 the sphere
surface is equipotential and the electric current leaving the sphere is
negligible and (b) for x<6 the electric field in the host oil is enhanced due
to the increased conductivity of the thin oil film in this region and
therefore most of the current passes through this region. The enhanced
conductivity of the thin oil film results from the non-ohmic conductivity
of the common non-polar oils used in ER suspensions. Their conduction
model then predicts that the attractive force between the particles is
proportional to Eon, where the exponent n ~ 1 at very high applied field
and n ~ 2 at very low applied field. Qualitative experimental agreement
with their conduction model was obtained for polyamide half-spheres (14
mm dia.) in mineral oil [27] and for a suspension of cellulose particles in
mineral oil [22].
Tang, Wu and Conrad [29], and Wu and Conrad [30] derived the
attractive force and current between conducting particles suspended in
non-ohmic conduction oil using an approach (to be discussed below)
which differs from that of Felici, Foulc and Atten [27]. They obtained
good agreement between the predicted values of both the shear yield
stress and current density and those measured experimentally.
Furthermore in their conduction models, the force and current is derived
as a function of the separation of the particles. In more recent work,
Atten et al [31] have however extended their conductivity model to
include separation of the particles. They proposed that the interaction
721

force decreases approximately linearly with small separation of the


particles. However, it will be shown below that the models of Tang, Wu
and Conrad [29] and Wu and Conrad [30] give a better agreement
between the predicted current density and the rheology of ER
suspensions and those measured.
In their study of non-ohmic conduction, Davis and Ginder [32] give a
simplified analysis for the shear modulus and shear yield stress based on
the work of Felici, Foulc and Atten [27]. They predicted that the shear
yield stress of an ER suspension with dc field is the 1.5th power of the
applied field. Subsequently, Davis [33] extended the non-ohmic
conduction model using an integral equation method, which gave good
qualitative agreement with the field dependence of the measured
attractive force between two particles [27].
Recently, Wu and Conrad [34] extended their work on the non-ohmic
conduction model with dc field to a suspension of metal particles coated
with a weakly conducting oxide film in a non-ohmic oil. Good agreement
occurred between the predicted values of both the shear yield stress and
current density and those measured. Subsequently, Wu and Conrad [35-
37] determined the dielectric and conduction effects on ER response at
any frequency of the applied field. Their predictions of both the
attractive force and the current density were again in good agreement
with those measured by: (a) Boissy, Foulc and Atten [23] for two
polymer particles in mineral oil, (b) Garino, Adolf and Hance [38] and
Miller, Randall and Bhalla [39] for ceramics particles in silicone oil or
dodecane, and (c) Inoue [18] for oxidized metal particles in silicone oil,
respectively.
In this chapter, section 1 gives a general review of the studies on the
role of electrical properties in electrorheological (ER) response. Section 2
first introduces the general physical equations for the combined
conduction and dielectric effects, and then gives the simplified conduction
models proposed to-date for: (a) bulk conducting particles, (b) low
conducting particles with a conducting film and (c) highly conducting
particles with a low conductivity film. This is followed by some
predictions by the conduction models under dc field and by comparisons
between the measured ER response and that predicted. Section 3
introduces our recent work on the conduction and dielectric effects in ER
response under ac field, and reviews the progress in studies of transient
response in ER fluids. Finally, section 4 proposes some design guidelines
for improved ER fluids.
722

2. CONDUCTION MODEL FOR ER RESPONSE UNDER DC


FIELD

2.1 General differential equation for the potential


Intheoretical considerations of electrorheology we first have to define
the local field (or the potential) distribution to determine the interaction
and current density between particles. If particles suspended in a liquid
have neither free bulk charge nor free charge at the interface, the
general differential equation for the potential in both the particles and
the host liquid under ac field is [40,41]:

V-(e*VV)=0 (9a)
or

V.(o*VV)=0, (9b)

where ~*=E+o/io)is the complex permittivity, o*=i~o~*=o+ic0~ the


complex conductivity, ~0 the field angular frequency, o and ~ the
conductivity and real component of the permittivity, V the unknown
complex potential, which is dependent on the frequency, time and
location. Once V is obtained, the electric field can be calculated by E---
VV. If o=0 and ~ is independent of electric field or the electric potential,
Equations (9a-b) become the Laplace equation

v2V=0. (10)

The dielectric polarization models [7,8,10,11] are based on the Laplace


equation.
One of the boundary conditions of Equation (9a) or (9b) at the
interface of the particle and the host liquid is that the current density
normal to the interface of the particle and the host liquid should be equal.
In a general form
n
(11)

where Ep=OVp/Onand E~=OVf/Onare the electric fields normal to the


interface in the particles and the host liquid respectively, and t is the
time. For an ac applied field Eo(t)=Eoexp(imt) (the potential V can be
assumed to be V=V*exp(icot), where V* is the unknown complex
723

amplitude depending only on the location coordinates), Equation (11) can


be written as

Ep*(OVp/On)=ef*(OVf/On), or (12a)
(12b)

where n denotes the normal direction at the interface of two media, and
again the subscripts p and f refer to the particles and host fluid
respectively. Another boundary condition is that the electric potential
should be equal at the interface of particle and the host liquid, i.e.,

Vp=Vr (13)

If we assume that the particles and the host liquid are non-conducting
or the frequency is very high, Equations (10-13) reduce to those
describing the dielectric polarization model [7,8,10].
As mentioned above, the conductivity of the host oil in ER fluids
generally depends strongly on the electric field. In dc or low ac field, the
non-ohmic character of the host oil has a significant effect on the ER
response. In this case Equation (9), together with its boundary
conditions (11)or (12) and (13), gives a strong non-linear complex
differential equation, which makes the exact solution extremely difficult,
even if a numerical method is used where iteration is involved. However,
from the view-point of engineering application it is generally
unnecessary to get the exact solution requiring extensive computations.
Of importance is that we obtain a reasonable estimate for the ER
response with an acceptable accuracy. Once we have the local field
distribution between the particles, we can easily calculate the attractive
force between the particles, the shear yield stress and the current
density flowing through a chain of particles or the ER suspension.

2.2 Conduction model for nearly-touching spheres


In the following we will discuss only dc electric field except for a
special note. Felici, Foulc and Atten [27] divided the region between two
nearly-touching spheres into two distinct zones: (a) the "contact zone"
and (b) the "non-contact zone". If the radius of the "contact zone" is
denoted by 6, non-ohmic conduction of the host oil is considered in the
"contact zone"(x_~6) and equipotential is assumed in the "outer zone"
(x>6). The distance 6 is obtained by setting the conductance of the solid
sphere equal to the conductance of the host oil. This gives
724

(a/5)ln(a/5)- F/~, (14)

where a is the sphere radius and F=~p/C~f(0) the ratio of the conductivity
of the sphere to that of the oil at low field. They then calculated the total
axial attractive force between the two spheres for low applied electric
field to be
fa = 4ua2~fF2Eo2/[uln(a/5]- (15)

Equation (15) gives a quadratic dependence of fa on F and E o. At high


electric field when x_~5the enhancement of the conductivity of the host
oil becomes significant because of the non-linear electric field dependence
of the oil conductivity; see Equation (8). The axial attractive force in
this case becomes

fa-- 2ua2efEcEo{ln[(10F/u)(~]2 Eo/Ec ]}2. (16)

Equation (16) gives an approximately linear dependence of the force on


the applied field and only a weak dependence on the conductivity ratio F.
In a more simplified estimate they give the attractive force

fa = 2ua2~f'EcEm, (17)

where E m is the saturation field in the host oil due to the strong non-
ohmic behavior of the oil. They estimate that E m ranges from
30kV/mm to 40kV/mm for the mineral oil they used. Thus, they
concluded that for two nearly-touching spheres, under low dc applied
field, the attractive force is proportional to Eo2, while under very high dc
field the force is proportional to the applied field Eo. Their analysis gave
qualitative agreement with force measurements on two nearly-touching,
large scale, semiconducting half-spheres (a=7mm). It should be pointed
out that although their model and experiment show that when
Eo>lkV/mm the dependence of the attractive force on the electric field is
linear, the shear yield stress does not have a linear dependence on the
field. This is because when a chain is sheared the particles in the chain
will separate and thus are no longer nearly-touching (see section 2.3).
The saturation field E m in the host liquid is an important parameter in
the non-ohmic conduction model. Wu and Conrad [30] give the following
empirical equation as an estimate of the saturation field
725

F o Ec
E ~ / E o - 30(~) "1(~0-0)09 (18)

and Davis and Ginder [32] give for the radius of the saturation region of
two nearly-touching spheres

5/a = ~]2Eo/Em (19)

Felici, Foulc and Atten [27] gave the following expression for the
relationship between the axial attractive force fa and the current I
passing through the spherical particle:

fa = 4~;a2~p2Vo 4/I2, (20)

where Vo is the potential difference between the two adjacent half-


spheres. Qualitative agreement with Equation (20) was obtained for
experiments on polyamide half-spheres (a=7mm) in mineral.
Although the conduction model for two nearly-touching spheres
presented a new concept for ER response, it could not predict the shear
behavior (i.e., the shear yield stress and the shear modulus) of a general
ER suspension, which is of interest in engineering applications. This
question and the approach will be addressed below

2.3 Conduction model for separated spheres


The strength of ER suspensions can be understood in terms of the
static electric interaction between the particles. A single-row chain of
the particles is the most basic and the simplest structure in ER fluids.
Hence, most of the theoretical studies on ER response [6-12,29-37] are
based on an analysis of the mechanical and electrical properties of a
single chain. Two geometric arrangements for the single chain are used:
(a) the chain is parallel to the applied field and (b) it makes an angle with
respect to the applied field. Klingenberg and Zukoski [7] used the
geometric arrangement (b) and gave an exact analysis for the restoring
force of the two-sphere system. Most studies, especially in the non-
ohmic conduction models, consider the arrangement (a) to obtain first
the distribution of the local field and then to estimate the shear modulus
and the shear yield stress when the chain is sheared. Method (a) is
simpler than (b) and has an acceptable accuracy [42]. All of the work on
the conduction model covered in this section uses this method.
726

Assume that with application of a dc field E 0 the chains formed in an


ER suspension consist of spherical particles with radius a and are
distributed uniformly between the two electrodes. Due to the symmetry
we need consider only two half-sphere neighboring particles as shown in
Figure 1. The following two approximate equations then apply
[29,30,34]:

OpEp(x)= of(E)E(x) (21a)

[R-h(x)]Ep(x) + h(x)E(x) - RE o (21b)

where R=2a+s is the distance between the two half-spheres' centers,


h(x) the gap between the two half-spheres at any location x, Ep(x) and
E(x) the local field in the particles and in the host oil, respectively.
Equation (21a) is the continuity condition of current density and
Equation (21b) that of the potential. Wu and Conrad [36] have shown
that Equations (21a-b) give a good estimate of the local field distribution.

- -~'I h(x)
of Ef _Io I eo

(a) (b)
Figure 1. Schematic of the conductivity model for two separated
spheres [29,30]: (a) The geometry and symbols employed; (b) an area
element in the horizontal plane, of is the conductivity of the host liquid,
Op the conductivity of the particles, Ef the dielectric permittivity of the
host liquid, epthe dielectric permittivity of the particles, a the radius of
the particles, s the separation of the particles, and h(x) the gap between
the two particles at location x.
727

It was found that most of the current passes through the contact zone,
which dominates the main behavior of an ER fluid. Good agreement
occurred between the predicted and experimental data [21] for the effect
of applied electric field on the quasi-static yield stress XE of a
zeolite/silicone oil suspension.
Based on the work of Felici, Foulc and Atten [27], Davis and Ginder
[32] predicted the shear modulus of an ER suspension with non-ohmic
conductivity at dc field to be

G= 3EoKf0EoE m (22)

and the shear yield stress

4 1
XE = ~EoKfOE ~ . 5 ~ (23)

where ~ is the volume fraction of the particles in the suspension. Wu and


Conrad [30] derived the saturation field Em in the "contact zone" to be
that given in Equation (18) (note: they used the conductivity definition
o=j/E, which will be used in the remainder of this chapter except where
noted otherwise). Outside the "contact zone" they used the equipotential
assumption. Figure 2 shows the maximum local field in the liquid layer
Ef=E(x=0) versus the separation of the particles. It is clearly seen how
104
~u~otential: E/Eo:I+I/S :
103 ~- ~ A=0.007
=. Ec:O.ZikV/mm
o
r.l.l 10 2 ; Eo(kV/mm)= ~ r=lo 7

: ;4 -' " '


101 . . . . . . . . .

10o ' ~ , ,A,,,,I , ,, ,,,,,I , ,, ,,,,,I , ,, ,,,,

10-4 10-3 10-2 10-1 10 0


S=s/2a

Figure 2. The ratio of the maximum local field in the fluid layer to the
applied field v s the normalized separation for different applied fields [30].
728

the saturation field varies with the normalized separation S and the
applied field E o.
Figure 3 shows the normalized attractive force F (f~=na2~oK~o2F)
between the two particles. If Em/Eo<I+i/S, no saturation of the field
occurs and the normalized attractive force
F = 0.955/S. (24)
If Em/Eo>I+I/S , a saturation field occurs in the liquid layer and the
normalized attractive force is

F m -- 66(F/A)~ l(Ec/Eo )n (25)

in the range ofEo=l-10kV/mm , Ec=0.001-1kV/mm , A=0.0005-0.5 and


F=103-109. We have n=0.92178Ec ~176 when Ec=0.01-1kV/mm and
n = l when Ec=0.1-0.3kV/mm. Therefore, the general form of the
normalized force can be written as
F = min.{Fm, 0.955/S}. (26)

The shear stress of an ER suspension is given by

3 (~oK~o2F(?) ~ (27)

lO~ ~ " ' ' '''"'1 ' ' '''"'! ' ' .... "1 ' ' '"'~"_~

F=]O 7
A=0.007
]o3 ~- ~ Ec=0.21kV/mm
:: ~ j F=fa/[:n;eoKfEo 2a21
-_ 1 . "~ =0.955/S
102

101 - lO .....

10 0 I/ , , , ~,~,,I , n ...... I , , , , ,,,,! . . . . .

10-4 10 .3 10 -2 10-1 ]00


S=s/2a

Figure 3. Normalized attractive force between two particles US


normalized separation for different applied fields [30].
729

The shear yield stress can then be obtained by maximizing Equation


(27).
Wu and Conrad [30] give the following estimate for the relationship
between the saturation field and the saturation radius for any
separation of the spheres:

Em/E o - (l+S)/[S+1_41_(5/a) 2 ]. (28)

When 5/a<<1, we have

(29a)
or

(5 = ~ a~(I+S)Eo/E m -S (29b)

For nearly-touching spheres (S--0)

(5 = a%]2Eo/~, (30)

which is same as that given by Davis and Ginder [32] for two nearly-
touching spheres; see Equation (19).
The average current density of a single chain is
a

If a saturation field occurs in the liquid layer


a

For two nearly-touching spheres


a

J : ~of(~)Eo + ~ rxE(x)of(E)dx. (33)


0,.6

If extreme saturation occurs in the liquid layer, the current density is


approximated by

J = 2(~f(Em)Eo (34)
730

with an error of less than ~10% compared with Equation (32). Equation
(34) gives a reasonable estimate for the current density of a common
ER fluid chain. For an ER suspension with particle volume fraction ~,
Equations (31-33) should be multiplied by 3~/2 for an ideal arrangement
of the particles.
Recently, Atten et al [31] presented an extension of their earlier
conductivity model [27] to the case of small separation between two
half-spheres (a=7mm) in the saturation regime of the liquid phase.
However, their prediction does not show good agreement with
experimental results; see Figure 4, which includes the prediction by the
conduction model of Wu and Conrad [30]. The conduction model by the
latter authors shows better agreement with the experiments than Atten
et al's model. The non-ohmic conductivity parameters of the oil employed
in Wu and Conrad's calculation were obtained by fitting the experimental
data of Atten, Foulc and Banqassmi [43] to Equation (8), which gives
of(0)=3xl013S/m, A=1.35 and E~=l.49kV/mm. The dielectric constant of
the oil is Kf=2.2 and the conductivity of the particles is o p= 1.7xl0SS/m.
Figure 5 further compares the dependence of the measured attractive
force on the applied field of two half-spheres [27] and that predicted by
the non-ohmic conduction model [30] and the polarization model [8].
Again the non-ohmic conduction model shows better agreement with
experiment than the polarization model. At low dc field and large
separation the attractive force is proportional to the square of the
applied field, and at high dc field or small separation it is approximately
proportional to the applied field. This behavior is in good accord with that
predicted by the non-ohmic conduction model. Davis [33] uses an
integral equation method to calculate the attractive force between two
conducting spheres suspended in a non-ohmic oil and gives a prediction
similar to that of Wu and Conrad [30]. However, his polarization model
gave a quadratic dependence of the attractive force on the applied field.
It should be pointed out that the non-ohmic conduction models
[29.30,32,33] predict that the shear yield stress of ER
suspension is proportional to approximately Eo 15 at dc applied field
ranging from 0.5kV/mm to 5kV/mm, which differs from the prediction for
the axial attractive force between particles. This is because the yield
stress in shear occurs when the particles are separated by the shear,
with the result that the saturation region of the local field in the liquid
between the particles decreases compared with the case of nearly-
touching particles.
As mentioned above, one of the most important differences between
the non-ohmic conduction model and the complex polarization model is
that the former predicts that the shear yield stress dependence on the
731

1.2
Atten et al [31]
Wu & Conrad [30]
0.8 Experiment [31]"
Z 9 1.357kV/mm
0.6
it 1.071kV/mm
9 0.714kV/mm
0.4 @ 0.357kV/mm

0.2

0 100 200 300 400 500


s (/~m)

Figure 4. Comparison of the measured attractive force [31] vs the


particle separation for two polymer half-spheres (a=7mm) in mineral oil
Elf-T50 with that predicted by the non-ohmic conduction models" ~
Wu and Conrad [30]; ...... Atten et al [31].

101 ' I I I I I I rl I ~ I I I I ~

! Wu & Conrad [30] ~~162


100 - ..'"
:_ -.
Z 10-1

lOCi --- :-s:O 1


10-3 . . . . . "ill ; ~ I I I I , i i/

0.1 1 10
E o (kV/mm)

Figure 5. Comparison of the measured attractive force for two nearly


touching polymer spheres of radius a=7mm (s=0) and separated spheres
(s=100ttm) in mineral oil Elf-T50 [31] with that predicted by the non-
ohmic conduction model of Wu and Conrad [30] and the polarization
model (t$*=t$c=1) of Chen et al [8]: s=0; ...... s=100ltm.
732

applied field is ~Eo 15 while the latter predicts the dependence to be


E02. Davis and Ginder [32] provided a qualitative comparison for several
model ER suspensions. Most of the suspensions show t h a t the shear
yield stress dependence on the dc applied field is ~Eo 15.
Of great interest is the effect of temperature on the ER response of
water-activated ER suspensions shown in Figure 6. The non-ohmic
conduction model [30] gives a reasonable prediction of the ER response
in the temperature range 20--160~ [14]. It is proposed [30] t h a t the
peak in the ER response at about 100-110~ is due to the combined
effect of the temperature p e r se plus its influence on the water content
in the host oil. The conductivity of the oil decreases with decrease in
water content, but increases with the temperature. As a result, the
host oil exhibits a strong non-ohmic behavior at low temperature
(high water content) and high temperature (>ll0~ In the contrast,
the particles may loose some of their water content at high
temperatures (i.e., >100~ giving a decrease in their conductivity
opposite to the temperature effect on the oil conductivity. The combined
effects of temperature and water content in both the particles and the
host oil give the highest conductivity mismatch for the ER suspension
at about 100-110~
It should be pointed out, the above non-ohmic conduction models
predict that the interaction force between the particles increases
continuously with the conductivity ratio of F-Op/Of(0). However,

800. . . . . I .... I .... t ....


Zeolite/Silicone 0il Open" Exp.
~=0.28 Filled: Pred.
600 " W=6wt.% 2kV

400-- //"
" / ./A A "'A A ~,
: / ~ A \
200 " - 9 / ~ .--'~ ~ "~
- ~1~ ~ ) Eo=lkV/mm u"O..
0-, l, , I , , , , I , , , , I I I , ,

0 50 100 150 200


T (~
Figure 6. Comparison of the measured shear yield stress dependence on
temperature for a zeolite/silicone oil suspension and that predicted [30].
733

experiments [44] show that the interaction force first increases, reaches
a maximum and then decreases with F. Boissy, Atten and Foulc [45]
and Wu and Conrad [46] found that when F is small the ER response is
negligible and when F<<I a "negative" ER response occurs. However,
when F is very large a complete chain spanning the two electrodes does
not form and a weak ER response occurs [45]. They proposed that
when F is too large, the two contacting particles will exchange their
charges in a very short time and separate before a complete chain
forms.

2.4 C o n d u c t i o n m o d e l for particles wit h a surface film


In the design of an electrorheological suspension, two important
methods to improve the ER response with respect to the particles are:
(a) add a weakly conducting film (often called insulating film) on highly
conducting particles (e.g., metal particles) and (b) add a reasonably high
conducting film on the surfaces of dielectric particles with low
conductivity. Regarding (a), uncoated metal particles are too conducting
and thus give an unacceptable level of current density for a suspension
of metal particles in a weakly conducting oil [18,20]. Also, due to the
early exchange of the charge on highly conducting particles, suspensions
with such particles exhibit poor ER response [24]. Hence, a low
conductivity surface film must be applied to particles with a high
conductivity. An oxide film is the common method for metal particles
[18,20], and a similar approach was used for highly conducting I2-Doped
Poly particles [47]. Regarding method (b), one procedure is to let the
particles absorb a certain thickness of water on their surface, e.g.,
humidified glass beads [5,48]. The conducting path for these two kinds of
coated particles is however totally different.

2.4.1 Low conductivity particles with a conducting film


For low conductivity particles with a conducting film, the current
passing through the particle mainly flows along the surface film. Felici,
Foulc and Atten [27] have given an analysis for two nearly-touching
dielectric particles with a conducting surface film. Tang, Wu and Conrad
[49] developed a conduction model for dielectric particles with a
conducting film. A schematic of their arrangement and conditions are
shown in Figure 7. Taking glass spheres with a water surface film as an
example, they give that the dependence of the axial attractive force
versus the applied field is quadratic at low applied field or with large
separation between the particles, but becomes linear at high applied
field and small separation. This result is similar to the prediction for
734

I h(x)
E0

/xL r

Figure 7. Geometry of two low conducting particles coated with a


conducting surface film. o = conductivity, E - dielectric permittivity, the
subscripts p, f and I denote the particle, host fluid and the surface film,
respectively, a - radius of the particles, s - the separation the particles,
R = the distance between the two sphere's centers, and h(x) - the gap
between the two particles at location x.

bulk conducting particles suspended in a weakly conducting oil presented


in section 2.3. They further found that increasing E c gives a strong ER
response, which is the same as indicated in Equation (25) for bulk
conduction in the particles. Although the current density passing
through the chain increases rapidly with increase in water film
thickness, the attractive force does not increase as much. This
indicates that a thick water film is unnecessary.
The role of a conducting surface film on a dielectric particle is mainly
to change the conductivity of the particle. Both experimental
observation [50,51] and theoretical considerations [49] exclude the so-
called water-bridge assumption [52], which proposes that the interfacial
tension between the water and host fluid provides a source of the shear
resistance in a water-active ER suspension. For example, in
experiments when Pda=2.1 (the radius of the particles a=100-110ttm),
the separation of the particles s=10ttm. This separation is 3 orders of
magnitude larger than the water film thickness (~220/k) and thus no
"water-bridge" could form between the particles. Nevertheless, the
measured attractive force between the particles is still considerably
large (see Refs.[49,51]).
735

In Figure 8 the experimental measurement of the current density for


a single-row chain of humidified glass spheres (200-220/zm dia.) in
silicone oil [50,53] is compared with that predicted by the conductivity
models [49,30]. Rather good agreement occurs between the measured
values and those predicted by Tang, Wu and Conrad [49] for dielectric
particles with a conducting film (absorbed water on their surface). In Wu
and Conrad's conduction model [30] a composite conductivity of the
particles Op=8XlOsS/m was used, which also gives a good agreement
with the measured result. Figure 9 compares the measured current
density vs shear strain for a single chain of glass spheres in silicone oil
with those predicted. Again good agreement occurs.
A comparison of the measured shear yield stress of the single chain
with that predicted by the non-ohmic conduction models [30,32,49] and
the polarization models [7,54] is given in Figure 10. The non-ohmic
conduction models predict the shear yield stress to be the ~l.5th power
of the applied field, whereas the polarization models give the field
dependence to be quadratic. In the polarization models, we assumed that
Op>>Of, giving the conductivity polarizability at dc field f~= 1. Figure 10
shows that the non-ohmic conduction models predict the shear yield
stress better than the polarization models.

100 ,. i i[lll i ! . . . . . . . [ i
1
_

i i ! i : , , 1

-
Experiments.:
, . 1 _T T i i.t , t |

o 3 Spheres ~,~[
~ 4 Spheres ...............................
~ ~ ~
~_,1~ 0 A 5 Spheres ~ ~ ! -

! i ii
I '!-...........J..........i-~-i-i~--i---
i-'-'~-i -!....!---i......Theoreticati
_ T h e o r e t i ..........
.c...a tii-.......
i i---
i
I i ~. ~ ..... Tangetal[49]
i -----Wu & Conrad 13(
i i :

01 ~ ~ ' ' '


0.20.3 0.5 2 3 45
E o (kV/mm)

Figure 8. Comparison of the measured current density of a static single


chain under de field [50,53] with that predicted [30].
736

51_ . . . . I . . . . . . . . I . . . . I .... -I
~.~_. _Expefimen.tal_
: :._ _~
4 1: ~ ................i............J o 3 Spheres ]-]
~" ~ 0 \ I A 4Spheres/:t
~~i 0 5Spheres1
,~ 2 i iA i ......; ................................
*'~ 1 - ..... : ~.,...................... !...................... i ....... o .........

o ?ii i,, i, :i i i i i
0 0.1 0.2 0.3 0.4
,,1 0.5
u
Figure 9. Comparison of the measured current density of a sheared
single chain of humidified glass beads v s shear strain with that
predicted by non-ohmic conduction models under dc field" 9 Tang, Wu
and Conrad [49]; Wu and Conrad [30].

I L : 1 1 !i J
9 Experiment [51,55] J /, ~ ~
......... Wu & Conrad [30] J ~Z'.,~";"
E] " Tang etai [49] [IIIIZ]III]I]I::]I::I~]:I]L~~I!!!-_
[i:J - Davis & Ginger [32] J:::::::::::::::::::::..~ ~:::::::::-
[---J -- ---Klingenberget al [7] [........i - ~ .....!.............Z
k if!___.~..- .- --Conrad et al [54] ..........!2~iiiiiiiiiiiiiii~iiiiiii--

.......i:i_..:i:!i:i::::ii:ii~ii!:i::::::i~iiii:iiliZ.ii~:i:2:1111!iiiiii:i~ii:ii:ii
_:_:.i_i

0.1 ~ ,J'2" i
0.6 0.8 1 2 3 4
E o ( k W m m )

Figure 10. Comparison of the measured shear yield stress of a single


chain of humidified glass beads [51,55] with that predicted by the non-
ohmic conduction models: Tang, Wu and Conrad [49], ..... Wu and
Conrad [ 3 0 ] , - - - - D a v i s and Ginder [32]; and the polarization models:
Klingenberg and Zukoski [7], Conrad, Chen and Sprecher
[54]. In the predictions by the polarization models it is assumed that
*= 13c=1, i.e., Op>>Of
737

2.4.2 Highly conducting particles with a low conductivity film


For highly conducting particles with a weakly conducting surface
film, the current first passes across the surface film and then continues
in the highly conducting bulk (core). Wu and Conrad [34] developed a
model for such particles. They define a combined conduction parameter
A a OI
= 6~f(0) ' where 6 is the thickness of the surface film, oI the

conductivity of the surface film and gf(0) the conductivity of the oil at
low field. Taking silicone oil as the host liquid (with of(0)=2.4x1012S/m,
Ec=0.21kV/mm, A=0.007), they found that increasing A (i.e., increasing
o I or decreasing 5) increases the ER response under dc applied field. If
the surface film is an ideal insulator and thick (i.e., ai=0), there is no ER
effect under dc field. If 5 is too thin, electric breakdown will occur in the
surface film when the applied field exceeds some value, which was
observed in the experiments [18,34]. Therefore, a reasonable thickness
of the surface film is desired to avoid breakdown. The shear yield stress,
shear modulus and the current density of an ER suspension with a
volume fraction ~ of particles suspended in silicone oil were determined to
be given by the following empirical equations:

T E (Pa) = 19r (35)


G ( P a ) - 55r nG (36)
J(pA/cm z) - 6xl 07r 1 22c~f(0)AnJ (37)

for the range 10<A<10 10 , F_>10 12 and lkV/mm<Eo<4kV/mm, where


nG=l.55Eo ~176n~=0.78Eo ~176 (E o is in kV/mm).
Finally, we will compare theoretical predictions with the experimental
results by Inoue [18] on a suspension in silicone oil of dried spherical
aluminum particles with a so-called "non-conducting" thin oxide skin.
The average diameter of the particles was 20#m with various
thicknesses of surface film. With a dc electric field only the ER response
for a surface film thickness 6=0.2ttm was reported. The reported
composite conductivity of such particles was 101~ which gives
A=83.3 for 5=0.2ttm. For the aluminum core we will take Fp-1018. The
silicone oil used was dimethyl silicone oil with 100cst viscosity at 25~
Inoue [18] did not report the non-linear conductivity parameters for the
738

silicone oil. Here we will use the conductivity parameters of Dow


Corning 200 silicone oil (50cst at 25~ determined by Wu et al [50].
Figure 11 compares the predicted shear yield stress with that measured
[18]. Also shown are the analytical results by Davis [17] based on the
complex polarization model. Good agreement occurs between the
measured yield stress and that predicted by the non-ohmic conduction
model of Wu and Conrad [34], whereas that predicted by complex
polarization model of Davis [17] is about an order of magnitude lower. It
should be pointed out that Davis [17] assumed that the effective
conductivity of the host fluid is 108-107S/m, and thus gives ~*=~o=-1/2
(oi<<of). In this case there would be a negative ER effect [24,46] not a
positive ER effect as reported by Davis [17]. This indicates that his
assumption is not correct and thus leads to a low estimate for the shear
yield stress. If we take ~c=1/2, which requires that o1=0.75o f, the
complex polarization model gives the same value for the shear yield
stress as that given by Davis with ~o=-1/2 (see Figure 11), because his
equation at dc field contains 132. If we take the upper-limit case, i.e.,
~o=1, Davis's analysis gives a 4 times larger yield stress than when

103
- 1;EOc EO 1"46 .
l02

11~ lO 1
_. . . . . . . . . . ~
# f

I
f Experiment [18]
10 o Wu & Conrad [34]
.... Davis [17] (~c=0.5)
. . . . . . . Davis [17] (~c=l)
10-1 -- I I I

1 2 3 4 5
E o (kV/mm)
Figure 11. Comparison of the predicted shear yield stress [34] with that
measured [18] for a suspension of oxidized aluminum particles in silicone
oil (~=0.2) under dc electric field. Also shown are the analytical results
by Davis [17] assuming ~*=~c=0.5 and ~*=~c=l (the upper-limit),
respectively. The average particle diameter is 20;um and the surface
film thickness 5=0.2/1m.
739

~c=1/2, which is still smaller than the measured value. Furthermore, the
complex polarization model predicts that the shear yield stress is
proportional to E02 not ~Eo 15, the latter being predicted by the non-
ohmic conduction model and supported by Inoue's experiment. Of
interest is that Davis' analysis [17] provides a good prediction for the
shear yield stress with ac field. The reason for this may be that with ac
field (especially at high frequency) the conductivity effect will be weaker
than the dielectric effect. The dielectric constant of the materials
employed does not usually change considerably with the electric field and
thus Davis' linear analysis seems more reasonable for ac field than for
dc field. A detailed study of the frequency dependence of the ER response
will be given in sections 3.2 and 3.3.
Figure 12 compares the predicted yield stress with that measured
[34] for ER suspensions of silicon particles (l~5~um diameter) with an
oxide surface film in Dow Corning 200 silicone oil. The conductivity of
the silicon is 4.35x104S/m [56], giving Fp=l.83x10 s. The conductivity of
the oxide film was taken to be oi=4.8x10llS/m. The average diameter of
the particles is 2.5~um and the surface film thickness 5=0.2;um. Good
agreement between the predicted and measured values of the yield
stress is indicated in Figure 12. Figure 13 compares the effect of the

100 , , , , , , rr I I' ", , , , , ,,_ _

- O x i d i z e d S i / S i l i c o n e oil
_

r 6=0.2/~m J
_

10

Exp.
Pred.
I I ,, I I I i I i i ] I _I I I I i i

0.1 1 lO
E o (kV/mm)

Figure 12. Comparison of the predicted shear yield stress vs. applied dc
field with that measured for a suspension (r of oxidized silicon
particles in silicone oil. The particles are of average diameter ~2.51~m
and coated with an oxide skin of thickness 5=0.2~um [34].
740

5o i i i i i i i i ]

Eo=lkV/mm o Exp.
40 --- Pred.
- Oxidized Silicon/Silicone Oil-
"~" 30
o ,=0.23 :
20

10

, , , , , , ,,! , I I I I I I I

0.01 0.1 1
(t,m)
Figure 13. Comparison of the predicted effect of the oxide film thickness
6 on the yield stress with t h a t measured for a suspension (r of
oxidized silicon particles in silicone oil with applied field E o - l k V / m m [34].

10-1 _- 1 i ! i i i i I ]

- Oxidized Silicon/Silicone Oil


- ~=0.23, 6=0.05ym
,~, 10-2 - T = 2 3 ~ /x A A
e,i
: A
_ A 9

1 0 -3
- o
10-4
F

10-5 I I ,, I I , , , ! ,

0.1 0.3 1 3

E 0 (kV/mm)
Figure 14. Comparison of the predicted and measured current densities
of a suspension (~=0.23)of oxidized silicon particles in silicone oil. The
averaged particle diameter is ~2.5/tm and the surface oxidized film
thickness 6=0.05/tm. o Experimental data without electrical breakdown
in the suspension; h Experimental data after several breakdowns;
Predicted by the non-ohmic conduction model [34].
741

thickness of the oxide film. Again, good agreement between prediction


and experiment occurs.
Figure 14 compares the measured and predicted current densities for
the suspension of oxidized Si particles (5=0.05ttm) in silicone oil. When
no electrical breakdown occurred in the suspension, the predicted and
measured current densities are in reasonable agreement. But after
several tests in which breakdown occurred, the measured current
density became an order of magnitude higher than predicted. In
contrast, a few breakdowns had almost no effect on the yield stress.
The breakdown field of a suspension of oxidized conducting particles
depends on the dielectric strength and the thickness of the oxide surface
film. The thicker the film, the higher is the breakdown field. After
several breakdowns, many of the particles no longer have a satisfactory
film to provide the desired degree of "insulation". These particles are
then responsible for the observed dramatic increase in the current.

3. C O N D U C T I O N AND DIELECTRIC EFFECTS IN ER


R E S P O N S E U N D E R AC FIELD

In section 2 we mainly introduced the non-ohmic conduction models


which apply to ER response activated by a dc field. We also gave the
framework of the polarization models which to-date have been applied to
two limiting cases: (a) high frequency ac field (dielectric polarization) and
(b) dc or low frequency ac field (conduction polarization). In this section,
we will present investigations into the combined effects of conduction
and dielectric properties on ER response at any frequency.
Electric polarization consists of three components [57]" (a)
orientational, (b) atomic and (c) electronic polarization. Each
corresponds to the motion of different kinds of particles (molecules and
ions, atoms, and electrons, respectively) with different characteristic
times. Thus, as the electric field frequency increases, non-equilibrium
effects will appear successively in the different components of
polarization. The motion of molecules and ions in weakly conducting
materials will lag behind changes in electric field in the frequency range
of 102-106 Hz. However, atoms can follow the field at characteristic
times of the order of 10-14s. At very high frequency the field may change
so fast that even the electrons cannot follow. Since in ER suspensions
the interested frequency range is usually less than 104 Hz and only few
materials exhibit a Debye-type dielectric relaxation in this frequency
range, we will not include the Debye-type effect in the following
calculations. However, if we know the frequency dependence of the
742

dielectric constant or the conductivity of the materials, it can be easily


considered in the calculations.

3.1 Complex polarization model


For the simplest analysis of the combined conduction and dielectric
effects in ER fluids under steady ac field, the complex polarization model
based on the point-dipole approximation gives a reasonable qualitative
prediction for the ER response [15-17,58]. As mentioned above, these
studies predict that the interaction between the two particles is
proportional to the square of the complex polarizability [3*. Tang and
Conrad [58] have calculated the I~*(r versus the frequency for
conducting particles suspended in a host oil with ohmic conduction and
showed that ER particles with high conductivity ratio Fo=Op/Of and low
dielectric ratio FE=Kp/I~ will have a high polarizability at low frequency
fields, which however decreases rapidly beyond a critical frequency. In
contrast, ER particles with low Fo and high F~ have a low 113"(~o)12 at
low frequency fields, but a higher 113"(co)12 at high frequency.

lO 4

10 a

10 2 : 10 --"

lO far
1 01

10 o
lO o lO ~ lO2 lO3 l~

Figure 15. Calculated diagram of active and inactive ER response


regions for several normalized electric field frequencies ~2. For a given
frequency the region to the left of the curve is the inactive ER region.
Region I is the inactive region; Region II the conductivity domain, which
is only active below a certain frequency; Region III the dielectric domain,
which is only active above a certain frequency; Region IV the active ER
effect region independent of the frequency [58].
743

According to the complex polarization analysis [58], the strength of ER


fluids increases with increase in the I~*(co)12. If we empirically take
I ~*(o))12 -0.5 as the boundary between an "active" and "inactive" ER
effect, we obtain a diagram which defines the active and inactive ER
response regions; see Figure 15. The critical frequency at which 113"(00)12
=0.5 is

where fc has a solution only in two cases: (a) when 0<F~8.243 and
Fa_>8.243 or (b) 0<Fa<8.243 and Fe~8.243. The four regions in Figure 40
represent four kinds of ER response. Region I (F~<8.243 and Fa_<8.243)
is an inactive region, where 113"(o))12is always less than 0.5. Region II
(0<FE<8.243 and F~>8.243) is conductivity dominant, where ER fluids
only show a strong effect below a critical frequency fc" Region III
(0<Fo<8.243 and FE>8.243) is dielectric dominant, which becomes active
when the frequency exceeds f~. Region IV (Fa>8.243 and F~>8.243) is
always active independent of the frequency.
The above analysis is based on the assumption that the particles
have the same dielectric and conduction properties in the bulk and on
the surface. The polarization process becomes more complicated with
surfactants around the particles [15] or composite particles, e.g., highly
conducting particles coated with a thin dielectric surface film [17,18,34]
or dielectric particles with a conducting film [49]. Tang and Conrad [58]
also calculated 1~*(o))12 in the frequency domain for several conducting
surface film thickness ratios. In a special case a new type of frequency
response for the dielectric particles with surfactants was found, which
can not be obtained by bulk polarization alone.
Although the complex polarization analysis is supported qualitatively
by experimental observations [15,20], it does not give a reasonable
estimate of the magnitude of the shear yield stress of ER suspensions.
In the following we will introduce a simplified analysis for the ER
response considering both the dielectric and conduction effects under ac
applied field, which gives better agreement with experiment.
744

3.2 Simplified analysis

3.2.1 Analysis method


Consider a single chain with an infinite number of spherical particles
aligned in the direction of the applied electric field and surrounded by a
nonpolar liquid. Two half-spheres in such a chain were shown above in
Figure 1. Assume that the time variation of the applied voltage is

V(t)=V 0cosc0t=Re(V0 ei~t) (39)

where co is the angular frequency of the applied voltage and t time. Let
E* denote the complex amplitude of electric field. Referring to Figure 1,
along the y-axis (x=0), we have the following approximate equations
with ac field [41]

2aE*p + sE*f= V0, (40a)


O p E p * - O'rE*f, (40b)

so that
~ _ (~176
E*f = 2ao,f + SO,p - 2a(of +ic0~f) + s(o +ic0e ) (41a)
p P

E* - ~176 (~176163176
- = (41b)
P 2ao*f + SO*p 2a(of +ic0s + s(o p +ir163

When c0=0 (dc voltage), O*p - o and o*f = of, the fields are induced by
P
the particle/liquid conductivity mismatch and Equations (41a-b)
become

E* ~176 (42a)
f- 2aof + SOp

E* - ~176
- (42b)
p 2aof + SOp
745

which can be also obtained from Equations (21a-b) when x=0. When o) is
very high or o p =of =0, the field is determined by the particle/liquid
dielectric mismatch and Equations (41a-b) become

epVo
(43a)
E*f = 2aef + Sep

E* = efV~ (43b)
P 2aef + Sep

The complex amplitude of the current density is

j*= c~*pE p = o*fE*f = ~176 (44)


2ao*f + SO*p

Let a-2~eo/Of, f]=afKf (here f is the frequency of the applied electric


field) Fe=Ep/~f,Fo=Op/(3f(0), S=s/2a, we then obtain

. (1 + S ) E o (I+S)E o
= (45)
E f : o,f/o, + s ~+ if]
+S
F o + iQF~

where S=s/2a is the normalized separation of particles, Eo=Vo/(2a+s)


the amplitude of the applied electric field, and ~ a non-ohmic parameter
for the host oil. For ohmic conductivity of the host oil

- 1, (46a)

and for non-ohmic conductivity

- 1 - A + Aexp( Exfi~-~c ), (46b)

where E is the rms value of the local field in the oil. Equation (47) is a
complex equation which can be solved by a computer. The electric field in
the host liquid layer is
Ef(x,t) - Re(Ef*e i~~ = ~E(x)cos(o~t+0 E (x)) (47)
746

IEf*l
where E(x)= ~r~ is the rms value of Ef(x,t) and

1 Im(Ef*)
0E = t a n [ ~ 1 (48)

is the phase angle shift from the applied field and is a function of the
location x.
Outside the contact center (x>0), an equation similar to Equations
(42a-b) can be used to estimate the local electric field distribution
between the two adjacent particles. However, a and s in Equations (42-
47) should be replaced by a'=a~/1-(x/a) 2 and s'=s+2a[1-~/i-(x/a)2].
Further, the normalized separation S should be replaced by
S'=(S+I)A/1-(x/a) 2 -1. This simplified analysis gives an estimate of the
local field with good accuracy [36].
Having obtained the distribution El* (x) of the complex amplitude of the
local electric field, we can get the complex amplitude of the current
density
9 $ $ $

J(x) =(:If (x)Ef (x). (49)

The current density at location x at any time t is

j(x,t) - ~/2 j(x)cos(~ot + 0j(x)) (50)

where j(x)= Ij(x)* I / ~ is the rms value of j(x,t) and

-lr Illl(j (X)*).


(51)
5-
is the phase angle of the current density shift from the applied electric
field, which is a function of the location x.
The rms value of the average current density along a chain is
2~
r

j = {~_ [ ~:a12 ~x~j(x)eos(mt+~(x))dx]2dt}l/2 (52)


747

The attractive force between two particles is given by


a

fa(S,t ) = ~soKf
1 f 2 ~ IEl(X)* 12cos2(eot+0E(x))xdx

= e 0 g a 2 I ~ ~ 2 F(S,t), (53)

where the normalized force F(S,t) is


1
F(S,t) = [ 2(E(~)/Erm~)2 COS2(cot+0E(~))~d~. (54)

Here ~=x/a and E ~ is the rms value of the applied electric field. The
rms value of the normalized attractive force is
tl

Fr~s = co ~ F(S,t)2dt (55)

and the mean value of F(S,t) is


2n
O)
(9
Fmean 2~ (56)
J F(S,t)dt
The shear yield stress is given by

Tnm=SfE,.m~ 2 max. (F,.m~(~,)q:+?2) (57a)

or

TEmean--E:Ferms 2 m a x . (Fmean(Y)q;~2) (57b)

In most experiments, the measuring meters give the rms value of the ac
signal; but some meters may give the mean value.
748

In a real ER fluid, many chains span the two electrodes, Equations


(57a-b) should then be multiplied by the factor 3(]), where ~ is the
volume fraction of the particles in the ER suspension. Conrad, Chen and
Sprecher [54] suggested that a structure parameter As should also be
included if the particles are not sphere-like and the structure consists of
clusters of chains. If the chains of spherical particles in an ER
suspension are considered to be ideal, single-row chains and to act
independently, we can take A~=I.
Calculation of the current density of an ER suspension differs slightly
with ac field from that with dc field. With dc field the current passing
directly through the liquid phase between two electrodes is negligible
compared with that passing along the chains. However, at high
frequency ac field this current can not be neglected. If the currents
passing through both the chains and liquid phases are considered, the
current density of the suspension is:
2n
0)

J-(~- [ nOx~j (x)cos(~ot+Oj(x))dx


I ~ a0~3
0
3 ) ~ j (a)cos(mt+Oj (a))]2 dt}l/2
+ (1-~r (58)

or it can be roughly estimated by omitting the phase angle effects of the


current passing through the chains and the pure liquid, giving
2yt

03

j _ ~{ ~ _ 3co [ a21- u x ~ j (x)cos(o)t+Oj(x))dx]2dt}~/2+ (1_2~)j(a) (59)


q

where j(a) is the current density (rms value) of the pure host liquid under
the applied electric field and is given by

j (a) = ErmsqOf2+(2nf~:f)2' 9 (6O)


749

3.2.2 Ohmic conductivity of the host oil


Figure 16 shows how the ER response changes with the normalized
frequency g~ of the applied field for a single chain. If Fa=F, the frequency
has no effect on ER strength. If F > F , the conductivity mismatch
dominates the ER strength at low frequency of applied field, and if
F~<FE, the dielectric mismatch dominates the ER effect at high
frequency of applied field. Figure 17 is the corresponding normalized
average current density (rms value) in the chain. The current density is
independent of the frequency below a critical value (the conductivity
domain), but it increases in a linear fashion with the frequency after a
critical value (the dielectric domain).
Kim and Klingenberg [15] and Webber [16] have pointed out that the
attractive force is proportional to ]~* ]2Eo2f(fl*) , where f(fi~) is a
complex function of 13" and that it is difficult to find a theoretical solution
for a pair of nearly-contacting spheres. In our calculations it is found
that the attractive force is mainly determined by 113"12 and the phase
angle 0~ of ~*. The normalized attractive force (fa/~a2ef Erms2) can be
roughly estimated by the following empirical equations (-37~176

Fmean='~ff2"~Frms 0.48(24 1[3" 12)exp([3'5 )


-
(61a)

lO3 _= I I I I I I I l l l I 10 3
1

r =ro=l~ l

=]02 ~
:
o'10] ~
7 r~=ro=lO . ~
, 10 ]
- r~=}, ro=]'O--.~ M
L s=o.oo5 r~=ro=s
10 ~ 9 ~ i i i i i I i i I 0o
10-4 10-2 10 o ]o2 lO4
Q

Figure 16. The normalized attractive force between two particles versus
the normalized frequency [35].
750

when [~'2- I~* COS0~[2 < 0.85, where 13 is the real component of 13",and

F m e a n ~ F r m s - 24 1[~* 12exp(2~ ,32 ) (61b)

when 13'2 ~ 0.85. Here

1~*12= (KpKf)2 +(~P'~f)2/(~ e~ (61c)


:)

The mean and rms shear yield stresses of a single chain are predicted by
the following empirical equations when ~'2<0.85:

TEmean~'I:Enns- 2.61EfErms21 ~* 12exp(~ '12 ), (62a)

and when [3'2zO.859

"~Emean---~3"~Erms= 1.1~fEr~2exp(exp[3.1([~'2-0.04)2~ (62b)

Boissy et al [23,24] measured the attractive force between two


polyamide half-spheres of radius 7mm immersed in the mineral oil Elf-
TF50. The reported conductivity of the solid spheres Op=l.7xl0SS/m.
]o8 I I I I I I I ! I I I
- - .... F~--10 3, F o - 1 0 ,-
106 -- - F~-5, Fo-106 ...-~/
~ r'~-5, 1"o=10 . . ~ . . ~

s-o.oos

~_..2--:-.~
10 ~ ~ I I I I I i I I ! ~
10-4 10-2 100 102 104 106 108
f~
Figure 17. The average current density of a single chain versus the
normalized frequency [35 ].
751

The conductivity of the oil and, therefore, the conductivity ratio Fo=
Op/Of was varied by adding an ionic surfactant (AOT) to the oil. The
dielectric permittivity ~_ of the half-spheres decreased with the
frequency of the applied efectric field. The dielectric constant of the oil is
Kf=2.2. Figure 18 compares the measured attractive forces [23] with
those predicted by our model [35]. In the experiments, the exact
separation of the two half-spheres was difficult to observe and thus no
data were reported on their separation. By taking the normalized
separation S=0.005 (R/a=2.01), good agreement occurs between the
predicted and measured attractive force of the two nearly-touching
polymer spheres. It should be pointed out that in the high frequency
regime (f>20Hz) the conductivity ratio has no effect on the ER response.
This is indicated by both prediction and the experiments. Furthermore,
the theoretical calculation shows that when the normalized separation
S<0.01 there is almost no effect of the frequency on the attractive force
at high frequency (compare curves a and a' in Figure 18). However, at
low frequency the separation of particles has a significant influence on

10 o
dr
~
j ' Erms=0.202kV/mm -
10-1<
( ....~ S = 0 . 0 0 5 ( R / a = 2 . 0 1 ) -=-
" a "'-,

10-2 l :2!-! .... ,:, ............. i"',


~d ""'", -
10-3

...... Fmean :.
10-4 I _ . . . . . . . . I ........ J ........ I . . . . . . . ,! .......
0.01 O. 1 1 10 100 1000
f (Hz)
Figure 18. Comparison of the predicted attractive force between two
polymer half-spheres immersed in mineral oil (Elf TF50) with that
measured (radius of the particles a=7mm, S=0.005, i.e., tUa=2.01).
Experimental: open symbols; theoretical: solid curve for Frm~, dashed
curve for Fmean. Experimental [23,24]" O Fo=l.4xl04, EIF(j= 1.2x103, A
['0=300, ~ ['0=30. Theoretical [35]: (a) Fo=l.4xl04, (b) Fo=l.2xl03, (c)
I"o=300, (d) Fa=30 , and (a') S=0.001, Fo=l.4xl04.
752

the attractive force, especially at a high conductivity ratio Fo. In Figure


18, we give two theoretical curves: (a) S=0.005 and (a') S=0.001 for Fo
-1.4x104 and the two dc measured points given in Refs[23] and [24],
respectively. The difference between the two measurements for the
same condition is apparently from experimental error, which is also
indicated but smaller in the high frequency regime.
Figure 19 compares the predicted shear yield stresses vs. applied field
for three ER suspensions with those measured (open symbols) [38]
employing 400Hz ac applied field. Good agreement occurs between the
experimental results and the predictions. For the BaTiO3/dodecane
suspension, slightly different results in different figures in [38] were
given. This may be due to measurement error. Figure 20 compares the
predicted and measured shear yield stress dependence on the frequency
of the applied field for the BaTiO3/dodecane suspension. Due to the small
conductivity and high dielectric constant of the BaTiO 3 particles, there is
no effect of frequency in the range from 10 to 4000Hz. This indicates
that for the BaTiO3/dodecane suspension the transition frequency
between the conductivity and dielectric domains is lower than 10Hz.

| [ ' ! i !

lO00 r ,:0.2 2 BaTiO3 -


- f=400Hz ~ z x SrTiO3
lO0 TiO2

10
Predicted [351
o/x r3 Experiment [38] ~-
1 , 1 , : I ,,J I,,,

0.8 1 2 3 4
Eo (kV/mm)

Figure 19. Comparison of the predicted (rms) [35] and measured [38]
shear yield stresses v s the amplitude of the 400Hz ac applied field for
three ceramic particles/dodecane suspensions" TiO2, Kp=70; SrTiO 3,
Kp=270; BaTiO3, Kp=2000; I~=2. In all the calculations Fo=10 and
of=10llS/m were taken.
753

103 ....... I ........ I


i" I I ,,1r

Predicted [35]
Experiment [38]
~" 102 =

O O 0 O 0

101 BaTiO3/dodecane _
,=0.2 :"
Eo- 1kV/mm
100 I I I I I llJ I I I I I I lli I l I I I I II

10 102 lO3 lO4


f (nz)
Figure 20. Comparison of the predicted (rms) [35] and measured [38]
shear yield stresses vs the frequency of applied ac field for
BaTiO3/dodecane suspension (Kp=2000, Kf=2, ~=0.2, F~-10 and of=10
llS/m).

3.2.3 Non-ohmic conductivity of the host oil


In most of the following calculations we will take Dow Corning 200
silicone oil as an example of a non-ohmic host liquid. Experimental
measurements [50] for this oil give the following non-ohmic conductivity
parameters: of(0)=2.4x10"12S/m, A=0.007, Ec=0.21kV/mm. In the non-
ohmic conduction model we can not simply conclude in which case the
polarization effect or conduction effect dominates the ER behavior.
However, it is found that if F >>FE, the non-ohmic conduction effect will
dominate the ER behavior (please note F is here defined as
ro=Op/Of(O)>>op/Of(E)). If F < F , the opposite occurs. Furthermore, a high
local electric field occurs mainly in the "contact zone" of the two
particles, which is same as predicted above. Another characteristic of
the local field in a liquid with non-ohmic conductivity is that there exists
a saturation regime near the contact zone.
Figures 21a-b give the shear yield stress and current density versus
the frequency for different values of Fo and F . If Fo>>F e, a higher
shear yield stress occurs at low frequency than at high frequency. If
Fo<F ~, the opposite occurs. The rms value of the shear yield stress is
754

about 1.2 times the mean value, which differs slightly from the
attractive force. The reason for this may be that the shear strains at
which yielding occurs varies with frequency. Of interest is that if the
dielectric effect is larger than the non-ohmic conductivity effect the
transition regime from conductivity domain to dielectric domain is larger

2.4

2
~, 1.5
1.6

(a) 1 . .
3
. .~ . 3 . V/~~l 1.2
0.8
~
0.5 - k

0 ~ 0"4

10-7 10-5 10-3 10-1 101 103 10 50


f (Hz)
I I i I i I i I /
I05 1. F~=5, Fo=106; 2. F~=IO3, Fo=106
3. FE=103,Fo=I(P; 4. FE=103
103 -
1,2
(b) ,~ 101 -

10-1 _ S=0.005
- . ~ Erms=3kV/mm
10-3 I I I I I I I I I
10-5 10-3 10-1 101 103 105
f (Hz)

Figure 21. The ER response in the transition regime, conductivity


domain and dielectric domain for a single chain (S=0.005,
Er~s=3kV/mm): (a) the shear yield stress and (b) the average current
density. (1) FE=5, Fo=106; (2) Fe=103, Fo=106; (3) FE=Fo=103; (4) rE-103,
ro=10 [36].
755

for non-ohmic conductivity of the fluid compared with ohmic


conductivity. The reason for this is that if the frequency increases, the
dielectric effect increases, i.e., the local field will increase. However, due
to the non-ohmic behavior the conductivity of the host fluid also
increases. It is the non-ohmic behavior that makes the transition regime
from the conductivity domain to the dielectric domain larger than in the
case of ohmic conductivity when the dielectric effect is larger than the
conductivity effect. Obviously, the stronger the non-ohmic conductivity
of the host fluid, the larger such a effect. However, if the dielectric effect
is smaller than the conductivity effect (Figure 21a, case 1), the
transition regime is similar to that for ohmic conductivity.
10 4 ' i i I

F~=5 .-3.5
.. 5E~:~rms <102Hz 1
,~ 103 Fa=10~ df ~ 103Hz -
=104Hz (a)

-i_--
2
_- /ECrErms
lO 1 , I , I I I I '

0.8 1 2 3 4 5 6
E rm s (kV/mm)
lO 4 I I I ' I I I '__

Fo=106 _"
F~= 103 103 Hz -
50Hz _
dc
~, 10 3 (b)
la

10 2 I I I I I i

0.8 1 2 3 4 5
Erm s (kV/mm)
Figure 22. The shear yield stress (rms value) dependence on the applied
field for a single chain: (a) F~=5, F,j=10 a and (b) FE-10 a, I"o=106 [36].
756

Figure 22a shows the shear yield stress (rms value) dependence on
the applied field when the dielectric effect is smaller than the
conductivity effect (F <<Fa). It is seen that the ER response at low
frequency (f<102Hz) is stronger than that at high frequency and further
shows a field dependence of E r ~ 15. At high frequency (f>103Hz) the
shear yield stress shows a Erms2 dependence. It is interesting that,
although f=103Hz is in the transition regime (see Figure 22a), the shear
yield stress still shows a quadratic field dependence. In contrast to
Figure 22a, Figure 22b shows the case when the dielectric effect is larger
than the conductivity effect. This gives a weaker ER response at low
frequency compared with high frequency. The shear yield stress shows
a field dependence of E ~ 151 , E r ~ 153 and Em~ T M for the frequency
f=0, 50 and 103Hz, respectively.
Figure 23 compares the predicted [36] attractive force (rms value,
S=10 5) with that measured [23] on the two 7mm radius polyamide half-
spheres in mineral oil with electrode spacing of 14mm. The conductivity
parameters employed in the calculation are the same as given in section
op=l.7xl0-8S/m,of(O)=3xlO-13S/m,
2.5.1, i.e., A=1.35 and Ec=l.49kV/mm.
The dielectric constant of the particles decreases with the frequency.
When the frequency f=10, 50 and 1000Hz, the dielectric constant of the
particles is Kp=34, 24, and 18, respectively. It is seen in Figure 23 that
the predicted attractive force is in good agreement with that measured.
The reason the measured force at low field and high frequency is slightly
smaller than predicted is probably that the force was too small to
measure accurately.
Figure 24 compares the dependence of the attractive force (rms
value) on the applied field. With ac fields of frequency f=50Hz and 200Hz
respectively, Wu and Conrad [36] predict that the attractive force is
proportional to E~m~2, which is in good agreement with experiment. With
dc field they predict that the attractive force is proportional to E ~ 13 at
S=10 5, compared with the experimental F n ~ E ~ m ~1.6 . If S=10 2 is
taken, it is predicted that F r a t E r n a l 4 5 (dashed line in Figure 24), which
is in slightly better agreement with experiment. However, with ac field
there is no difference between the predicted and measured field exponent
for this range of S. The reason the field dependence of the attractive
force at de field depends slightly on S is that at high applied field the
saturation of local field occurs over a large range of S. Thus over this
range the attractive force has almost a constant value. At low field,
however, only within small range of S does saturation of the local field
757

10o " ''"'"1 ' ''"'"1 ] ''"'"1 ' ''"'"1 , ,,,,,u_

- - Pred . [36] Exp . [23 ] 9 -:_


10-1< o O. 1kV/mm
l
A 0.2kV/mm -
~ ~ / m m :_
~" 10-2

~-, 10-3

10 -4 I , ,,,,,,I , ,,,,,,,I , , , ,,,...

10-1 100 101 102 103 104


f (Hz)
Figure 23. Comparison of the predicted frequency dependence of the
attractive force (rms value) between two polymer half-spheres
immersed in mineral oil (Elf TF50) with those measured [23]. Radius of
the particles a=7mm, normalized separation S=10 -5. Experimental[23]"
open symbols; theoretical[36]" solid curves.

10o _-- J'" ~ ' ~ I ' ~ '


i i I _-

- dr
10-1

10-2 50Hz

10-3 OHz

10-4

10-5 I I I I I I I I I I z t

0.06 0.1 0.3 0.50.7


Erm s (kV/mm)
Figure 24. Comparison of the predicted field dependence of the
attractive force (rms value) between two polymer particles immersed in
mineral oil (Elf TF50) with those measured (radius of the particles
a=7mm. Experimental[23]: open symbols; theoretical[36]: solid curves
for S=10 -5 and dashed curves for S=0.001. Note: when f=50 and 200Hz,
the solid curves for S=10 5 and the dash curves for S=0.001 come
together).
758

occur. Hence, in this case increasing S leads to a small decrease in the


attractive force; see the dash line in Figure 24. A more detailed
treatment of field saturation is given in Ref.[24].
Figure 25 compares the predicted gain-frequency characteristic of
the polyamide half-sphere/mineral oil system and that measured, where
I is the measured current and Vrm~ the applied field (rms value). At high
frequency the predicted value is in good agreement with experiment,
irrespective of whether the non-ohmic conduction of the host oil is
considered. However, at low frequency, the prediction using the
assumption of ohmic conductivity does not agree well with experiment.
Although the predicted value considering the non-ohmic conduction is
smaller than experiment, the tendency of the gain to increase with
increase in the applied voltage is similar to what occurs experimentally.
Possible reasons for this are: (a) experimental error and (b) the
conductivity given in Ref.[23] of(0)=3xl013S/m of the oil at low field used
in the calculation is a low estimate. Ref.[24] gave of(0)=l.21xl0q2S/m.
If we take of(0)=1012S/m, the predicted gain is in good agreement with
experiment over the entire frequency range.

-120
k

~" - 160 " Pred.[36]:


I-Non-ohmic
~/~ 5.6kV
"- -200 [--|2.8kV _ ~
Exp.[23]: _
O 0.TkV :
1~ -240 ~ A 1.4kV -
~ I" ~:"" 0 2.8kV _

9~ -28o
~,~ 10 -4 10-3 10-2 10-1 10 0 101 10 2 10 3 LO4
f (Hz)
Figure 25. Comparison of the predicted gain of two polymer particles
immersed in mineral oil (Elf TF50 without AOT) versus the frequency
with those measured. Radius of the particles a=7mm, the normalized
separation S=0.10 "5. Experimental[23]: open symbols; theoretical[36]"
solid curves for non-ohmic conductivity and dashed curve for ohmic
conductivity.
759

lO3 -

9 BaTiO3/dodecane, ~=0.2, f=20Hz [59] _


9 BaTiO3/silicone oil, ~=0.1, f=60Hz [39] -

Wu and Conrad [36]: / i


- - - BaTiO3/dodecane ./~L"
lO2 - B aTiO3/silicone o i l ~ . '"

101 _ i ..J i i i ,il _ l 9 i i i J I J_

0.1 1 10
Erm s (kV/mm)
Figure 26. Comparison of the predicted shear yield stress (rms value)
[36] with that measured by Miller, Randall and Bhalla [39] for
BaTiOs/silicone oil suspension (f=60Hz and ~=0.1, Kf=2.8, Kp=2000,
=10l~ of(0)=2.4x1012S/m, A=0.007, Ec=0.21kV/mm) and by
der and Davis [59] for BaTiOJdodecane suspension (f=20Hz and
r K~2.0, Kp=2000, ~p=10l~ of(0)=2.4xl012S/m, A=0.007,
Ec=0.2 lkV/mm).

Figure 26 compares the predicted [36] dependence of the shear yield


stress on the ac applied field for BaTiO3/silicone oil and BaTiOs/dodecane
suspensions and that measured [39,59]. Again, there is good agreement
between the predicted and measured values for both suspensions.

3.3 E R r e s p o n s e of c o n d u c t i n g particles with a dielectric


s u r f a c e film
In section 2.4, we discussed the ER response of particles with a
surface film under dc electric field. In this section we will consider the
ER response of highly conducting particles with a dielectric film under
ac electric field. Inoue [ 18], Yu et al [47] presented experimental evidence
that highly conducting particles with a dielectric surface film suspended
in a weakly conducting oil show significant ER response. Davis [17]
gave a theoretical analysis of this based on the point-dipole
approximation. Subsequently, Davis and Ginder [37] proposed the
following expression for the shear yield stress at high frequency ac field:
760

3 ~/gla
- 2 *Q rm 2Ef (63)

Wu and Conrad [34] developed a non-linear conduction model for


highly conducting particles with a dielectric surface film under dc electric
field; see section 2.4. Their recent work [37] shows t h a t this type of ER
suspension gives an exciting ER response under ac field if the surface
film has a high dielectric constant. For simplicity we will only consider
ohmic conduction of the host oil in the following, because non-ohmic
behavior is generally not very important at high frequency ac field.
Figure 27 shows how the shear yield stress depends on the normalized
frequency ~2=2~fe0t~o f. For the general case, the normalized shear yield
stress ~E--~E/[E0t~rm~ 2] is given on the right side of the figure. To give
some idea of the magnitudes involved, the dimensional shear yield stress
is given on the left side for the specific parameters listed (i.e., coated
metal particles suspended in silicone oil). It is seen that if the

f~ 20 - i , i i l i i _ 100
- I', I._lO 3

15- ' ~ - 80
- FoI= 10 2 _
-
~) 10 - Fo=IO~2 - 60
- Fs=lOl o - I[,~
_ _ 40
F I=IO 3
5 2
"~ - FoI=2~ 5/a=O.O1 - _ 20
~,~ -
_
0 I i I i 1 I i
1 0 - 5 1 0 - 4 1 0 - 3 1 0 - 2 10-1 10 0 101 10 2 10 3

fl=2n'fEf/af
Figure 27. Shear yield stress (rms value) of a single chain of highly
conducting particles coated with low conducting film vs the normalized
frequency. The right ordinate is the normalized shear yield stress YE------TE
/[e0KfErms 2] with Fa=(jp/(Jf=1012, F=~p/~f=lO 1~ FEI=ei/~f=103, 5/a=0.01.
The left ordinate is the dimensional shear yield stress XE for the special
case of coated metal particles suspended in silicone oil with (Jf=2.4xl0-
12S/m, Kf=2.5, Erms=3kV/mm and f=50Hz [37].
761

conductivity ratio F~ of the surface film to the host oil is less than the
permittivity ratio F J , the ER strength is higher at high frequency ac
field than at dc or low frequency. For the special case of coated metal
particles suspended in silicone oil with af=2.4xl012S/m, the transition
frequency fc is about 10 -4-10-2 Hz. This transition frequency is
proportional to the ratio of the conductivity af to the permittivity ef of
the host oil, i.e., fc=ff2c(Jf/(2~tef). If we take of=10SS/m and s163 fc=l -
100Hz, which is same as the estimate by Davis [17]. No effect of
frequency on the shear strength occurs when the conductivity ratio
F~ equals the permittivity ratio F I. However, it should be pointed out
that when the ratio FI is high, the local field in the liquid layer between
two particles increases. This increased local field will increase the
conductivity of the host oil if it exhibits super-ohmic behavior, and the

I~ 20, ' "', ..... i ........ J ........ l -_ I00

- ~=2897.5 5/a=0 01 :
15 F,pl012 " --: 80

- F=IOI 0
60
@) 1 0 - ro~=1.67
- -_ 4O I~
5
~ -_ 20

O --- , ~ , ,,,,,I ' I I tl ~1 I I ~ i~,~l 0

1 10 100 1000
FEI=ei/e f
Figure 28. Shear yield stress (rms) of a single chain of highly conducting
particles coated with low conducting film vs the permittivity ratio
F~I=ei/ef of the surface coating to the oil. The right ordinate is the
normalized shear yield stress YE---'~E][s 2] with ro=(~p/Of=1012,
F~=ep/ef=lO 1~ I4o=0i/0f=1.67, g~=2897.5. The left ordinate is the
dimensional shear yield stress ZE for the special case of coated metal
particles suspended in silicone oil with of=2.4xl012S/m, Kf=2.5,
Erms=3kV/mm and f=50Hz [37]
762

real conductivity ratio may be lower than suggested. In this case the
non-ohmic conduction effect of the host oil should be considered
especially at low frequency.
Figure 28 gives the shear yield stress versus the dielectric constant
ratio of the surface film to the host oil for a single chain of particles. The
yield stress increases with the dielectric constant ratio, which indicates
that conducting particles with a high dielectric constant surface film
can give an exciting ER response. For example, if the volume fraction of
the particles ~=0.4, the surface film thickness 8=0.01a and the surface
dielectric constant ratio FEI=I~/~1000, and if we choose a structure
parameter As=10 [54], we can get a shear yield stress of 100kPaI Even
if we consider the structure parameter A~=I, one still obtains a shear
yield stress of more than 10kPa. In their recent work [60], Wu and
Conrad give the following empirical equation to estimate the shear yield
stress of highly conducting particles coated with a dielectric film
suspended in ohmic oil (As = 1):

8 grE 2
x E (Pa) - O.O06+(~FEI)O.6 , (64)

where A=5/a and the units of the applied ac field Eros are in kV/mm.
Figure 29a compares the experimental shear yield stress [18] for
aluminum particles with an alumina surface film with t h a t predicted
[ 17,37]. The frequency of the applied field is 50Hz. Davis [ 17] assumed
that the surface dielectric constant KI=5 for A1203 , but Randall et al
[61] report I~=10, whereas Thurnauer [62] report Ki=8-9. In Figure 29a
we give two cases, when KI=5 and KI= 10. Both values give predictions in
good agreement with the measurements of the yield stress. Even though
the surface conductivity ratio FoI may range from 1.67 to 100, no
significant difference occurs in the predicted values of both the shear
yield stress and the current density. This indicates that when the
frequency f=50Hz the surface conductivity ratio has no effect on the ER
response, which is seen more clearly in Figure 27. Figure 29b compares
the measured [18] current density with that predicted [37]. Again good
agreement occurs.
Figure 30 compares the dependence of the measured shear yield
stress on the oxide film thickness with that predicted [37,17]. It is clear
that good agreement again occurs between the predicted and
experimental values. The term "EG yield stress" is defined by Inoue [ 18]
763

to mean that when the surface film thickness is too small a field higher
than the breakdown strength of the film material could not be applied.
From the analyses by Davis [17] and Davis and Ginder [32] and the
calculation by Wu and Conrad [37], the smaller the surface film
thickness 6 the higher the shear yield stress. However, a decrease in the
film thickness also raises the local field in the film, which increases the
possibility of dielectric breakdown. In this case a high field can not be
applied, the result being that the ER suspension exhibits a low shear
yield stress. This can be seen in Figure 30. According to the above
analysis, the surface film should have a high dielectric constant, a high
dielectric breakdown strength and a reasonably small thickness.

1000 . . . . ~...... I............................................. ! .......................... ! ........ / ! . ...........


-!!!!!i!!!i~!i!!!!!!!!!~!!!!!!!!!!!~i~!!!!!!!!!!~!!!~!!!!!i!!!!!!!!!!!!!!!!!!!~!!!!!~i~;ii!!!~!~t!!!!!~!~!!!---

.... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . " . . . . . . . i!!!!:::!!!:!:-


...... i ...... i ........................................... ! ....... -,~.-.~'-"- ....... i .................. i .............
...... i...... i................................ -/.---~----':~'.. ................................... i .............

-~ 100 :::::::-,
i!::!!,'
...... , .............. <...:
t..--.............. , ............ :::::::::::::::::::::::::::::::::::::::::::::
i i:! (a)
:::::::::::::::::::::::: Wu & Conrad [371 (K,:5)
...... !...... !........... Wu & Conrad [37] (K,=10) [--
...... i ...... i........... Davis [17] (K,=5) ~-

10 i i i i
I "

i S=O.O01 ,'
10 ...........!.............................................~.........................................................

!i (b)

i -- Pred. [37] (KI=5) [ _J


1 -!!!{{!!{!!i!!!!!!!!!!!!:!!!!!!!!!!!!!!. - Pred.[371 (K,:10) [!!]!!!::

0.8 1 2 3 4 5
Erm s ( k V / m m )
Figure 29. Comparison of the measured ER response of oxidized
aluminum particles suspended in silicone oil [18] with those predicted
[17,37] (5=0.2/zm, a=10/zm, (I)=0.2, f=50Hz, of=2.4x1012S/m, Kf=2.5,
ro=Op/~f=1012, r~-ep/ef=101~ and FIo=oi/of=l.67): (a) the shear yield
stress and (b) the current density.
764

.-. 800 E,,,~l,,,,i .... i , , , i , ~ , i , , , ' t , ' " l " ' ~ ' -
~ i ---Wu & Conrad [37]-
~ .... Davis & Ginder[32]-
600 L- ! o o 9 Experiment [18]
Erms=3kV/mm
400 ... 50Hz -

: ""0 "
200 ~- i'~'-" :". .... / m c

0
0 0.2 0.4 0.6 0.8
5 (Hm)

Figure 30. Comparison of the "EG" shear yield stress measured by


Inoue [18] for oxidized aluminum particles suspended in silicone oil with
those predicted by Wu and Conrad [37] (solid curve) and by Davis and
Ginder [32] (dashed curve): a=10~um, r f=50Hz, ~f=2.4x1012S/m,
Kf=2.5, KI=5 , Fo=Op/Of=1012, r = e p / e f = l O 1~ and lqo=oi/of=l.67.

15 I -I I i

Exp. [18]
- Pred.[37] (K~=5)
10 Pred.[371 ( K , - I _ ~ "

wooa~176176

5
f S=0.001
9 E~b=400kV/mm

0 I I I I I _1 I
0 0.1 0.2 0.3 0.4
a qum)
Figure 31. Comparison of the measured [18] a p p l i e d breakdown field
for oxidized aluminum particles suspended in silicone oil with those
predicted [37]: a=10/tm, ~=0.2, f=50Hz, of=2.4x10lZS/m, K~2.5,
Fo=Op/Of=1012, Fe=13p/Ef=1010 and FIo---oi/of=l.67.
765

The dielectric strength of a surface coating depends on its density,


temperature, thickness and the purity. Moulson and Herbert [63]
reported that the dielectric strength of purified, high density alumina of
lttm thickness is about 500kV/mm at room temperature and decreases
with the specimen thickness and density of the alumina material.
Whitehead [64] reported that the dielectric strength of hollow lead-glass
spheres increases with specimen thickness, but that of clear ruby
muscovite mica decreases with the specimen thickness. Figure 31
compares the breakdown applied field measured by Inoue [18] for
oxidized aluminum particles suspended in silicone oil and that predicted
by Wu and Conrad [37], by taking the breakdown strength of the
alumina surface coating to be EIb=400kV/mm. Good agreement occurs
between the measured breakdown applied field and that predicted. In
our model, when the peak value of the maximum local field in the surface
film reaches 400kV/mm, the corresponding applied field is considered to
be the breakdown applied field for the ER suspension.

3.4 T r a n s i e n t r e s p o n s e in E R fluids
In Section 1 it was pointed out that for dc field the conductivity
mismatch between particles and fluid is the dominant factor in
determining the ER strength. Using pulsed dc excitation, Ginder et al
[65-68] and Tang and Conrad [58] have demonstrated the separate roles
played by the conduction and dielectric properties of the particles and
host liquid in transient ER response. An interpretation of the
experiments, based for simplicity upon a point-dipole analysis, sheds
considerable light on the mechanism of ER activity. Recently, Davis [33]
studied the ohmic (linear) ER response to a pulsed field and the non-
ohmic (non-linear) ER effects at high dc fields using an integral equation
method. Based on the point-dipole approximation, Ginder, Davis and
Ceccio [68] proposed a model to qualitatively predict the transient
response. The most important prediction of Ginder, Davis and Ceccio
[68] is that the shape of transient stress curves depend sensitively on
the relative contribution to the particle polarization mechanism of the
dielectric mismatch ~d=(Kp-Kf)/(Kp-2Kf) and conductivity mismatch
~c=(CJp-~f)/(~p-2(~f). If the ER response is mainly induced by a
conductivity mismatch (~d/~c<<l), then when an electric field is suddenly
applied the shear stress will increase over a relaxation time
t*=~o(Kp+2Kf)/((~p+2af), and finally reaches a maximum in the steady
state. If the ER response is instead induced by a dielectric mismatch
(~d/~>>l), sudden increase in the field produces an instantaneous peak
766

in the stress, which then decays over a time t* because of the residual
conductivity mismatch. These distinct qualitative behaviors enable a
determination of the mechanisms of electrorheological transient
response. The predictions of the ER transient response were fitted to
the measured stress transients using nonlinear-least-squares curve-
fitting algorithm; the solid curves in Figures 32-34 demonstrate that
qualitative fits were obtained. Good agreement also occurred between
the values of ~d/~c and t* derived from independent fits of the unipolar
and bipolar transients in amorphous aluminosilicate and barium
titanate suspensions.
For the amorphous aluminosilicate fluid the fit to the date gives
~d/~c=0.15, which suggests that ~d/~c<<l. As stated above, a

200 (I) ' ' ' ' -' 300 r--' , ,


5 S -1
7 = 5 s

#
O~ 1SO
200

100

20 . . . . 40 ' , ' , ' , "' e /


rphous Aluminosilicot

lO 20

-lO -20

I--~ -20 . . . . ' " ' - 40 ' -

2 I 2 r-- ! "i

5.0 Hz I (TTT~I) 1.0 Hz

O'
I I I 1 --2 1 I l .

0.2 0.3 0.4 0.5 0.0 0.5 I .0 I .5 2.0


0.0 0.1

time (s) time (s)


(a) (b)
Figure 32. The time dependence of (I) the shear stress 1; (dots); (II) the
current density and (III) the applied field E o in the amorphous
aluminosilicate ER fluid. The sample was sheared in a Couette cell at a
constant rate Y=5s"1 and was excited by a lkV/mm peak amplitude.
The solid line in (I)is a fit to the stress by the dynamic ER response
model [68]. (a) 5Hz unipolar square wave; (b) 1Hz bipolar square wave.
767

conductivity mismatch mechanism is associated with the finite time


required for the growth of the shear stress in this system (see Figure
32a). That ~d/~c <<1 is expected in ER systems in which the transport
of real charge carriers --ions, electrons, or holes -- to the interface
between particle and host liquid promotes the ER effect[9,11,12]. Within
the transient model of Ginder, Davis and Ceccio [68], such a situation
arises from a substantial mismatch between the particle and host liquid
conductivities. It may also result from surface conductivity effects due
to absorbed polar impurities. When the field direction is reversed (Figure
32b), a finite time is required for the charge carriers to redistribute and
thus for the ER response to recover. The parameters obtained by fitting

200 , , , 200 r .... , ' I


-1 i
(I) 5 ~-' 15o ~ (I) 5 s i
r ~5o
7
I~ 100 100 ~ ~ . ' i
50 50 q
0 1 i 1
0 1 ,. J

~ 80 , i 80 i i

(II) ' Barium Titanate Barium


~ 4o 40
Titanate

o 0 "

-40 -40 7
8O

2 1 I
10 Hz I) _ _ ~ ~ 10 Hz
2 I) I

~ 0 -I
o I
I , | ~ 2 . , i | i I

0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20

t i m e (s) t i m e (s)
(a) (b)
Figure 33. The time dependence of (I) the shear 1: (dots); (II) the
current density j and (III) the applied field E o in the barium titanate ER
suspension. The sample was sheared in a Couette cell at a constant rate
Y=5s 1 and was excited by a 2kV/mm peak amplitude, 10Hz unipolar
square wave. The solid line in (I) is a fit to the stress by the dynamic ER
response model [68]. (a) 10Hz unipolar square wave; (b) 10Hz bipolar
square wave.
768

the field reversal data again indicate f~d/f~c=0.16<<l, as expected for a


system dominated by conductivity mismatch. However, in the barium
titanate system, the fit to the data gives f~d/f~c=7.2,which indicates that
f~a/[tc>>l. This is due to the high dielectric constant and low conductivity
of the barium titanate particles. Since dielectric-mismatch polarization
occurs almost instantaneously with the application of the field, the large
dielectric mismatch gives rise to the initial particle polarization and to
the shear stress maximum observed when the field is applied (see
Figures 33a-b). Because a finite particle-liquid conductivity also exists,
charge is carried to the particle surface on a time scale t*, finally
screening the dielectric polarization and thus causing the shear stress to
decay to its steady-state value. When the field suddenly returns to zero,
800 r" , , , , '"
L
-t
(I) "y = 5 ,
600

400

200

02 , , ,

~ o.1
~ o.0
-0.1
~-~ -0.2 '

10 Hz

1 . . . . .

o ! .. | I

0.00 0.04 0.08 0.12 0.16 0.20

t i m e (s)
Figure 34. The time dependence of (I) the shear x (dots); (II) the current
density j and (III) the applied field E 0 in the crystalline aluminosilicate,
or zeolite, ER suspension. The sample was sheared in a Couette cell at a
constant rate Y=5s1 and was excited by a lkV/mm peak amplitude,
10Hz unipolar square wave. The solid line in (I) is a fit to the stress by
the dynamic ER response model [68]. The relatively high conductivity of
this fluid is reflected in high current density in (II).
769

the dielectric-mismatch polarization disappears almost instantaneously.


The residual polarization due to the screening charge caused by ~o gives
rise to a smaller transient stress peak (see Figure 33a) that also decays
due to charge transport on the time scale t*.
The unipolar stress transient in the crystalline aluminosilicate
system (Figure 34) indicates that both dielectric and conductivity
mismatches contribute to ER response, since the fit to the ER dynamic
response implies that ~d/~c~0.6. The dielectric mismatch gives rise to
the initial fast component of the stress growth, while the slower growth
toward the steady state stress can be assigned to the conductivity
mismatch. However, if the ~d/~c=l, the ER response will not change
after it reaches the initial value induced by the dielectric mismatch. The
ER response of this suspension will not depend on the field frequency; see
also sections 3.2.2 and 3.3.
It should be pointed out that Ginder, Davis and Ceccio [68] did not
consider the frequency dependence of the dielectric constants and the
conductivities, which may affect the transient response in ER fluids.
Also, they did not include the non-ohmic conductivity of the host oil,
which may give rise to field-dependent transient phenomena. Further,
Block et al [69] proposed that the rotations of the particles in shear flow
may assist charge transport and thus affect the characteristic time t*.
Also, the time-dependent growth of the chains or clusters of particles
induced by the application of the field may influence the transient
response, which was not considered. The observed transient stresses
might be associated with the buildup of stress as the structures are
sheared. The time dependence of the dipole moment was calculated
based on the assumption that the particles were isolated in an external
field which differs from the case of an ER fluid with a high volume
fraction of particles.
Finally, it should be mentioned that good accord was obtained by Tang
and Conrad [58] between their calculated stress-time response using
their simplified analysis based on the work of Ginder, Davis and Ceccio
[68] and the measurements. Davis [33] also gave a good prediction of
transient ER response using the integral equation method.

4. D E S I G N OF I M P R O V E D E R F L U I D S

Recent review articles on ER materials [70-75] give some guidelines


for the design of desired ER fluids. To improve ER response, both the
host oil and particles should be given attention. It is desired that the host
oil has low and stable conductivity, which is not strongly super-ohmic,
770

and has high dielectric permittivity. For the particles, we desire that
they have both a reasonably high conductivity and a high dielectric
permittivity. To improve the electrical properties of the particles two
methods can be used: (a) improve the bulk electrical properties, (b)
improve the surface electrical properties. Our studies suggest the latter
to be an exciting possibility. In Table i some design guidelines are given
for the host oil. In Tables 2, 3 and 4, some recommended design
parameters are given for ER suspensions activated by: (a) bulk
conduction of the particles, (b) dielectric particles with a conducting
surface film and (c) highly conducting particles with an "insulating
surface film", respectively. The recommendations in Tables 1-4 are
based on the theoretical and experimental considerations reviewed
above.

Table 1 Recommended design parameters for


the host oil of ER suspensions

" Conductivity of (a) As low as possible, usually <10l~


(b) Weak non-ohmic behavior.
(c) Low water content.
(d) Not sensitive to the ambient.
9 Dielectric constant High.
9 Density Approximately that of the particles.
9 Viscosity Recommended 20-200Pa-s.

Table 2 Recommended design parameters for


bulk conducting particles in ER suspensions

9 Conductivity Op (a) (~p=106-109S/m, making 107>Op/Of >104.


(b) Not sensitive to the ambient.
9 Dielectric constant As high as possible.
9 Density Match that of the host off.
9 Size Diameter = l---20/zm
9 Shape Spherical, but this needs fin'ther study.
9 Possible Max. x E l-SkPa (Eo=3-4kV/mm)
771

Table 3 Recommended design parameters for "insulating"


particles with a conducting surface film

9 Particles: (a) High dielectric constant


(b) Density close to that of the host oil (~0.97-0.98)
(c) Spherical shape, but needs further study.
9 Film: (a) Conductivity o I =104--107S/m, or oi/of >104
(b) Thickness 5= 10 -3--l~um, or 5/a<0.1
(c) Uniform, smooth and continuous.
9 Possible max. shear yield stress ~E"
DC Field: 1-5kPa (E0=3-4kV/mm)
AC Field: Not clear, but appears to be smaller than with DC field.

Table 4 Recommended design parameters for conducting particles


with "insulating" surface film

9 Particles: (a) High conductivity, Op>104S/m.


(b) High dielectric constant, Kp>1000.
(c) Density approximately that of the host oil.
9 Film: (a) Conductivity, o I =10s-10llS/m.
(b) As high dielectric constant as possible.
(c) High dielectric breakdown strength.
(d) Small thickness: 5/a<0.05, (depending on
its electrical breakdown strength)
9 Possible max. shear yield stress ~E"
DC: ~5-10kPa.
AC: ~10-100kPa.
772

REFERENCES
1. W. M. Winslow, U.S. Patent 25 (1947) No.2417850.
2. W. M. Winslow, J. Appl. Phys., 20 (1949) 1137
3. C.A. Coulson, Electricity (Interscience Publishers, Inc., New York,1961)
4. A. P. Gast and C. F. Zukoski, Advanced Colloid and Interface Sci. 30
(1989) 153
5. A. F. Sprecher, Y. Chen and H. Conrad, Proc. 2nd Int. Conf. ER Fluids,
ed. J. D. Carlson, A. F. Sprecher and H. Conrad (Technomic, Lancaster,
1990), pp. 82-89
6. T.C. Halsey and W. Toor, Phys. Rev. Lett. 65 (1990) 2820
7. D.J. Klingenberg and C. F. Zukoski, Langmuir, 6 (1990) 15
8. Y. Chen, A. F. Sprecher and H. Conrad, J. Appl. Phys. 70 (1991) 6796
9. L.C. Davis, Appl. Phys. Lett., 60 (1992) 319
10. R. Tao, Q. Jiang, H. K. Sim, Physical Review E, 52 (1995) 2727.
11. R.A. Anderson, Proc. 3th Int. Conf. ER Fluids, ed. R. Tao and G. D. Roy
(World Scientific, Singapore,1992), pp.81-92
12. L. C. Davis, J. Appl. Phys., 72 (1992) 1334
13. Y. Chen, and H. Conrad, Developments in Non-Newtonian Flows, ed.
D. Siginer, W. Van Arsdale, M. Aitan and A. Alexandrou (ASME AMD,
New York, 1993) 175, pp.199-208.
14. H. Conrad, and Y. Chen, Progress in Electrorheology, ed. K. O.
Havelka and F. E. Filisko (Plenum Press, New York and London,1995),
pp.55-86.
15. Y. D. Kim and D. J. Klingenberg, Progress in Electrorheology, ed. K. O.
Havelka and F. E. Filisko (Plenum Press, New York, 1995), pp.l15-130
16. R. M. Webber, Progress in Electrorheology, ed. K. O. Havelka and F. E.
Filisko (Plenum Press, New York, 1995), pp.171-184
17. L. C. Davis, J. Appl. Phys. 73 (1993) 680.
18. A. Inoue, Proc. 2nd Int. Conf. ER Fluids, ed. J. D. Carlson, A. F.
Sprecher and H. Conrad (Technomic Publ. Co., Lancaster-Basel, 1990),
pp.176-183
19. H. Conrad, and A. F. Sprecher, J. Statistical Phys. 64 (1991) 1073
20. W. J. Wen and K. Q. Lu, Appl. Phys. Lett., 67 (1995) 2147
21. H. Conrad, Y. Shih and Y. Chen, Development in Electrorheological
Flows and Measurement Uncertainty, ed. D. Siginer, J. Kim, S. Sherif,
and H. Coleman (ASME, New York, 1994), FED-Vol.2051, AMD
Vol.190, pp.69-82
22. J. N. Foulc and P. Atten, Proc. 4th Int. Conf. Electrorheological Fluids,
ed. R. Tao and G.D. Roy (World Scientific, Singapore, 1994) pp.358-371
23. C. Boissy, J. N. Foulc and P. Atten, ed. R. Tao and G. D. Roy (World
Scientific, Singapore, 1994), pp.453-462
24. C. Boissy, P. Atten and J, N. Foulc, J. Intelligent Material System and
Structures, 7 (1996) 599
25. H. Conrad, Y. Chen and A. F. Sprecher, Proc. 2nd Int. Conf.
Electrorheological Fluids, eds. J. B. Carlson, A. F. Sprecher and H.
Conrad (Technomic, Lancaster, 1990), pp.252-264
773

26. J. N. Foulc, N. Felici and P. Atten, C. R. Acad. Sci. Paris. 314 II (1992)
1279.
27. N. Felici, J. N. Foulc and P. Atten, Proc. 4th Int. Conf.
Electrorheological Fluids, ed. R. Tao and G. D. Roy (World Scientific,
Singapore, 1994), pp.139-152
28. L. Onsager, J. Chem. Phys., 2 (1934) 599
29. X. Tang, C. W. Wu and H. Conrad, J. Rheology, 39 (1995) 1059
30. C. W. Wu and H. Conrad, J. Phys. D: Appl. Phys., 29 (1996) 3147.
31. P. Atten, K. Q. Zhu, C. Boissy and J. N. Foulc, J. Intelligent Material
System and Structures, 7 (1996) 573
32. L. C. Davis and J. M. Ginder, Progress in Electrorheology, ed. K. O.
Havelka and F. E. Filisko (Plenum, New York, 1995), pp.107-114.
33. L. C. Davis, J. Appl. Phys., 81 (1997) 1985
34. C. W. Wu and H. Conrad, J. Appl. Phys., 81 (1997) 383.
35. C. W. Wu and H. Conrad, J. Phys. D: Appl. Phys., (1997) in print.
36. C. W. Wu and H. Conrad, "Dielectric and conduction effects in non-ohmic
electrorheological fluids", Submitted to Phys. Rev. E, (1997).
37. C.W. Wu and H. Conrad, J. Appl. Phys., 81 (1997) 8057
38. T. J. Garino, D. Adolf and B. Hance, Proc. 3rd Int. Conf. ER Fluids, ed.
R. Tao (World Scientific, Singapore, 1992), pp.167-174
39. D.V. Miller, C. A. Randall, A. S. Bhalla, et al, Ferroelectrics Letters, 15
(1993) 141
40. M. J. Chrzan, and J. P. Coulter, Proc. 3rd Int. Conf. ER Fluids, ed. R.
Tao (World Scientific, Singapore, 1992), pp.175-191
41. M. Zahn, Electromagnetic Field Theory: A Problem Solving Approach
(John Wiley and Sons Press, New York, 1979)
42. C.W. Wu and H. Conrad, Submitted to J. Materials Research, (1997).
43. P. Atten, J. N. Foulc and H. Banqassmi, Progress in Electrorheology,
ed. K. O. Havelka and F. E. Filisko (Plenum Press, New York and
London, 1995), pp.231-244.
44. H. Block, and J. P. Kelly, Proc. 1st Int. Symp. ER Fluids, ed. H. Conrad,
A. F. Sprecher and J. D. Carlson (North Carolina State University
Publications, Raleigh, 1989), pp.l-26
45. C. Boissy, P. Atten and J. N. Foulc, J. Electrostatics, 35 (1995) 13
46. C. W. Wu and H. Conrad, J. Rheol., 41 (1997) 267
47. W. C. Yu, M. T. Shaw, X. Y. Huang and F. W. Harris, J. Poly. Sci., 32
(1994) 481.
48. Y. H. Shih and H. Conrad, Proc. 4th Int. Conf. Electro-Rheological Fluids,
ed. R. Tao and G. D. Roy (World Scientific, Singapore, 1994), pp.294-313
49. X. Tang, C. W. Wu and H. Conrad, J. Appl. Phys., 78 (1995) 4183
50. C. W. Wu, Y. Chen, X. Tang and H. Conrad, Int. J. Modern Phys. B,
10 (1996) 3315.
51. C. W. Wu, Y. Chen, X. Tang and H. Conrad, Int. J. Modern Phys., B,
10 (1996) 3327
52. Stangroom, J. E., Phys. Tech., 14 (1983) 290.
53. C. W. Wu and H. Conrad, "Electrical properties of electrorheological
particle clusters", To be published (1997)o
774

54. H. Conrad, Y. Chen and A. F. Sprecher, Int. J. Modern Phys. B, 6


(1992) 2575
55. C. W. Wu and H. Conrad, "The shear strength of electrorheological
particle clusters", To be published (1997).
56. G. E. McGuire, Semiconductor Materials and Process Technology
Handbook (Noyes Publications, Park Ridge, NJ, 1988), pp.28-49
57. C . J . F . Bottcher and P. Bordewijk, Theory of Electric Polarization, II
Dielectric in Time-Dependent Field (Elsevier, New York, 1978)
58. X. Tang and H. Conrad, J. Appl. Phys., 80 (1996) 5240
59. J. M. Ginder and L. C. Davis, Proc. 4th Int. Conf. ER Fluids, ed. R. Tao
and G. D. Roy (World Scientific, Singapore, 1994), pp.267-282
60. C. W. Wu and H. Conrad, 6th Int. Conf. on ER and MR Fluids,
Yonezawa, Japan, July, 1997
61. C.A. Randall, D. E. McCauley, C. P. Bowen, T. R. Shrout and G. L.
Messing, Proc. 4th Int. Conf. Electrorheological Fluids, ed. R. Tao and
G. D. Roy (World Scientific, Singapore, 1994), pp.60-66
62. H. Thurnauer, Dielectric Materials and Application, ed. by A. R. Von
Hippel (Technology Press of MIT and John Wiley & Sons, New York,
1954), pp.179-189
63. A. J. Moulson and J. M. Herbert, Electroceramics: Materials, Properties,
Application (Chapman and Hall Press, London, New York, 1990)
64. S. Whitehead, Dielectric Breakdown of Solids (Oxford, Clarendon, 1951)
65. J. M. Ginder and S. L. Ceccio, 65th Annual Meeting of the Society of
Rheology, Boston, MA, Oct. 17-21, 1993.
66. J. M. Ginder and S. L. Ceccio, Polym. Preprints, 35 (1994) 315-320
67. J. M. Ginder and S. L. Ceccio, J. Rheol., 39 (1995) 211
68. J. M. Ginder, L. C. Davis and S. L. Ceccio, Progress in Electrorheology,
ed. K. O. Havelka and F. E. Filisko (Plenum Press, New York, 1995),
pp.281-294
69. H. Block, K. M. W. Goodwin, E. M. Gregson, and S. M. Walker, Nature,
275 (1978) 632
70. K. D. Weiss and J. D. Carlson, Advances in Electrorheological Fluids,
ed. M. A. Kohudic (Technomic Publishing, Lancaster, 1994), pp.30-52
71. K. O. Havelka, Progress in Electrorheology, ed. K. O. Havelka and F. E.
Filisko (Plenum Press, New York, 1995), pp.43-54
72. H. Block, and J. P. Kelly, A. Qin, and T. Watson, Langmuir, 6 (1990) 6
73. H. Block and P. Rattray, Progress in Electrorheology, ed. K. O.
Havelka and F. E. Filisko (Plenum Press, New York, 1995), pp.19-42
74. F. E. Filisko, Proc. 3rd Int. Conf. on Electrorheological Fluids, ed. R.
Tao (World Scientific, Singapore, 1992), pp.116-128
75. F. E. Filisko, Progress in Electrorheology, ed., K. O. Havelka and F. E.
Filisko (Plenum Press, New York, 1995), pp.3-18.
775

R H E O M E T R Y OF E L E C T R O R H E O L O G I C A L F L U I D S

Rex C. Kanu a and Montgomery T. Shaw b

aDepartment of Industry and Technology, Ball State University, Muncie, IN


47306

bDepartment of Chemical Engineering and Polymer Program, University of


Connecticut, Storrs, CT 06269-3136

1. R H E O L O G I C A L M E A S U R E M E N T S

1.1 Introduction
The purpose of this section on rheometry of ER fluids is to describe
methods that have been used for rheological characterization of ER fluids in
enough detail so that the experimentalist interested in studying ER fluids will
have a good start on the designs that have been tried, along with an
appreciation of their strong and weak poims. We will exclude engineering
tests in this section unless they result in true theological information.
Auxiliary characterization such as electrical and optical properties will be
covered briefly.
The inclination in rheometry is to adapt existing equipment to the task, and
with ER fluids this has been followed successfully. Most investigators have
taken commercial instruments and replaced existing fixtures with ones that are
electrically isolated from each other and from the rest of the instrument. Both
rotary and linear motions have been used successfully. Pressure-driven flows
have generally required custom-built equipment.
When working with ER fluids, two precautions are necessary: (1)
potentially lethal voltage sources are needed for most ER experiments and
safety interlocks thus become very important, and (2) high voltage connected
to even well-isolated fixtures can create stray currents to parts of the
instrument that should not be exposed to any currem. This may require
auxiliary grounding of parts leading from the fixtures to the bulk of the
instrument. For example, rather than counting on a ground through a motor
shaft or a load cell, a separate ground lead should be used [1]. Janocha and
776

Rech [2] have described some additional problems with high-frequency, high-
voltage fields.
Aside from protecting the experimentalist and the instrument, one should be
aware that the geometry of the fixtures, along with the dielectric properties of
all materials in the electric field influence the direction and strength of the
electric field. For example, if the classical parallel-plate geometry is used,
then the field near the edges of the plates will be concentrated near the corners
unless the sharp corners are rounded. The strength of the field will further be
influenced by the nature of the dielectric surrounding the plates (e.g., air vs.
fluid). At the electrical stresses typically required, air may break down [3].
According to Paschen's law, the breakdown voltage is a function of the
product of pressure x distance. At 1 ram, a field of 1 kV will break down air,
and fields of half this value will create corona discharge, adding to any
observed current. Thus successful designs will reduce the field before
encountering a fluid-air interface. Details concerning the design of cup-and-
bob viscometers have been described [4].
Suspensions are the most common ER fluids. Suspended particles in fluids
subjected to fields polarize, and then migrate together by attraction of the
electric dipoles. This process is indistinguishable to a more general process
called dielectrophoresis. Dielectrophoresis forces result when there is a
gradient in the electric field. The general relationship for dielectrophoretic
force on a sphere is:

F - 2~g o a3 ~2 - g l VE 2 (1)
g2 + 2 9 1

where E is the intensity of the electric field, e2 is the permittivity of the


sphere, el is the permittivity of the oil, e 0 is the permittivity of free space,
and a is the radius of the sphere. Note that the direction the sphere moves
depends on the sign of e2-~l and not on the direction of the field. Thus
particles of high dielectric constant relative to oil will migrate to the regions of
high field. Rheometers must therefore be designed to avoid large values of
VE 2 "
A larger force by far is the electrophoretic force on a charged particle in the
fields typically needed for ER experiments. By using ac fields, electrophoretic
migration can be avoided.
ER fluids also sediment and migrate due to the influence of gravity and
hydrodynamic forces, respectively. Sedimentation necessitates reloading,
777

restirring or other precautions. Kawai et al. [5] used a system to circulate ER


fluid from a reservoir through the rheometer, thus avoiding sedimentation.
Banding in torsional flows has been observed [6], and other flows may feature
similar hydrodynamic separation processes.

1.2 Simple Flow Observation


The simplest ER experiment is often used as a demonstration: a field is
placed across the fluid to solidify it, and then the voltage is reduced until the
fluid slumps or flows. The geometry often comprises two parallel cylindrical
wires. The electrical stress E between two parallel cylinders is spatially
varying and is the highest near the electrodes, with a magnitude of
approximately

AV (2)
E
2a In da

where A V is the voltage between wires of radius a spaced at distance d. The


mechanical shear stress o 21 is the highest also at the wires and will depend
upon the volume of fluid that has solidified. A rough estimate is

_ d 2 (3)
o" 21 Pg / 2 a

where p is the density of the fluid and g is the gravitational constant. By


progressively lowering the applied voltage, a value of yield stress can be
estimated.
This simple idea has been modified for a parallel plate arrangement where
the equations are now

AV (4)
E
H

and

o 21 = H p g (5)

where H| is the gap between the plates. Stipanovic and Schoonmaker [7]added
a Teflon block above the fluid (Figure 1), which increases the shear stress to
778

cy21 = (Hwtpg + W) / wt (6)

where W is the weight of the block, w is the width of the device and t is the
depth of the fluid beneath the block. Assumed here is that the width of the
block is the same as the width of the fluid, and there is no resistance to motion
contributed by the fluid between the block and the electrode, a good|
approximation because of the low electric field in that region. The Teflon
block increases the stress range and gives one the ability to develop a ~ 2~ vs.
E relationship over a| wide range of stresses without changing anything but the
weight of the Teflon block.
An alternative geometry for measurement of this kind is the annulus. Some
advantages of the annulus would be the fairly well-defined electrical field
combined with the ability to drive the fluid with gas pressure, which can be
easily varied during the experiment. Such arrangements diverge, however,
from the simplicity offered by the gravitational flow devices already described.

1.3 Sliding plate rheometers.


This class of rheometers involves a linearly driven plate that also serves
as the electrode. While the latter restriction is not imperative-the voltage of

PTFE
Block
j ,, z/
. , , /,

2 m m ~

Figure 1. Gravity-operated apparatus for determining the yield stress of ER


fluids (From Stipanovic and Schoonmaker [7])
779

a movable plate could be allowed to float between two other fixed electrodes-
most investigators have found it convenient to arrange the movable plate as
either the ground or the high-voltage electrode. Linear motions have been
applied by a variety of methods, usually using a tensile testing machine.

+C~~ Stationary Pivot Point

J:D jf-~,
~ j Microscope

) ~ To Microscope Stage

t E>
Shear
Light

Field of View "

Figure 2. Pendulum apparatus for applying small loads to ER fluids.


Simultaneous observation is possible in this device. From Sprecher, Chen and
Conrad, [8].

Another similar concept uses the weight of the plate itself by suspending the
top electrode on a pendulum and moving the bottom plate underneath this
swing (Figure 2) [8]. This arrangement has gives a nonlinear change between
the movement of the bottom plate and the stress, which is advantage in the
accurate determination of yield stress at low fields. Clearly if the bottom plate
is confined to a horizontal plane, the gap will change slightly; compensation
would be possible by mounting the bottom plate on another pendulum.
Still another sliding plate design uses axial annular drag flow between a
fixed, double cup and a light sleeve attached to the load cell mounted on the
cross head of a tensile testing machine (Figure 3) [9]. This geometry lends
itself to installation of optical paths in the field direction. Block et al. [10]
780

(+)
Teflon Insulator

Crosshead
T Crosshead Motion

II 'QI ,an,
! !

ner Electrode
,1-rr
-mm Gap (h)

ER
Io - 5 0 m m Fluid
uter Electrode

Figure 3. Axial annular drag flow device attached to a testing machine. From
Sprecher, Conrad and Carlson [9].

have described a horizontal sliding-plate rheometer that employs a very


compliant, spring-leaf load cell to push very gently against the sliding plate.
This apparatus was useful for determining the yield stress of the fluid, as
opposed to steady flow properties.

1.4 Rotational Viscometric Flows

1.4.1 Torsional Flow between Parallel Plates


The choice of geometry influences not only the homogeneity of the electric
field, but the nature and uniformity of the deformation of the fluid. While
781

parallel plates can give a nearly uniform electrical field, the deformation field
varies spatially. As important is the variation of the contribution of the
material at various radii to the observed torque. The outcome is that the
outermost edges of the fluid contribute the most to the torque reading, but may
experience the most nonuniform electric field. This will lead to torque
readings which will be difficult to interpret in terms of the stress in the fluid by
the conventional methods [11]. In addition, particle migration by
dielectrophoresis will be a problem. The advantages of the parallel plate
geometry include ease of construction, small sample size, and the relative ease
of adding optical paths. The normal force can also be measured. Under static
conditions the attractive normal force F provides a measure of the dielectric
properties of the fluid. The relevant equation is:

F- 1 ~cR2~ 0~ f E 2
-~ (7)

where 'R is the radius of the plates, and g f is the effective permittivity of the
ER fluid. Figure 4 shows the application of Equation 2 for an experimental
ER fluid. Note that the normal force is quite linear in E 2 up to quite high
fields. The value of g f derived from this plot is close to the value obtained
with a time-domain dielectric instrument [12].

1.4.2 Couette Flow


The Couette geometry features reasonable homogeneity of both electric and
deformation fields if the gap is small relative to the radius of the cup and bob.
This advantage has attracted many advocates [4, 5, 13-27]. In addition, the
Couette geometry offers large torque for a given diameter, making it
particularly appropriate for measuring yield stress [ 18, 28]. The disadvantages
are the usually large sample volume, and the inertia of the moving fixture. In
many cases it is somewhat more cumbersome to introduce optical paths into a
Couette cell than for parallel plates, although complex birefringence
measurements on ER fluids have been performed in a Couette geometry with
the light propagating in the neutral or vorticity direction, i.e., parallel to the
axis of the cylinders (see section 3.2 for more details). This is virtually
impossible with the parallel-plate geometry.
While Couette flows generally have lower edge effects than parallel-plate
flows, such effects apparently are not absent. They derive from a coupling of
the normal hydrodynamic effects with fringing of the electrical field. A study
by Janocha et al. [4] compared the response of Couette devices featuring
782

different bob geometries, including ones with conical bottoms and insulating
flat bottoms. The use of an insulating bottom removed the influence of the
contribution from the bottom area. While true with all ER geometries, these
authors found that stirring the fluid with the field off between each
measurement was essential to counteract the effects of dielectrophoretic
migration of the particles.
The Couette geometry is one of the best choices for the determination of
yield stress. Because applications of ER fluids may call for holding of parts or
fluids in position for a considerable time, the measurement of yield stress is of
importance to the technology. The subject of yield stress itself, irrespective of
the fluid, has generated lively debate; the reader is referred to, for example,
Astarita [29], Bennington et al. [30], Bonnecaze and Brady [31], Harnett and
Hu [32], Nguyen and Boger [28], Schurz [33], and Yoshimura et al. [18] to
sample the arguments. For ER fluids, most investigators favor the application
of a steadily increasing stress to the solidified ER fluid, finding the value that

1600

1200 55.35 Pa mm2/kV 2


g 8
0
13_
800 --
N
b

4OO

I I I
0.0 5.0 10.0 15.0 20.0 25.0
(E, kV/mm} 2

Figure 4. Attractive force between parallel plates is proportional to the static


composite permittivity ~c of the ER fluid. Slope is equal to Sc So/2. From
Jordan, McLeish and Shaw [ 1].
783

causes distinct flow. Torque-controlled drag-cup motors are most convenient


for this type of measurement. Application of torque to a rotor using a
Helmholtz coil working on a permanent dipolar magnet embedded in the bob
enabled Woestman [19] to apply a torque that would drop slightly if the fluid
yielded. This resulted in the fluid resolidifying. As this was repeated, the
fluid's yield stress increased by over a factor of three, suggesting that the fluid
acquires a more stable structure as the result of the disruption caused by the
slippage. The ratio between the static yield stress (the point where motion
started) and the dynamic yield stress (the point where the motion stopped) was
about 1.2.

1.5 Pressure-Driven Flows


Applications of ER fluids include valves wherein the ER fluid is required to
resist pressure. There are also many designs of devices featuring pumping or
squeezing flows that involve a combination of pressure-driven and drag flows.
While one might hope that these flows could be fully predicted from
rheometric drag flows, the nature of ER suspensions suggests that the
coincidence of a solid surface at the maximum shear stress can lead to special
characteristics. In addition, pressure-driven flows, especially in a thin
channel, can provide opportunities for close examination of the fluid's
structure by optical or other means. Because of the requirements for a
uniform electrical field, experimentalists have focused on channel flows [34,
35]. Axial, pressure driven, annular flows find use in applications [36] but are
less common in rheometry [37].

1.5.1 Channel (Slit) flows


Two rheometers using slit flows have been described fully in the recent ER
literature. One, due to Nakano et al. [35] uses a 10-mm-wide slit with a 0.6-
mm gap. The slit is 18 mm long, but only the last 10 mm are energized (this
was increased to 20 mm in later work [34]). Fluid is feed from a cylinder,
using a screw-driven piston, through a streamlined section into the slit. A
pressure transducer gives a measure of the shear stress. By measuring
pressure with and without the field, the cylinder and entrance pressure losses
can be found because these areas are not energized. Transients upon applying
the field can be related to the buildup of structure in the ER fluid.
The apparatus describe by Weiss and Carlson [38] is similar, but has the
added convenience that the fluid can flow back into the reservoir (a syringe) by
gravity at the end of the experiment, requiring very little fluid.
784

V/A 77v////////A..i.GV/////~ _ . . ~,.I

%
0

~ J

-=-V

Figure 5. Pressure-driven, annular axial flow device for use with ER fluids
[37]. The annular gap is between parts A and B. This material has been
reproduced from the Proceedings of the Institution of Mechanical Engineers,
Part C, Journal of Mechanical Engineering Science, 16 1974 by J. Arguelles,
H. R. Martin and R. J. Pick by permission of the Council of the Institution of
Mechanical Engineers.

A somewhat different approach was taken by Thurston and Gaertner [39]. In


their apparatus, an oscillatory pressure is applied along the slit. The slit is
vertical with the driver piston at the stop and the bottom resting in an open cup
785

of ER fluid. Because of inertia, this flow is complex; however, at low


frequencies in thin slits, the complex shear stress c~21 is given by the usual
equation:
9 _p,
o-21 H/2L (8)

where P* is the complex pressure, H is the slit height and L is the length of the
slit.

1.5.2 Axial annular flows


Arguelles et al. [37] have described an apparatus that exploits axial
pressure-driven flow in an annular space between two cylinders, labeled A and
B in Figure 5. The fluid, driven by air pressure of up to 500 psig (350 kPa),
flows from the reservoir Co through holes C1 into volume C2 and then through
the gap labeled h. It exits through opening O.

1.60therSound Propagation
Korobko and Chernobai [40] depict a chamber in which a sending and
receiving transducer are arranged to propagate sound parallel to the electric
field. Both velocity and attenuation could be measured, although only velocity
measurements were reported. Fields ranged up to 2 kV/mm. The sound
velocity was linearly proportional to the field and did not show a threshold,
even at particle volume fractions as low as 2 %.

2. E L E C T R I C A L MEASUREMENTS

Of concern in this section will be measurements of the dielectric properties


of ER fluids during flow, and preferably during the measurement of shear
stress. The measurement of steady, transient or periodic current as a result of
be considered a routine accessory of the rheometer. Of more value to the
diagnosis of the structure of the fluid is the dielectric spectrum of the fluid
during flow. In addition, and also of technical value, is the measurement of
dielectric strength of the fluid, again under flow conditions.

2.1 Complex permittivity


While several groups have measured dielectric properties of ER suspensions
under quiescent conditions (e.g., [25, 41]), measurement of permittivity under
flow conditions is far from trivial. A Couette cell designed to do this has been
786

described by Block et al. [17]. By using a spli.t-stator design, their apparatus


could measure shear stress simultaneously and provide guard electrodes
thereby. According to the descriptions, high voltage could be applied to the
electrodes, but details were not provided. Separation of the high-voltage
activating potential from the low-voltage measuring signal has been discussed
by Placke et al. [42], but circuit diagrams were not provided. Their apparatus
centered around a commercial rotational rheometer with a rotating cup and
fixed bob.

2.2 Ultimate dielectric properties


The ER effect obtainable in a fluid can be limited by the dielectric strength
of the suspension. This may well be a function of shear rate. Inoue [20]
apparently made measurements of the ultimate dielectric properties under flow.

3. O P T I C A L P R O P E R T I E S

Of the methods for diagnosing structure of quiescent and flowing ER


suspensions, optical methods have proven to be both adaptable and
informative. Optical methods range from direct observation with a microscope
to light scattering, to measurement of dichroism and birefringence. Many
groups have been successful at combining one or more of these methods with
the simultaneous measurement of rheological properties. Some of the methods
will be explored below.

3.1 Direct observation under flow


The observation of suspensions during continuous flow is not trivial; but, by
using shallow depth of focus or sheets of light, individual particles, or sets of
particles, can be observed. Another approach is to use dilute suspensions,
with the appreciation that the structures . Exploiting one or a combination of
these techniques, many researchers have arranged cells whereby the deforming
suspensions can be observed [43]; a typical apparatus is illustrated in Figure 6.
However, most of these do not simultaneously provide rheological
information. Exceptions are the pendulum apparatus of Sprecher et al. [8] and
Bossis et al. [6].
The observation of periodic oscillatory flow is somewhat easier than steady
flow, because illumination can be synchronized with the motion of the
particles. Equipment for such an experiment has been described by Bossis et
al. [6], and this is reproduced in Figure 7. In this apparatus, motion is derived
787

I
0
I Q
J
I

/
Delrinbase I

~la/s slid.~ ' l ,1\


Stainless steel plates Motor
Figure 6. Apparatus for observing with a microscope the deforming structure
of ER fluids. Direction of observation is perpendicular to the field. From
Jordan, Shaw and McLeish [ 1].

from the output of the accelerometer, whereas the current to the vibrator is
proportional to the load. The light passes parallel to the field, which is applied
to the suspension using tin-indium oxide coatings on the glass plates. The gap
between the two plates is kept small, e.g., 100-200 ~tm.

3.2 Birefringence and Optical Dichroism.


As light passes through a material, it is either transmitted, scattered or
absorbed. The scattered and transmitted light can be measured directly; the
absorbed light by difference. In anisotropic systems, such as aligned ER
fluids, the absorption and scattering can be expected to depend on the direction
of the particle columns with respect to the direction of the polarization of the
light. Similarly, one expects the refractive index of the medium to be
anisotropic, which will result in birefringence. Birefringence and dichroism
can be interpreted as phenomena resulting from the real and imaginary
components of the complex refractive index tensor, and thus are related. See
Fuller [44] for more details, including typical optical arrangements.
788

Equipment for measuring birefringence under electrical fields has been


described by Block et al. [22] and by Smith and Fuller [24]. Both arrange for
the light to propagate in the vorticity direction in a Couette cell with the field
applied in the direction of the velocity gradient. The apparatus of Smith and
Fuller shown in Figure 8 is set up to give the dichroism, which shows dramatic
changes on application of the electric field.

3.3 Light Scattering


Systems can be chosen that absorb very little light, thus simplifying the
interpretation of the measured transmitted and scattered light. From these
measurements, inferences concerning the spatial variation in structure, and the
changes with time (say, after applying the field or shearing the fluid) can be
made. Unlike microscopy, the signal in a scattering experiment is independent
of the bulk motion of the fluid, an important advantage in flowing systems.
However, the information concerning structure is never as complete or certain
as with direct observations.

I/ / JI Phase detector Camera


Sine-wave generator
, !\
i '" Microscope
Electromagnetic vibrator
I
I

1 X o cos (~t - 6)
E , ; , ~ / / /i/
I
~ / 1
///Glass Plates

l .
. . . . .
. . .
/ ,/ //,
/

," ,, ,~
~_/ , i~///

Micro mete r h i , " ","" ":"'~"/x,~


Z, , ~ / \ ) 1/

Accelerometer ~

~ Chopped light source

Figure 7. Design for observing the structure of ER fluids along the field
direction while subjecting the fluid to oscillatory deformation. From Bossis et
al. [6]
789

l Xenon Arc Lamp


I
Monochrometer
l
f "~ Lens

i i 11 Polarizer
Wave Plate

E-field ~ FlowCell

-I CPU I I I PMT
I ldc ~r4co
iI
LPF

Figure 8. Equipment for observing the dichroism developed by an ER fluid.


From Smith and Fuller [24].

Suspensions such as ER fluids tend to be highly scattering. As a result of


this, light that is scattered tends to be rescattered, which convolutes the signal
even more. If the path length is large enough, but not large enough for the
light to strongly absorbed, then the light that appears to be "transmitted" has
actually been scattered many times; it diffuses through the suspension like a
gas though a porous medium. At this limit, the observed transmittance no
longer obeys Beers law [45, 46]. All information concerning structure is lost,
save a single length scale corresponding to the mean-free path of the light. At
larger path lengths, the transmittance begins to fall exponentially because of
absorption, and obeys the rule"

T ~ ~10l* e _L/labs (9)


3labs
790

where T is the transmittance (ratio of transmitted light to incident light) labs is


the absorbance length, 1" is the diffusion length and L is the path length. From
the variation of T with L, both length scales (/*and labs) can be found. Path
lengths up 7 mm for commercial ER fluids have been reported [45] using the
equipment depicted in Figure 9. Deformation was accomplished by moving
the first electrode back and forth.
To derive more information about the structure of ER fluids, most
researchers use ER fluids comprising particles and liquid of nearly equal
refractive index. (To reduce multiple scattering, short path lengths are another
option, but path lengths are limited by the particle size. A third method-
diluting the suspension-is not available because of the strong effect that
particle concentration has on the behavior of ER fluids.) If the refractive
indices of the particle and suspending fluid are matched, then the fluid is
nearly transparent and the full two-dimensional light scattering patterns can be
observed and interpreted in terms of well-established scattering theory [47].
An example of such equipment, with the light transmitted perpendicular to the
field (vorticity direction in a Couette geometry) is shown in Figure 10. The
path length through the oriented fluid was 2 mm. While deformation was
available, torque could not measured simultaneously.

3.4 Other optical techniques


Non-centrosymmetric media should generate a weak, but measurable signal
at twice the frequency of an incident light beam. Such media could include ER

XY A2
A1 -

Laser
"I - - ~

<
/ /
f

\
\

L1
L2

Figure 9. Diagram of apparatus for measuring diffuse transmittance of ER


fluids in the field direction. From Ginder [45].
791

Figure 10. Apparatus for measuring the time-resolved, two-dimensional light


scattering for light propagating perpendicular to the field [26]. Reprinted with
permission from J. E. Martin and J. Odinek, Journal of Rheology, 39(5) 995-
1009 (1995). Copyright 1995 Society of Rheology

fluids subjected to DC fields [48]. The strength of the second harmonic


generation (SHG) should reflect the nature of polarization in the ER fluid.
Figure 11 shows schematically an apparatus for measuring SHG in an ER
fluid. No provision was made for deforming the solidified fluid, but this could
be added. Time and field are the principal variables in SHG-ER experiments.

4. SUMMARY

While ER fluid rheometry presents special problems, equipment has been


developed that addresses these problems and opens up opportunities for
investigating structural changes in other two-phase materials as well. The
relationship between structure and stress in highly structured materials such as
ER fluids is a constant challenge for rheologists, and techniques will continue
to evolve. Imaging methods for studying the details of particle motion will
undoubtedly be developed. One unfortunate aspect of ER fluids is that
792

presently NMR imaging techniques, used successfully for other suspensions,


cannot be applied because of the magnetic fields produced by the current
passing between the plates.

Sample Box
HIGH
/ 0 VOLTAGE
1--.

IR Absorber
I l IR Filter

532nm IF
rY
w
< PMT
._1

0
<
i>-
BOXCAR
COMPUTER

Figure 11. Equipment for studying second harmonic generation in an ER


fluid. From Wu et al. [48].

REFERENCES

1. T . C . Jordan, M. T. Shaw, T. C. B. McLeish, J. Rheol., 36 (1992) 441.


2. H. Janocha and B. Rech in R. Tao and G. D. Roy (eds.)
Electrorheological Fluids, Mechanism, Properties, Technology and
Applications, World Scientific, Singapore, 1994, pp. 344-357.
3. R. Bartinkas and E. J. McMahon (eds.) Engineering Dielectrics Vol. 1,
Corona Measurement and Interpretation, Am. Soc. For Testing and
Materials, Philadelphia, 1979.
793

4. H. Janocha, R. Bolter, and B. Rech, Paper presented at the 5th


International Conference on ER Fluids held in Sheffield, UK, 7/1995.
5. A. Kawai, K. Uchida, K. Kamiya, A. Gotoh and F. Ikazaki, Advanced
Powder Technol. 5, (1994) 129
6. G. Bossis, E. Lemaire, J. Persello, and L. Petit, Progr. Colloid Polym.
Sci. 89 (1992) 135.
7. A. Stipanovic and J. Schoonmaker, ACS Polym. Prepr., 35 (1994) 365.
8. A. Sprecher, Y. Chen and H. Conrad, in J. D. Carlson, A. F Sprecher,
and H., Conrad (eds.), Proc. 2nd Inter. Conf. ER Fluids, Technomic
Publishing Co., Inc. Lancaster, P A , 1990, pp. 82-89.
9. A . F . Sprecher, J. D. Carlson, and H. Conrad, Mater. Sci. Eng., 95
(1987) 187.
10. H. Block, J. P. Kelly, A. Qin and T. Watson, Langmuir, 6, (1990) 6.
11. G. V. Gordon and M. T. Shaw, Computer Programs for Rheologists,
Hanser Publishers, New York, 1994.
12. R. Kanu and M. T. Shaw, in Progress in Electrorheology, K. O. Havelka
and F. E. Filisko (eds.) Plenum Press, New York, 1995, pp. 303-323.
13. C. J. Gow and C. F. Zukoski IV, J. Colloid Interface Sci., 136 (1990)
175.
14. H. Uejima, Japan. J. Appl. Phys., 11 (1972) 319.
15. G. V. Vinogradov, Z. P. ShuI"man, Yu. G. Yanovskii, V. V.
Barancheeva, E. V. Korobko, and I. V. Bukovich, J. Eng. Phys., 50
(1986) 429.
16. W. M. Winslow, J. Appl. Phys., 20 (1949) 1137.
17. H. Block, E. M. Gregson, A. Qin, G. Tsangaris, and S. M. Walker, J.
Phys. E: Sci. Instrum., 16 (1983) 896.
18. A. S. Yoshimura, R. K Prud'homme, H. M. Princen, A. D. Kiss, J.
Rheol., 31 (1989) 699.
19. J. T. Woestman, Phys. Rev. E, 47 (1993) 2942.
20. A. Inoue, in J. D. Carlson, A. F Sprecher and H., Conrad (eds.), Proc.
2nd Inter. Conf. ER Fluids, Technomic Publishing Co., Inc. Lancaster,
P A , 1990, pp. 176-183.
21. T. Y. Chen and P. F Luckham, Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 78 (1993) 167.
22. H. Block, E. M. Gregson, W. D. Ions, G. Powell, R. P. Singh, and S.
M. Walker, J. Phys. E.: Sci. Instrum. 11 (1978) 251.
23. Y. Kojima, T. Matsuoka, and H. Takahashi, J. Appl. Polym. Sci., 53
(1994) 1393.
794

24. K. Smith and G. Fuller, J. Colloid Interface Sci., 155, (1993) 183.
25. M. Jordan, A. Schwendt, D. A. Hill, S. Burton, and N. Makris, J.
Rheol. 41 (1997) 75.
26. J. E. Martin and J. Odinek, J. Rheol. 39 (1995) 995.
27. K. Tanaka, R. Akiyama, and K. Takada, Polym. J., 28 (1996) 419.
28. Q. D. Nguyen and D. V. Boger, Rheol. Acta, 26 (1987) 508.
29. G. Astarita, J. Rheol., 34 (1990) 275.
30. C. P. J. Bennington. R. J Kerekes, and J. R Grace, Can. J. Chem. Eng.,
68 (1990) 748.
31. R. T. Bonnecaze and J. F. Brady, J. Rheol., 36 (1992) 73.
32. J. P. Harnett and R. Y. Z. Hu, J. Rheol., 33 (1989) 671.
33. J. Schurz, Rheol. Acta, 29 (1990) 170.
34. M. Nakano, R. Aizawa, and Y. Asako, paper presented at the 5th
International Conference on ER Fluids, Sheffield, UK, July, 1995.
35. M. Nakano and T. Yonekawa, in R. Tao and G. D. Roy (eds.)
Electrorheological Fluids, Mechanism, Properties, Technology and
Applications, World Scientific, Singapore, 1994, pp. 477-489.
36. D. Brooks, Int. J. Modern Phys. B, 6 (1992) 2705.
37. J. Arguelles, H. R. Martin and R. J. Pick, J. Mech. Eng. Sci. 16 (1974)
232.
38. K. D. Weiss and J. D. Carlson, in R. Tao (ed.), Proc. Int. Conf.
Electrorheol. Fluids, 1991, World Scientific, Singapore,1992, pp. 264-
279.
39. G. B. Thurston and E. B. Gaertner, J. Rheol. 35 (1991) 1327.
40. E. V. Korobko and I. A. Chernobai, J. Eng. Phys. 48 (1985) 153.
41. Y. Kim, and D. Klingenberg, in K. O. Havelka and F. E. Filisko (eds.),
Progress in Electrorheology, Plenum Press, New York, 1995. pp. 115-
130.
42. P. Placke, R. Richert, and E. W. Fischer, Colloid Polym. Sci. 273 (1995)
848.
43. E. Lemaire, G. Bossis, Y. Grasselli, and A. Meunier, in C. Gallegos, A.
Guerrero, J. Mufioz and M. Berjano (eds.), Progress and Trends in
Rheology IV, Verlag, Darmstadt, 1994, pp. 140-142.
44. G. G. Fuller, Optical Rheometry of Complex Fluids, Oxford University
Press, New York, 1995.
45. J. M. Ginder, Phys. Rev. E, 47 (1993) 3418.
795

46. J. M. Ginder and L. D. Elie, in R. Tao (ed.), Electrorheological Fluids:


Mechanisms, Properties, Structure, Technology and Applications., World
Scientific, Singapore, 1992, pp. 23-36.
47. T. C. Halsey and J. E. Martin, in R. Tao and G. D. Roy (eds.)
Electrorheological Fluids, Mechanism, Properties, Technology and
Applications, World Scientific, Singapore, 1994, pp. 115-128.
48. J. Y. Wu, L. K. Shen, and L. W. Zhou, in W. A. Bullough (ed.), Proc.
5th Int. Conf. Electro-Rheological Fluids, Magneto-Rheological
Suspensions and Associated Technology, World Scientific, Singapore,
1996, pp. 698-703.
797

SOME APPLICATIONS OF NON-NEWTONIAN FLUID FLOW

J. w. Hoyt

Mechanical Engineering Department


San Diego State University

1. NON-NEWTONIAN FLUID FRICTION R E D U C T I O N -


OVERVIEW

1.1 Introduction
Drag reduction in the turbulent flow of solutions of polymers or
surfactants has been studied for almost 50 years, beginning with the
discoveries of Toms [1] and Mysels [2]. The scientific and technical aspects
of these non-Newtonian fluid flows are still under active study, since the
underlying mechanisms involved in these flow effects, like those of
turbulence itself, are only dimly understood.
Nevertheless, a great deal of empirical information has been obtained,
outlined in several recent books: Sellin and Moses [3]; Gyr and
Bewersdorff [4]; and Choi, et al [5].
There are also numerous reviews: Hoyt [6]; Virk [7]; Shenoy [8]; and
Kulicke, et al [9] among many others. A very comprehensive bibliography
listing over 4,900 references has recently appeared: Nadolink and Haigh
[10]. Conference proceedings are another excellent source of current
information: the most recent being sponsored by the American Society of
Mechanical Engineers (Hoyt, et al, eds., [11]).

1.2 Basic Drag-Reduction Concepts


If the pressure loss per unit length in turbulent pipe flow is less than that
found with a Newtonian fluid such as water, the fluid is said to be drag-
reducing. Solutions of high molecular weight linear polymers (above, say
798

50,000), surfactants forming aggregates of rod-like micelles, and fiber


suspensions all reduce the pressure drop or fluid friction, and thus can be
inferred to interfere somehow with the three-dimensional fluctuations u',
v', and w' associated with turbulent flow.
There are at least three amazing aspects of the drag-reduction effect:
1) Drag reduction only occurs in turbulent flow, defined (in pipe flow)
as VD/v (the Reynolds number) larger than 2000 or so, where V is the bulk
fluid velocity in the pipe, D is the inside diameter of the pipe, and v is the
kinematic viscosity of the liquid.
2) Only a small amount of high polymers, surfactants, or fibers are
required to dramatically reduce the friction. Solutions of less than I part-
per-million of the most effective polymers show large friction reductions,
while surfactant and fiber concentrations of several hundred ppm give
substantial reductions.
3) Smaller quantifies of additive are required for a given effect as the
molecular weight is increased, or (in surfactants) the aggregate-forming
ability is increased. The better additives have molecular weights in the
millions, or form aggregates of similar molecular weight. If the polymers
are degraded, or the aggregates broken up, by exposure to high shear
stresses, the drag-reduction effect disappears.
Savins [12] introduced the term "drag reduction", defined in pipe-flow
as a percent:

DR, % = ([AP~- APpI/APs} xl00

where AP~ is the pressure drop per unit length in a pipe flowing the solvent
alone and APp is the pressure drop for the same flow rate, using additives.
Another commonly found term is the "engineering friction factor", ~,,
defined as:

1
= [D aP] / [Tp V2]

where D is the pipe diameter and p is the fluid density. Friction factor -
Reynolds number plots are extremely useful in demonstrating the presence
(or absence) of drag reduction, since for a Newtonian fluid, the
relationship was deduced by Karman and Prandtl. A convenient
computing formula for the Karman-Prandtl (pure solvent) value is:
799

1/~/~ = 1.8 log [Re/6.9]

Comparing values obtained with this relationship with data from


additive-solution flow at the same Reynolds number leads again to the
drag-reduction percentage, since

DR, % = [{~s- Xp} / ~,~]x 100

The Karman-Prandtl relation is for a smooth pipe only; textbook


expressions can be used for pipes known to be rough in nature. Care must
be taken in examining literature values for friction factor, since chemists
and occasionally other scientific workers may use the "Fanning" friction
factor which has a value 1/4 that of the D'Arcy or engineering factor
defined above.
The maximum value of drag reduction which can be obtained by using
additives seems to be about 80%; a more refined estimate has been
suggested by Virk [13], which is often referred to in the literature as "Virk's
asymptote".
Some of the most effective water-soluble drag-reducing polymers such
as poly(ethylene oxide) and polyacrylamide can be obtained in very high
molecular weights (ca. 5 million or more) and their properties have been
extensively studied in laboratory investigations. Their use in industrial or
commercial applications has been extremely limited due to the breakdown
or degradation of the polymer molecules caused by shearing or agitation in
pumps or other mechanical devices. The fragile long chains (which are
extended in dilute solution) are readily fractured by passing through
pumps. Another factor is the fairly high cost of the polymers, which limits
their application to the transport of high value products. As explained in
the next section, these objections to the use of polymers to reduce pipe
friction are overcome in petroleum pipelines, where there may be
hundreds of miles between pumping stations, and an increased flow rate is
highly advantageous.
Drag-reducing aggregates of surfactant molecules are also broken up
by passing through pumps, but unlike polymer molecules, the aggregates
reform when the shear stresses are relaxed, returning to their original
effectiveness. Economic considerations suggest that surfactant solutions
are best fitted to recirculating flow systems. Thus surfactants may play an
800

important role in systems such as large-scale domestic heating or cooling,


as described in a following section.
The concepts outlined here in a very general form are given as an
introduction to the more detailed applications of polymers and surfactants
in fluid flows found in the following sections.

2. OIL-PIPELINE FRICTION R E D U C T I O N

2.1 Introduction
The use of high-polymer additives to reduce the pipeline friction of
crude-oil and petroleum products has been a spectacular success, and
currently forms an important part of industrial petroleum transport
technology. Several factors account for this success. First, although the
drag-reducing polymers are not inexpensive, their use in parts-per-million
quantities to greatly increase the flow of a very high-value product is often
much more economical than the capital and operating expense of
installing additional pumping capacity. Secondly, the additive can be
applied with relatively minor portable equipment, which can be easily
removed in case flow augmentation is no longer needed. Another factor is
that there is no need to remove or treat the additive-containing product in
any special way - the additives disappear in the refining process, or are
innocuous in other pipeline products. Finally, of importance in crude-oil
transport in cold climates, the greatly reduced heat transfer of additive-
containing fluids helps keep the product viscosity at a lower level.
These advantages have led to installations of drag-reducing additives
at more than 80 locations worldwide (Motier, et al, [14]), since the initial
trials in the Trans Alaskan Pipeline in 1979. Steady improvements in
polymer properties have accelerated this use; Motier, et al, show that
currently only about 1/14 the quantity of polymer (compared with the 1979
additive) is needed for the same amount of friction reduction.

2.2 Additive Characteristics


The commercial chemicals used as drag reducing additives are highly
proprietary. Motier and Carrier [15] suggest that the additives are o~-
olefin polymers or copolymers; the molecular weights are extremely high -
35 million or more. A surprising aspect of these high molecular weight
801

polymers is their resistance to shear degradation; in the 1.2 m dia Trans


Alaskan Pipeline, high drag reduction is maintained over a 353 km section
with initial polymer concentrations currently on the order of a few ppm.
The additives are supplied commercially as 10% solutions in a carrier
similar to kerosene. With this polymer concentration, the resulting fluid is
extremely viscous; injection pumps require gas pressure boosting at the
suction side to function properly. No particular injection nozzles or
equipment seems to be required - the additive is simply pumped into the
pipeline and disperses via turbulence. Higher polymer-concentration
suspensions are under study to reduce the supply volume required.
Although initially there was concern that additives might present a
problem in crude-oil refining, the o~-olefin polymers appear perfectly
acceptable to refinery processing, and at the current levels of pipeline
application (1-2 ppm; Motier, et al, [141] would be practically undetectable
anyhow.

2.3 Applications
The first and probably the major success story in polymer drag
reduction was (and is) the Trans Alaskan Pipeline System (TAPS). The
highly competitive and proprietary nature of the petroleum industry has
meant that very little detail regarding applications is available in the open
literature, but Burger, et al [16] have given some actual test data on
flowing Alaskan crude through 2.66, 5.25, 33.4, and the TAPS 119.4 cm dia
pipes. From these d a t a , a rather convoluted scheme based on estimated
molecular parameters was used to predict the TAPS performance from
laboratory data on the smaller pipe sizes. Although the scheme worked
well enough to warrant installation of drag reduction in the TAPS
pipeline, the much simpler method described later in this Chapter, based
on hydrodynamics, leads to even better predictions as shown in Figure 1.
From Berretz, et al, [17]; Beaty, et al, [18]; Hom, et al, [19]; and Motier,
et al [14], we learn that the initial success of drag reduction lead to the
cancellation of plans to build two additional pumping stations on the TAPS
pipeline. The 1.45 million barrels per day pumping capacity of the pipeline
was raised to 2.1 mbpd by polymer injection. Above 50% drag reduction in
the 1287 km long pipeline is achieved with 28 ppm or less polymer. The
logistics of providing some 95,000 liters per day of 10% polymer solution in
kerosene to remote locations in Alaska are impressive.
802

0.100 II II I _ -- IIII I I

.<

d
o
.w
o
4-1

0.010
o
-r4
4J
O

h~

0.002 . ,I . I . . . . . . .

I 000 I O, 000 I 00,000 I , 000, O(

Reynolds number

Figure 1. Pipe-flow data for Alaska crude oil with 10 ppm drag-reducing
additive. Actual data for 3 pipe sizes, with scaling from the two smaller
(2.664 and 5.25 cm dia) to the full-scale (119.4 cm dia) pipeline. Adapted
from Burger, et al [16] tabulated data. I"1 2.664 cm pipe dia; O 5.250 cm dia;
A 119.4 cm dia; E1'119.4 cm prediction from 2.664 cm dia data; (~ 119.4 cm
prediction from 5.250 cm dia data; ---Karman-Prandtl friction line for
Newtonian fluids.

Similar formulations have now been applied to other crude oil pipelines
(Beaty, et al, [18,20]; Lester, [21]; Motier and Prilutski, [22]; Motier, et al,
[14], as a apparently viable commercial technique. The low installation
cost, and impressive flow increase performance, together with the "use
only when needed" feature has lead to increased acceptance in crude oil
production, replacing expensive fixed assets which may have only a short
usage requirement.
Polymer additives have been used with great success in petroleum
product pipelines. Carradine, et al, [23]; Motier, et al, [24]; Muth, et al,
[25-27]; and Motier and Carrier [15] describe applications to diesel oil,
803

gasoline, and natural gas liquids pipelines. Drag reductions of over 40%
appear to be routine, and this offers the possibility of eliminating marginal
pumping stations.

2.4 S u m m a r y
The use of high-polymer additives to enhance flow in petroleum pipe
lines has been a great commercial success. Additive effectiveness has been
improved so that only a few parts-per-million are required to give friction
reductions of 50% or more. The economics of polymer addition seem to be
very favorable in the pipeline transport of high value products such as
petroleum, reducing the need for pumpiv.g stations as well as increasing
throughput.

3. SURFACTANT APPLICATIONS TO LARGE-SCALE HEATING AND


COOLING SYSTEMS

3.1 Introduction
Many regions are served with district-wide heating or cooling systems,
which, in the case of heating for example, may involve transporting hot
fluid over considerable distances. The heat source is usually lower-grade
energy from electrical generating stations or industrial operations which
would otherwise be rejected to the environment. Offsetting this low-cost
source is the energy requirement to pump the hot fluid, typically several
kilometers or more, to heat exchangers servhag domestic needs. The use of
surfactants as drag reducers to reduce this substantial pumping
requirement is under intensive study. Mound 7% of the heating
requirement in Germany, 35% in Sweden, and as much as 40% in Denmark
and Finland is centrally supplied, while district heating serves six hundred
thousand households in Seoul, Korea. Hence the opportunities for energy
saving are plentiful. Moreover, in new designs, smaller transmission
pipes might be used, if drag reduction by surfactants achieves acceptance.
3.2 Surfactants - Basics
Surfactants are a broad class of surface-tension reducing chemicals,
characterized by having both a hydrophilic (watersoluble) and a
hydrophobic (water-repelling), often oil-soluble, component in the same
molecule. In solution, depending on the detailed chemistry, surfactants
804

may be ionic or non-ionic. The anionic group forms the basis of household
detergents, which are produced in enormous quantifies world-wide.
Cationic and non-ionic surfactants are less well-known, but readily
available industrial chemicals.
Thirty years ago, Savins [28] showed that some anionic surfactants
could greatly reduce the turbulent pipe friction as compared with the
water in which they were dissolved. At higher shear stresses, the drag
reduction disappeared and the friction was identical to water.
Remarkably, when the shear stress was lowered, the solutions were again
found to be drag-reducing. Many surfactants of various ionic classes have
now been found to be friction-reducing. All studies show that pipe friction
reductions of up to 80% can be observed, but at higher shear stresses, the
drag reduction effect abruptly disappears, promptly reappearing when the
shear stress is lowered.

3.3 Surfactant Micelles


Originally it was thought that drag reduction in surfactants was due
to the formation of rod-like (or worm-like) "micelles", which caused the
effect, but were then broken up and reformed as the shear stress exceeded
some critical level and then was reduced. Ohlendorf, et al [29] have now
shown that the rodlike micelles take appreciable time (up to hours) to form
initially, and are not broken up in high shear-stress situations. Rather, the
rods aggregate into much larger structures, which can align themselves in
the flow direction, or form a network in the flow which results in a
lowering of the turbulent friction. When the aggregates are exposed to
shear stresses exceeding a certain level, they are dispersed (i.e. the
aggregate binding forces are overcome) and the drag-reduction effect is
lost. However, unlike long-chain polymers, the surfactant micelles will
reform again (in a few seconds) into the drag-reducing aggregates when
the shear stress is reduced below the critical level. Thus the drag-reducing
ability of surfactant solutions remains essentially intact after passing
through pumps, valves, etc., which would destroy the effectiveness of
longchain polymer solutions.
The detailed chemistry involved in micelle formation and the
subsequent aggregation is quite complex. Apparently, micelles are initially
roughly spherical, and begin to form at a surfactant concentration which
depends on the surfactant itself, as well as temperature and presence of
805

counter-ions. This "critical micellular concentration" (CMC) is often on


the order of a few hundred parts-per-million or less. At first, the
hydrophobic regions of the individual molecules group together to
minimize their area exposure. Electrostatic forces also play a role, but in
the presence of counter-ions the initially spherical micelles (at least in
drag-reducing surfactants) tend to rod-like form, at somewhat higher
concentrations. Ohlendorf, et al [29] estimate the rod length for a typical
drag-reducing surfactant micelle to be around 10 to 40 nm, depending on
temperature. These rods form at a surfactant concentration from two
to five times the "CMC", again depending on the temperature.
Explanation of the existence of aggregations of these rod-like forms is less
straight-forward.
The two principal phenomena which suggest aggregation are shear-
thickening and drag-reduction itself. In Figure 2 we see how the viscosity
of a typical drag-reducing surfactant solution suddenly jumps from
approximately that of water to a value 10 times larger as the shear stress
is increased. This astonishing behavior must be due to the micellular rods
(which do not overlap in the solvent at rest) suddenly forming a network
or a lengthy aggregate. Flow birefringence studies suggest the latter.
That significant drag reduction occurs in turbulent flow is another
indication of lengthy aggregates.
In Figure 3 we note the pipe-flow behavior of a drag-reducing
surfactant solution, with the engineering friction factor, ~, ,plotted as a
function of the Reynolds number VD/v, where V is the average flow
velocity in the pipe, D the diameter, and v the viscosity of the solvent
(water). While the laminar pipe flow friction factor is somewhat higher
than water (due to higher viscosity as noted before), in the turbulent region
we notice a large reduction compared with the water solvent. This action
is similar to that of high-molecular-weight polymer solutions, where it is
thought that, in turbulent flow, the polymer coils unwind to form lengthy
strands which act to dampen the u', v', W r velocity fluctuations in the flow.
Similarly, turbulent surfactant flow also dampens these fluctuations, thus
supporting the idea that aggregates occur, and that they are of extended
length. Further, at a certain higher Reynolds number the friction factor
abruptly returns to the pure water value, suggesting that the aggregates
are dispersed by the more violent velocity fluctuations occurring as the
Reynolds number is increased. Curves such as Figure 3 are typical of drag-
806

10

q ,mPa

m,

Y. ! s e c
0.1 !
I | ,i J |1

10 10 10 10 4

Figure 2. Viscosity versus wall shear rate for 0.1% solution of C~6 TASal a t
35~ adapted from Ohlendorf, et al [29].

reducing surfactant flow; individual curves are functions of temperature,


surfactant type and concentration, counter-ion type and concentration, as
well as pipe diameter.

3.4 Basic Laboratory Experiments


Many surfactants of various ionic classes have been identified as
friction-reducers, and Shenoy [30] has given an extensive review of earlier
work. Work has continued in an effort to provide surfactants for district-
heating with improved temperature range, lesser environmental hazard,
lower cost, etc. Determination of the most useful surfactant for a
proposed application is a complicated balance involving shear stress in the
pipe, surfactant and counter-ion concentration, and range of operating
temperatures. The drag reduction which is obtained is a non-linear
function of all of these variables for a given surfactant.
Table I lists many of the surfactant formulations under active study for
large-scale district heating.
807

Table 1
Drag-Reducing Surfactants

Cationic

C14TASal myristyltrimethylammonium salicylate


C16TASal cetyltrimethylammonium salicylate
C16TAC cetyltrimethylammonium chloride
C16TAC (Arquad 16-50) with NaSal counter-ion
C18TAC (Arquad 18-50) with NaSal counter-ion
Ethoquad 0/12 with NaSal counter-ion
(Oleyl-N(CH3)(C 2H40H)2C1 )
Ethoquad T 13-50 with NaSal counter-ion
(tallow-tris hydroxyethyl ammonium chloride)
Habon G hexadecyl dimethyl hydroxyethyl ammonium-3-
hydroxy-2-naptholate (n-alkyl; n = 16)
Obon-G n - 18; also with added NaSal
Dobon-G n = 22

Zwitterionic

N-alkylbetaines with N = 15 or 17,


plus Na dodecylbenzenesulphonate

Selection of a surfactant compound for a given application is a balance


involving cost, environmental considerations, and useful lifetime, as well
as performance at various temperatures as a drag reducer, in pumps and
heat exchangers, valves and meters, etc. Laboratory tests have focused on
several of these factors.
Figure 4 (Chow, et al, [31]) shows the effect of operating temperature
on drag reduction performance of a typical alkyl trimethylammonium
chloride (Arquad 18-50) with equal weight NaSal counter-ion. For this
surfactant combination, the maximum drag reduction (70-80%) was
obtained at higher Reynolds numbers as the temperature was increased
from 30 to 90 ~ C. At 100 ~ C the drag reduction effect greatly decreased. A
808

0.080 I I ! I
I

0.040
X Water

o
~<
0.020
0
o
U.
r
0
o
Symbol
U.
A 0.090 in. - ~ ?

0.008
-

o 0.500 in.

0.004
10 3 10 4 10 5
Reynolds Number

Figure 3. Drag reduction with an equimolar solution of C~6 TABr and 1-


napthol to give a total concentration of 508 ppm. Adapted from White [51].

wide operating range is desirable in practical systems. Figure 5 (Schmitt,


et al, [32])shows that, if protected from oxidation, temperature recycling
from 30 to 120 ~ C gave reproducible drag-reduction results with a 1000
ppm solution of surfactant. Each temperature cycle extended over a
period of 34 hours.
Other work with C~4TASal and C16TASal solutions has been reported by
Bewersdorff and Ohlendorf [33], Bewersdorff [34], and Harwigsson, [35].
Recent studies using Arquad and Ethoquad compounds include Lu, et al
[36], Myska and Zaldn [37], and Usui, et al, [38]. Park, et al [39] found no
effect of a centrifugal pump on the friction-reducing properties of several
809

[c,a
I I I iG jlll l ji . . . . it- iii i _ i i _

7",," N - - CH 3 Cl
I
CH 3
C
o6
u

- i

0 SYMBOL T ( o C) iL
i i

o .30
+ 40

Z 50
o 70
9 80
X 90
o 1 O0 D -- 0.243 iN.
6
f~
I 10" 10 5 10 a
Reynoi(:J$ Number
Arcluod 18/50 -~- NoSol (2000 / 2000 IDIDm)

Figure 4. Effect of solution temperature on drag reduction for 2000 ppm


surfactant. At 100~ much of the drag-reduction effectiveness is lost.
Adapted from Chou, et al [31].

surfactant solutions, including C~6TASal and Habon-G, confirming earlier


results of Steiff, et al [40] and Gaslievic and Matthys [41].

3.5 Environmental Considerations


In spite of only moderate toxicity of most of the surfactants under study
for district heating applications, the large quantities involved make
environmental considerations very important. Harwigsson, et al [42]
estimate that about 60% of the 7 x 105 m 3 of water circulating in Swedish
domestic heating systems must be replaced annually due to leaks and
maintenance. At 500 ppm of surfactant, this would mean about 200 tons of
surfactant discharged into the environment each year. Therefore, more
benign and easily biodegradable surfactants having drag reducing
qualifies are being sought. Hellsten, et al, [43] have suggested Zwitterionic
surfactants as being possible candidates. (zwitterionic surfactants contain
810

DR
[%]

80-
L
6O

z.O i-

20
,,...,

40 60 80 100 120 T[oC]


Figure 5. Results of a temperature recycling test with a +10~ per hour
rate of temperature change. An oxygen stabilizer, hydrazene, has been
added to the 1000 ppm surfactant solution to avoid degradation. Each test
cycle lasted 34 hours, flowing the solution through a 12.7 mm dia pipe.
Adapted from Schmitt, et al [32].

both positive and negative charges.) Compounds investigated by Hellsten


exhibited drag reduction, but only over a limited range of temperatures.
A contrasting opinion is offered by Steiff, et al, [44], who present data
suggesting that Habon-G and Dobon-G are environmentally acceptable in
Germany, and solutions of these surfactants would require no further
811

treatment before discharge into a sewerage system. As a background in


considering the environmental concerns of surfactants in district heating
systems, it should be remembered that these systems usually contain
substantial quantifies of other substances, such as the oxygen scavenger,
hydrazene, so that evaluation of the effects of leakage, etc, becomes more
complex.

3.6 Heat Transfer Considerations


As noted in the Section on oil-pipeline friction reduction, drag-reducing
additives reduce heat transfer in pipes even more spectacularly than the
friction. Thus in the long-distance transport of heated fluid with
surfactants to reduce the pipe resistance, the overall heat loss also can be
reduced, offsetting this advantage is the problem of extracting the heat a t
the point of use. Some of the suggestions for overcoming this problem are
listed below:

Use a heat exchanger of the cross-flow type

Increase the velocity in the heat exchanger

Break up the micelles just before the heat exchanger

Cross-flow or plate-fin heat exchangers may be advantageous in


surfactant-solution heat exchangers, since the boundary-layer effect found
in pipes is reduced. Hoyt and Sellin [45] found little difference in cross-
flow heat transfer in a drag-reducing polymer solution compared with
water, when the results are correlated using a Reynolds number based on
the increased viscosity of the polymer solution. In large-scale tests,
however, as explained in Steiff, et al [44], it has been found that there is
still a substantial reduction in effectiveness in these types of heat
exchangers. Pollert, et al [46] concur in these results, but note that if the
fluid velocity through the exchanger is increased, there seems to be no loss
of heat transfer effectiveness.
Kawaguchi, et al [47] suggests that bringing the surfactant solution to a
temperature a few degrees higher before entering the heat exchanger
could destroy the micelles and thus restore the heat exchange properties of
the fluid to that of water. Others have suggested mechanically breaking
812

up the micelles by mixers or valves, before entering the heat exchange


region. The practicality of these ideas remains to be tested.
Contrary to expectations from laboratory results, field tests (Steiff, et
al, [40], Pollert, et al, [46]) seem to show little or no influence of the
surfactant additives on the overall system heat transfer. While this may
suggest that the principal thermal resistance was not on the surfactant
side of these heat exchangers, it cannot be assumed that this is the usual
case.
Matthys [48] has given a thoughtful review of current understanding of
the heat-transfer problem in the flow of drag-reducing fluids,
supplemented more recently by experimental work (Gasljevic and
Matthys, [49]).

3.7 Large-Scale Demonstrations


Brief reports on actual use of surfactants in district heating systems
have become available. Steiff, et al [40] describe results from full-scale
tests in Volkingen, Germany, which appear to confirm the energy savings
expected due to decreased pumping-power requirements. Similarly,
Pollert, et al [46] found in a demonstration in Kladno-Krocehlavy, Czech
Republic, that pumping power was reduced by 40% when Habon-G was
employed. Interestingly, the surfactant retained its properties in the
system for the two winter-heating seasons studied. As mentioned above,
no effect on heat exchanger operation was noted in these two
investigations.
Gasljevic and Matthys [50] give a very complete report on the use of
surfactants in the cooling system of a large building. This application of
surfactants is even more challenging than district heating since the many
fittings, valves and heat exchangers add much more complexity to the
flow patterns than the long straight runs expected in district systems. In
this chilled water system, a reduction of about 30% was achieved in the
required hydraulic pumping power. As expected, the heat transfer, both in
the "chiller" (where the circulating water is cooled by refrigerant), and in
individual room heat exchangers, was diminished. However, the
reduction was less than might be expected, due to large heat transfer
resistances found on the refrigerant and air sides of the heat exchangers.
In other words, the principal barrier to heat exchange was not the chilled
surfactant solution, but rather the other fluid components of the system.
813

A complex cooling system such as that studied by Gasljevic and Matthys


involves numerous components which interact in complicated ways,
making analysis difficult, and somewhat uncertain. Nevertheless, the field
test was regarded as very successful, no doubt leading to further
implementation of the use of surfactants in chilled-water cooling systems.

3.8 Summary
The use of surfactants in district and building heating and cooling
systems is an emerging technology with great promise for significant
energy savings. While there seems no doubt that the pumping-power
requirements of such systems can be greatly reduced by the use of
surfactants, appropriate formulations to avoid environmental concerns
and equipment designs to enhance heat transfer are important next steps
in securing commercial acceptance.

4. SCALE-UP

By scale-up, we mean predicting the pressure drop in large industrial or


commercial pipelines based on results from laboratory-scale small pipes.
Scale-up for the use of drag-reducing polymer solutions in long-distance
off pipelines or central-station heating or cooling schemes is an important
current problem.
Astarita [52], in an elegant article tracing references back to Vitruvius (35
B.C.), declared that a non-trivial scaling theory cannot be constructed for
non-Newtonian fluids. Many attempts to provide scale-up information
for drag-reducing fluids have tended to confirm this gloomy conclusion, by
introducing graphical or iterative procedures (Granville, [53,54]; Matthys
and Sabersky, [55]; Sellin and Ollis, [56]; Taylor and Sabersky, [57]), or by
admittedly empirical methods (Savins and Seyer, [58]; Gasljevic and
Matthys, [59]).

However, quite recently, accurate scale-up procedures for the same


fluid (homologous scaling, according to Astarita) have become possible,
based on better knowledge of the basic fluid dynamics involved in such
flows. We start by recalling the velocity profile for drag-reducing fluids"
814

24 ~ -

22

20 . - -

18 ---

16 . . . .

14 --
AB
12
/o
12 p p m I
10 ---

' J-
2 r

i1~

6 ppm

0 0.05 0.1 0.2 0.5 1.0


U ~ (fps)

Figure 6. Delta B as a function of u* for poly(ethylene oxide) in a one inch


d i a m e t e r pipe.

u+ =Alny
815

where A and B are constants having the approximate values of 2.5 and 5.0,
and AB is a function of the drag-reducing substance, its concentration, pipe
roughness, etc. The problem has been that AB is a highly non-linear (and
non-predictable) function which must be determined by experiment.
Figure 6 shows a typical result.
An assumption made in most scale-up techniques is that the value of AB
is the same for equal values of shear stress (or friction velocity, u*) in both
small and large pipes. Based on this assumption, Hoyt and Sellin [60]
introduced a relatively simple scale-up calculation method which views
the drag reduction as an "negative" analogy to the familiar pipe roughness
employed in Newtonian fluid flow. An even more recent advance has
obviated the need for such an analogy.
Anderson, et al, [61] provide data from a large-scale experiment which
demonstrates conclusively that, for equal values of shear stress, AB has the
same value in both small and large pipes. Figure 7 is an example of
measurements of AB plotted against the friction velocity (u*), for pipes
differing in diameter by a factor of 6. The same polymer solution was
delivered to the test pipes and the measurements made simultaneously,
side-by-side, so that the only difference was pipe size. Similar results were
obtained over a wide range of polymer concentrations. Scaling using the
invariance of the AB-u* relationship proved to be very precise. Hence we
can greatly simplify the scaling relations.
Extending Prandtl's logarithmic velocity distribution law by including
AB, we find that:

1/~/~,- 2 log [Req~,/2.51] + AB/~/8

where X is the engineering friction factor for flow in pipes and Re is the
Reynolds number VD/v. Solving for AB, and letting subscript I indicate
the small pipe, and 2 the larger,
AB~ = q8[1/q~-2 log {Req~/2.51 } ]
and

AB2 = ~/811/~/X2/2 log {ReqX2/2.51}].


816

5O

e D,4.IS'
40 0 0.,1~6"

30
m
g
<1
Oo
20

10
o

O , L , I . ,' 9 ! , I 9 1~ J . l , , , l ~ l ! ] , t , 1 . ~ , ! 9 I ~ I 9 J . [ . l . l . t , l '

2 3 4 se7 10-1 2 s 4 s87 100

U,-ftJs
Figure 7. Delta B as a function of u* for p o l y a c r y l a m i d e p o l y m e r solution
in two different size pipes. The m e a s u r e m e n t s were m a d e simultaneously,
side-by-side, using 33 p p m solution from the same reservoir.

Equating AB's and noting that Req~, - u*Dq8/v, where D is the pipe
diameter:

1/~/s - 1/~/)q +2 log [u*D2"48/2.51v] - 2 log [u*D1~/8/2.51v]

Hence

1/'4K 2 - 1 / ~/k~+ 2 log [D2/D,].


817

0
U o
0 o
c 10-2 0
0
.--

u
+ _

@@
@ 0=6.15" Meosured @
0 D=l.026"bteasured
0 0=6.15"Predicted
, : , t, t,l,t,l ! 1 t t t] , ', , t ,,,I,~,1 ! 1 ! I II t , t .,,!,~,

104 2 3 45 105 2 3 + 5 106 2 3

ReynoldsNumber
Figure 8. Friction factor as a function of Reynolds number for the data of
Figure 7, together with scaling predictions for the 6.15 inch dia pipe flow,
based on the data from the 1.026 pipe. Black dots are the actual test data
for 6.15 inch dia.

The friction factor k2 so obtained from a plot of ~a vs Re1 is to be replotted


at the same shear stress, and so, equating u* from the Reynolds number
expression given above"

Re2 = [ ~/X~/4K2] [D2/D~]Rel.

This extremely simple scaling procedure yields excellent results as


shown in Figure 8, where the data from a 1.026 inch diameter pipe test has
818

been scaled up to 6.15 inch diameter and compared with the four available
test points. Similar agreement of scaling and actual test data was
obtained for five other polymer concentrations, which suggests that the
procedure is reliable, as well as being grounded in fundamental
engineering relations. Astarita will be surprised!

The above scaling procedures work very well if the smaller pipe is 1/2
inch (13 mm) or more in diameter. As shown in Hoyt and Sel!in [60], the
velocity profile in pipes of lesser diameter is significantly different from
that of larger pipes, thus voiding similarity scaling. If only very small
diameter pipe-flow data are available, the empirical approach of Savins
and Seyer [58] has given useful results. Savins and Seyer replace the
actual shear stress of the polymer solution flow with that of the pure
solvent. In effect, this modifies the velocity profile to approximate that of
the larger-scale polymer pipe. If data from very small pipes are the only
available source of scaling information, the Savins and Sayer scheme
probably remains the only alternative. It has been used with considerable
success in off pipelines (Lester, [62]).

5. FLOW TRACERS

5.1 Introduction
Many techniques with varying levels of sophistication have been
proposed as methods of showing the flow path, vorficity, or velocities of
turbulent flow in boundary layers or around various objects. The simplest
idea is a dye streak, but dyes suffer from very rapid dissipation and are
almost useless in turbulent flow. Hence very elaborate techniques, such as
Particle Image Velocimetry, have been brought forward to provide
turbulent flow information. However, the equipment cost of the lasers,
shutters, cameras, etc., involved in these methods is beyond the reach of
many investigators.
By incorporating non-Newtonian properties into a dye-streak, a highly
effective flow tracer has been developed. Such a tracer is of great interest
since it allows one to follow turbulent and separated flows currently
inaccessible with simple methods.
819

5.2 Tracer Development


The non-Newtonian tracer has been developed at Bristol University
(Hoyt and Sellin, [63-65]) using ideas based on drag-reduction technology.
The basic concept is to incorporate both the shear-induced-state (SIS),
sometimes referred-to as shear thickening" or "strain-hardening" flow
aspects found in surfactant solutions, with the high "extensional viscosity"
or "thread-drawing" properties of polymer solutions. These are then used
to form (with a colorant) a tracer fluid which can be ejected into the
turbulent flow as a dye-streak which resists dispersion and breakup while
following the flow path.
Key to the development of the tracer is the amazing viscosity- shear
stress relationship of certain surfactant solutions. As shown earlier in this
chapter, solutions of C~6TASal (as well as C~4TASal) exhibit a sudden
order-of-magnitude jump in viscosity as the shear rate is increased from a
fairly low value. This attribute is used in the tracer to stabilize the ejected
strands of colored fluid, so that they remain coherent in rapidly changing
stress fields.
The tracer threads are held together axially by the high extensional
viscosity provided by high-polymers such as poly(ethylene oxide) or
polyacrylamide. This contribution is very important as otherwise the
surfactant strands would tend to snap. A final component of the tracer is a
small amount of white latex wall paint to act as a colorant. The paint also
acts to add additional stability to the ejected tracer threads. Since the
tracer is a mixture of surfactant, polymer, and paint suspension, the
rheology is rather complex, and largely unknown.
Most of the development work has been with the following formula, or
slight variations thereof:

2% C16TASal 500 ml
1/2% PEO 250 ml
Tap Water 1000 ml
White Paint 5 ml

The 2% C~6TASal solution is made up from 2% (by weight)


cetyltrimethylammonium bromide, mixed with 2% by weight sodium
salicylate. C~4TASal (made from myristyltrimethylammonium bromide
and NaSal in the same manner) gives a more responsive tracer thread;
820 .

mixtures of the two can be employed. As explained in a previous section


of this chapter, more dilute solutions of these compounds have been
studied for possible use in district heating systems.
PEO is polyethylene oxide) (Polyox WSR-301, Union Carbide), a
familiar drag-reducing additive. A~I/2% solution, made up by dispersing
the dry powder in methanol, and gently mixing in deionized water,
provides adequate elogational viscosity, evidenced by a high degree of
"thread drawing" (noticed when pouring the solution from a container).
Other high polymers such as polyacrylamide can be used.
The tracer components are stored separately and mixed together
shortly before use. The mixed tracer is dispensed from a rake of I m m ID
hypodermic tubing as shown in Figure 9, or from individual tubes, fed by
gravity or a small pump. The non-Newtonian character of the tracer is
shown by the slight "die-swell" as the fluid leaves the hypodermic tubes.

Figure 9. Method of dispensing tracer from rake of hypodermic tubing.

Figure 10 shows flow around a 51 mm dia cylinder at a Reynolds


number of 11,500. A mirror above the flow channel allows the three-
dimensional characteristics of the flow to be observed simultaneously with
the view from the side. The 3-D flow visualization obtained in this
manner is at present unavailable from any other technique.
A frame from a video recording using the tracer is shown in Figure 11.
A much more diute solution of tracer has been used in order to show the
smaller scales of the turbulent flow around a 51 m m dia cylinder at a
821

Figure 10. Side view of flow around 51 m m dia cylinder at Re = 11,500.


Plan view is seen in mirror above.

Reynolds number of 12,600. The turbulent character of the boundary-layer


flow leaving the cylinder, as well as details of the vortices being formed,
can be revealed by frame-by-frame examination of the video recording.
This flow tracer method permits the use of ordinary film and cameras,
as well as "home" video, and of course, visual observation. It is hoped that
the technique will find a useful place in education, inasmuch as it gives
more detail than presently available from computer simulations or other
flow visualization methods. The only disadvantage is that, in a
recirculating flow facility, drag-reducing polymer is returned to the test
section.
822

Figure 11. Video frame of flow around 51 mm diameter cylinder at Re =


12,600.

DEDICATION

This Chapter is dedicated to the memory of Preston Lowery HI, Chen


Liang, and Constantinos Lyrintzis, Engineering faculty colleagues and
friends at San Diego State University, whose lives were snuffed out by a
deranged graduate student, August 15, 1996.

ACKNOWLEDGMENTS

The support of the University of Bristol and the U.S. National Science
Foundation in preparing this chapter is very greatly appreciated. The
confidence in this work expressed by Dr. Michael Roco of the N SF,
through Grant No. CTS-9508409, is gratefully acknowledged.

REFERENCES

@ B.A. Toms, in Proc. 1st Int. Congress on Rheology, North-Holland


Publ., 2, 1948, 135.
823

2. K.J. Mysels, Flow of Thickened Fluids. U.S. Patent 2 492 173 (1949).
3. R.H.J. Sellin and R.T. Moses (eds.), Drag Reduction in Fluid Flows,
Ellis Horwood, Chichester, 1989.
4. A. Gyr and H.-W. Bewersdorff, Drag Reduction of Turbulent Flows by
Additives, Klewer Academic, Netherlands, 1995.
5. K.-S. Choi, K.K. Prasad, and T.V. Truong (eds.), Emerging
Techniques in Drag Reduction, IMechE, London, 1996.
6. J.W. Hoyt, Trans ASME, J. Basic Engrg., 94 (1972) 258.
7. P.S. Virk, AICHE Journal, 21 (1975) 625.
8. A.V. Shenoy, Colloid Polymer Sci., 262 (1984) 319.
9. W.-M. Kulicke, M. Kotter, and H. Grager, in Advances in Polymer
Science 89, Springer-Verlag, Berlin, (1989) 1.
10. R.H. Nadolink and W.W. Haigh, Applied Mechanics Reviews, 48
(1995) 351.
11. J.W. Hoyt, et al (eds.), Symposium on Turbulence Modification and
Drag Reduction, ASME, FED-Vol.237, New York, 1996.
12. J.G. Savins, Soc. Petrol. Eng. Journal, 4 (1964)203.
13. P.S. Virk, H.S. Mickley, and K.A. Smith, Trans ASME, J. Applied
Mechanics, 37 (1970) 488.
14. J.F. Motier, L.-C. Chou, and N. Kommareddi, in Proc. Symposium on
Turbulence Modification and Drag Reduction, ASME FED-Vol.237
(1996) 229.
15. J.F. Motier and A.M. Carrier, in Drag Reduction in Fluid Flows, Ellis
Horwood, Chichester (1989) 197.
16. E.D. Burger, L.G. Chorn, and T.K. Perkins, Journal of Rheology, 25
(1980) 603.
17. M. Berretz, J.G. Dopper, G.L. Horton, and G.J. Husen, Pipeline and
Gas Journal (September 1982).
18. W.R. Beaty, R.L. Johnston, R.L. Kramer, L.G. Wamock, and G.R.
Wheeler, in Third International Conf. on Drag Reduction, Bristol
University (1984) F.1.
19. A.F. Horn, J.F. Motier, and W.R. Munk, in Drag Reduction in Fluid
Flows, Ellis Horwood, Chichester (1989) 255.
20. W.R. Beaty, R.L. Johnston, R.L. Kramer, L.G. Warnock, and G.R.
Wheeler, Oil & Gas Journal, 82 (Aug. 13, 1984) 71.
21. C.B. Lester, Oil & Gas Journal, 83, (1985) 51; 76; 107; 116.
824

22. J.F. Motier and D.J. Prilutski, in Third International Conference on


Drag Reduction, Bristol University (1984) F.2,1.
23. W.R. Carradine, G.J. Hanna, G.F. Pace, and R.N. Grabois, Oil & Gas
Journal, 81 (1983) 92.
24. J.F. Moiler, D.J. Prilutski, Z.-J. Shanti In, and R.J. Kostelnik, in Third
International Conference on Drag Reduction, Bristol University
(1984) F.3.
25. C.L. Muth, C.J. Stansberry, G.J. Hussen, M.H. Lewis, and M.S.
Ziobro, Pipeline & Gas Journal, (June, 1985) 38.
26. C.L. Muth, T.U. Hannigan, R.S. Vruggink, and G.F. Pace, Pipeline &
Gas Journal, (June, 1986) 26.
27. C.L. Muth, M.J. Monahan, and L.S. Pessetto, Pipeline Industry, (July,
1986) 43.
28. J.G. Savins, Rheologica Acta, 6 (1967) 323.
29. D. Ohlendorf, W. Interthal, and H. Hoffman, Rheologica Acta, 25
(1986) 468.
30. A.V. Shenoy, Colloid & Polymer Science, 262 (1984) 319.
31. L.-C. Chou, R.N. Christensen, and J.L. Zakin, in Drag Reduction in
Fluid Flows, Ellis Horwood, Chichester,1989, 141.
32. K. Schmitt, F. Durst, and P.O. Brunn, in Drag Reduction in Fluid
Flows, Ellis Horwood, Chichester, 1989, 205.
33. H.-W. Bewersdorff and D. Ohlendorf, in Turbulent Shear Flows 5,
Cornell University (1985) 9.41.
34. H.-W. Bewersdorff, in Proc. Symposium on Turbulence Modification
and Drag Reduction, ASME FED-Vol.237, (1996) 25.
35. I. Harwigsson, "Surfactant Aggregation and its Application to Drag
Reduction". Ph.D. Thesis, Lund University, 1995.
36. B. Lu, Y. Talmon, and J.L. Zakin, in Proc. Symposium on Turbulence
Modification and Drag Reduction, ASME FED-Vol.237 (1996) 169.
37. J. Myska and J.L. Zakin, in Proc. Symposium on Turbulence
Modification and Drag Reduction, ASME FED-Vol.237 (1996) 165.
38. H. Usui, T. Itoh, and T. Saeki, in Proc. Symposium on
Turbulence Modification and Drag Reduction, ASME FED-Vol.237
(1996) 159.
39. S.B. Park, H.S. Suh, S.H. Moon, and H.K. Yoon, hi Proc. Symposium
on Turbulence Modification and Drag Reduction, ASME FED-Vol.237
(1996) 177.
825

40. A. Steiff, W. Althaus, M. Weber, and P.-M. Weinspach, in Drag


Reduction in Fluid Flows, Ellis Horwood, Chichester, 1989, 247.
41. K. Gasljevic and E.F. Matthys, in Recent Advances in Non-Newtonian
Flows, ASME, AMD-Vol.153 (1992)
42. I. Harwigsson, A. Khan, and M. Hellsten, Tenside Surf. Det., 30
(1993) 174.
43. M. Hellsten, I. Harwigsson, C. Blais, and J. Wollerstrand, in Proc.
Symposium on Turbulence Modification and Drag Reduction, ASME
FED-Vol.237 (1996) 37.
44. A. Steiff, K. Klopper, B. Zeidler, W. Althaus, and P.-M. Weinspach, in
Proc. Symposium on Turbulence Modification and Drag Reduction,
ASME FED-Vol.237 (1996) 235.
45. J.W. Hoyt and R.H.J. Sellin, Experimental Heat Transfer, 2 (1989) 113.
46. J. Pollert, P. Komrzy, K. Svejkovsky, J. Pollert, Jr., and J.L. Zakin, in
Proc. Symposium on Turbulence Modification and Drag Reduction,
ASME FED-Vol.237 (1996) 31.
47. Y. Kawaguchi, Y. Tawaraya, A. Yabe, K. Hishida, and M. Maeda, in
Proc. Symposium on Turbulence Modification and Drag Reduction,
ASME FED-Vol.237 (1996) 47.
48. E.F. Matthys, in Drag Reduction in Fluid Flows, Ellis Horwood, 1989,
129.
49. K. Gasljevic and E.F. Matthys, in Developments in Non-Newtonian
Flows, ASME, AMD-Vol.75 (1993) 101.
50. K. Gasljevic and E.F. Matthys, in Proc. Symposium on Turbulence
Modification and Drag Reduction, ASME FED-Vol.237 (1996) 249.
51. A. White, Nature, 214 (1967) 585.
52. G. Astarita, J. Non-Newtonian Fluid Mechanics, 4 (1979) 285.
53. P. Granville, in Proc. 2nd Int. Conf. on Drag Reduction (1977) BI.1.
54. P. Granville, in R.H.J. Sellin and R.T. Moses, eds. 3rd Int. Conf. on
Drag Reduction, Bristol University (1984) C3,1.
55. E. Matthys and R. Sabersky, Int. J. Heat Mass Transfer, 25 (1982)
1343.
56. R.H.J. Sellin and M. Ollis, I & EC, Product R & D, 22 (1983) 445.
57. D.D. Taylor and R.H. Sabersky, Letters in Heat and Mass Transfer, 1
(1974) 103.
58. J. Savins and F. Seyer, Phys. Fluids, 20 (1977) $78.
826

59. K. Gasljevic and E.F. Matthys, in Proc. Symposium on Development


and Applications of Non-Newtonian Flows III, ASME, San Francisco
(1995).
60. J.W. Hoyt and R.H.J. Sellin, Experiments in Fluids, 15 (1993) 70.
61. G.W. Anderson, J.J. Rohr, and J.W. Hoyt, in Proc. Symposium on
Turbulence Modification and Drag Reduction, ASME FED-Vol.237
(1996) 19.
62. C. Lester, Oil & Gas Journal, 83, No.5 (1985) 51.
63. J.W. Hoyt and R.H.J. Sellin, Experiments in Fluids, 20 (1995) 38.
64. J.W. Hoyt and R.H.J. Sellin, in Engineering Turbulence Modelling and
Experiments 3, Elsevier, Amsterdam, 1996, 381.
65. J.W. Hoyt and R.H.J. Sellin, in Proc. Symposium on Turbulence
Modification and Drag Reduction, ASME FED-Vol.237 (1996) 225.
827

PAPER COATING RHEOLOGY

D. W. Bousfield and A. Co

Department of Chemical Engineering, University of Maine


Orono, ME 04469-573 7 USA

1. INTRODUCTION
Paper and paperboard are often coated to improve the appearance of the sheet
and the quality of the print of the final product. Products such as magazines,
catalogs, labels, and consumer packaging are often coated. The coating is
normally composed of a pigment such as kaolin or calcium carbonate, a latex
binder, and a soluble binder such as starch. Coatings are formulated at high solids
content in order to minimize drying requirements and to improve quality. The
coating must be mixed, pumped, recirculated, and metered onto the moving paper
web. The final metering or coating operation is the critical step in the process; the
coated layer must be free from defects and near a target coat weight. The
rheology of the coating is an important issue in the control, design, and operation
of the coating process.
A roll applicator followed by a blade on the same roll is a common method to
apply coating onto the web. In the last 15 years, the short dwell applicator system
became popular; coating is pumped to a pond in front of the blade and the blade
meters on the final coating amount as depicted in Figure 1. The short dwell
applicator has one advantage short contact time and, consequently, minimal
water penetration into the web. Uses of fountain or jet applicators and "metered
size presses" are recent trends to apply coatings. High coat weights are normally
obtained by a rod metering step followed by an air knife component. However,
the blade metering element is a common last step to control coat weight and is the
focus of this chapter and much of recent research.
There are a number of unique aspects of paper coating that make this operation
complex. The web is rough, porous, and absorbent. In addition, it changes its
physical properties upon contact with water. The coating step is often
accomplished at speeds higher than 15 m/s. The final coating layer thickness
ranges from 7 - 30 ktm. The web is compressible, along with the rubber backing
828

~'\"\,\
Coated web

T
Excess to pond

Figm'e 1. Schematic of paper coating operation with a short dwell coater. The
region around the blade tip is enlarged to show the details of the blade tip.
Coating is applied to the web in the pond. The blade controls the final amount
of coating applied to the web.

that supports the blade loading. The blade is slightly flexible, it is mounted and
loaded in various ways, and it wears during operation. All of these factors,
including the high speeds, cause paper coating to be a complex operation.
Economic driving forces cause the coating to be applied at high solids.
Application at high solids reduces energy costs for drying and is known to
produce a better quality product in terms of light scattering and printing
properties, as reported by Van Gilder et al. [1 ]. Coating of high solids gives the
paper less chance to roughen upon contact with water and the coating layer does
not sink into the paper web. However, high solids content naturally brings
challenging rheological problems. There is a trade-off between solids and
operational speeds. Figure 2 shows data from Taylor [2] that is typical in the
industry; as the coating solids increase, the maximum speed for defect-flee
operation decreases. In order to increase the speed of a specific coating line, the
solids content of the coating often needs to be reduced. This reduction increases
drying costs and lowers quality of the product. Some pigments have different
"runnability" limits. Ground calcium carbonate pigments can often be coated at
arotmd 70% solids, whereas delaminated kaolin clays need to be coated at around
829

16 9 , , I 9 9 . I 9 . 9 I 9 I I I 9 9 9 I . , 9

-~ 14-
E
v 12-
.1
(J
1
o 10-
>
.o 8.

E 6.

.E 4. o
x
0
~ 2-

0 9 " " I " " ' I " " ' I " " " I " " " I " ' "

0.38 0.40 0.42 0.44 0.46 0.48 0.50


Solids V o l u m e Fraction

Figure 2. The relationship between coating solids and the maximum web
velocity. The open circles are data from Taylor [2]. The line is a model
calculation from Bousfield [3].

55% solids. In any case, there is a constant incentive to operate at high solids and
not to have operational difficulties.
The operational difficulties or "runnability" problems vary widely in the
industry, but common problems are the scratches, streaks, or skips in the coating
layer and the buildup of coating on the blade. The buildup on the blade in itself is
not a problem, but when the buildup breaks off the blade and ends up on the web,
problems in the calendering and printing operations develop. The blade buildups
are ot~en called "whiskers", "stalagmites", "weeps", or "spits", depending on the
operator and the nature of the building (dry or wet). Other operational problems
are related to coat weight control, both in the machine and cross-machine
directions. Another common problem is web breaks; these cause a loss of
production and correlate to high blade forces. Many of these problems are related
to rheology of the coating.

2. R H E O L O G I C A L TESTS OF COATING COLORS


The most common industrial test to monitor the viscosity of a coating
formulation is an apparatus with a spindle rotating in a container of fluid.
Although this type of apparatus cannot measure the "true" viscosity of non-
830

Newtonian fluids, it is used extensively in the coating industry because of its ease
of use and its value as a tool for quality control. However, to understand the
effects of the components of a coating formulation on its rheological
characteristics and to relate these rheological characteristics to its performance in
a coating operation, numerous researchers have made use of various standard
rheological tests. These include steady-state measurements in shear flow,
measurements in small-amplitude oscillatory shear flow, and transient
measurements in shear flow.
Of these measurements, viscosity measurement in steady-state shear flow is the
most commonly used. Examples of recent works are Triantafillopoulos and
Grankvist [4], Carreau and Lavoie [5], Purkayastha and Oja [6], Ghosh et al. [7],
and Ghosh [8]. To obtain the viscosity function over a wide range of shear rates,
one usually has to utilize several viscometers: the cone-and-plate system for low
shear rates, the concentric cylinders configuration for medium to high shear rates,
and the capillary type for high shear rates. An important experimental
consideration is that there is sufficient time for the coating to reach its equilibrium
configuration at a given shear rate. Also of interest is the yield stress of the
coating colors. Its importance in the actual coating operation is minimal since the
operation is performed at very high shear rates. However, it may play an
important role in the recovery stage after the coating application.
The behavior of the viscosity versus shear rate curve depends on the
components of the coating colors. Figure 3 shows the viscosity curves of a starch-
containing formulation and a formulation with latex particles (Roper and Attal
[9]). The curve of the starch-containing formulation exhibits a power-law shear-
thinning region and a region of almost constant viscosity at high shear rates. On
the other hand, the formulation with latex particles shows a region with shear-
thickening or dilatant behavior. Beyond the power-law shear-thinning region, the
viscosity rises sharply with increasing shear rate and then decreases slightly at
higher shear rates. Roper and Attal [9] indicated that the dilatant behavior
depended on the solids level and the latex particle size; the extent of dilatant
behavior was increased with higher solid levels and larger latex particle size.
Aside from the solid levels and the size and shape of solid particles, the
polymer additives dissolved in the suspending medium also have considerable
effects on the rheological behavior of the coating suspension. Examples are
shown in Figure 4, which depicts the viscosity curves of Carreau and Lavoie [5]
for kaolin suspensions with three different concentration levels of CMC (carboxy
methyl cellulose). They attributed the large increase in viscosity to the
interactions between kaolin and CMC, since the viscosity of the CMC solution
itself did not change significantly at these concentration levels.
831

101 I I , I I I I I

~, 1 0 o.
t~
n

(/)
O
o
.~-
> 10-1.. m

1 0 .2 I I ~) I I I I
10 -~ 10 0 101 1 2 10 3 10 4 10 s 10 s 1 7
Shear Rate (l/s)

Figure 3. The viscosity functions of a strach-containg coating (open circle) and a


coating with latex particles (open square). Data are from Roper and Attal [9].
4
10 I I I I i I

10- [] 11'

2
A 10-
w,
n 1 []
-~- 10-

0
10-

10 "1.

-2

10 ; I I ;0 01 ;2 ;3
1 -3 10-2 10-1 1 1 1 1
7, to (1Is)

Figure 4. Effect of CMC content on the viscosity and the dynamic viscosity for
44% vol. Kaolin suspensions. Data are from Carreau and Lavoie [5].
832

Also shown on Figure 4 are the dynamic viscosity versus frequency curves of
the suspensions. These are obtained from small-amplitude oscillatory shear
experiments. The dynamic viscosity rt'((o) is determined from the componem of
the shear stress that is in-phase with the rate of deformation and rt"(oJ), from the
component that is out-of-phase. Figure 4 shows that the viscosity curves and the
dynamic viscosity curves coincide, that is, rl(8)--r/'((o~,:~. This is analogous to the
Cox-Merz rule for polymer solutions or melts, which predicts that the magnitude
of the complex viscosity (r rl * ((o)I-x/rt;' ((o)+ rt"((o)) is equal to the viscosity at
corresponding values of frequency and shear rates. This empirical analogy can be
useful in predicting viscosity data when only linear viscoelastic data are available.
Alternately, linear viscoelastic data can be expressed in terms of the storage
modulus G' (-rl'a, ) and the loss modulus G" (-rt'a,). Figure 5 shows the storage
modulus of kaolin suspensions at various concentration levels of CMC, as
reported by Carreau and Lavoie [5]. As in the viscosity curves, a small
concentration level of CMC increases the storage modulus drastically, again due
to the interaction between kaloin and CMC. Another observation is that the
storage modulus approaches a constant value at low frequencies, a behavior that
is more like a viscoelastic solid. Similar behavior was also observed for the
coating formulations considered recently by Ghosh [8].
3
10 a , ... , I ,

1%CMC
2
10-
A

n
0.25 % CMC

1
10-

I
~00---0-0-(~ 0 ~ ' 0 ~ ~ ~ CMc
!

o
10 j
10" 10 "1 1() ~ 101 10'
(l/s)

Figure 5. Effect of CMC content on the storage modulus for 44% vol. Kaolin
suspensions. Data are from Carreau and Lavoie [5].
833

150" L ' .... J ' 50

G'
4O

100"

(.9
3O ~'

50
20

0-I- , " ,' , I0


1 0 -~ 1 0 "2 1 0 "1 100 101
Frequency (Hz)

Figure 6. The dependence of the storage and loss moduli and phase angle on
the frequency for a clay-based coating color containing 1.25 part of CMC per
100 parts of clay. Data are from Fadat et al. [10].
Figure 6 shows the linear viscoelastic data of Fadat et al. [10] for a clay-based
coating color with a different grade of CMC. Here the behavior of the storage
modulus is more like that of a viscoelastic liquid. The reduction of the phase
angle 5 (-tan -~G"/G') with increasing frequencies indicates that the suspensions
becomes more solid-like at higher frequencies. These behaviors are different from
those exhibited by the coating suspensions studied by Carreau and Lavoie [5] and
Ghosh [8].
Transient shear-flow experiments have been used to elucidate the characteristic
times of structure formation or breakup in coating suspensions. A common
industrial practice is to generate a hysterisis loop from shear stress measurements
completed over a strain-rate sweep for a finite time interval. Although this type of
measurement is valuable for quality control, it is of little use in characterizing the
structure formation or breakup in the coating suspension. More useful transient
experiments are the stress growth measurement upon the inception of steady
shear flow, as described by Ghosh et al. [7] and Ghosh [8], and the stress
relaxation measurement after a sudden sheafing displacement, as reported by
Young and Fu [11] and Ghosh [8]. These experiments were able to yield
estimated characteristic times of the coating suspensions studied.
834

Normal stress measurements at very high shear rates were conducted by Windle
ad Beazley [12] using a jet-thrust technique for starch-based coatings. Calculation
of the first normal stress coefficient from their normal stress data gives very low
values, indicating that normal stress effects may not be significant at the high
shear rates encountered in coating operation. This is corroborated by
Triantafillopoulos and Grankvist [4], which did not observe any jet swell in their
high-shear capillary flow experiments for several coating formulations.
Due to the converging-diverging geometry in the coating operation, the flow is
both shearing and extensional. Elongational viscosity may therefore play a
significant role. Measurements of elongational viscosity at the high deformation
rates experienced in a coating operation have not yet been reported. However,
Carreau and Lavoie [5] claimed that they did not observe any significant elastic
or extensional effect in the results from their lubrication equipment at high
deformation rates.

3. STRUCTURAL MODELS TO PREDICT RHEOLOGY


The application of paper coating has a unique aspect in that the clearance
between the blade and the web is only one order of magnitude larger than the
pigments in the suspension. In some places, on the top of a "high" spot of the
paper web, the coating layer is the same order of magnitude as the pigment. A
number of coating defects and the "stalagmites" or "whiskers" described by the
industry seem to be related to the particulate nature of the coating. Therefore, the
process calls for a good understanding of the flow properties of the coating at the
particulate level.
Calculations using Stokesian dynamics were introduced by Brady and Bossis
[13, 14] to relate the microscopic behavior of suspension to their macroscopic
behavior such as rheology. Stokesian dynamics is a technique to calculate the
trajectories of particles in a suspension undergoing flow. The technique has been
shown to predict various phenomena. These include the increase of viscosity with
increasing solids described by Durlofsky et al. [15] and Toivakka and Ekltmd
[16], the reduction of the viscosity of a mixture of large and small spheres given
by Chang and Powell [17] and Toivakka and Eklund [18], the increase of
viscosity at high shear rates shown by Boersma et al. [19], the Brownian shear-
thinning nature of suspensions characterized by Phung et al. [20], and the
influence of particle roughness reported by Bousfield [21 ].
The technique calculates the hydrodynamic force on every particle at a
particular instant of time. Other forces such as electrostatic and colloidal forces
can be included in the calculation. Once the net force on a particle is known, its
acceleration is calculated. The particle velocities and positions are updated with
835

numerical techniques. In the limit of small particle Reynolds number, the particle
velocities can be found through matrix techniques; the velocities must cause the
net force on every particle to be zero.
In a blade coating geometry, the clearance between the web and the blade can
be in the order of 15 ~xn. Bousfield [22] showed that structures of particles will
form in shear fields and that these structures can span the gap between the web
and the blade. This mechanism is the same as those reported by Brady and Bossis
[14] and by Boersma et al. [19]. When a cluster of particles forms in this
geometry, large forces are transmitted between the web and the blade. This
mechanism could be responsible for blade wear and blade deposits, if these
structures exit the blade. Figure 7 depicts this situation: particles are forced closer
together as the fast moving particles must past the slow moving particles near the
blade. If the electrostatic or steric repulsive forces are not significant to keep the
particles separated, particle clusters can span the gap between the blade and the
web.
A potential mechanism for the viscoelastic response of a highly dispersed
suspension was modeled by Toivakka et al. [23] with a Stokesian dynamics
calculation. When electrostatic or steric forces keep the suspended particles well
separated, an energy minimum is obtained by an ordered packing of the particles.
Any small flow field will disrupt the particle configuration and will push the
particles closer together in some regions, as demonstrated in Figure 8. If the
motion is slight, as in a small-amplitude oscillatory experiment, the suspension

Figure 7. Schematic of the results from a particle motion model showing the
clustering of pigments between the blade surface and the web at high shear
rates. This structure can transmit large forces between the web and the blade.
836

Static i| -i =v

I II

Figure 8. Slight motion caused by shear will bring some particles closer
together. This compression of the electrostatic or steric repulsive forces can
store energy and lead to a viscoelastic response of the suspension, even when
dissolved polymers are not present.

Figure 9. A schematic showing the pushing away of a larger particle from a


wall by smaller particles, as described by Toivakka and Eklund [18].

can store energy by the compression of the electrostatic or steric forces. This
storage of energy leads to the possible elastic nature of a suspension, even when
dissolved polymers are not present in the system. Example of this type of
experimental system was reported by Fadat et al. [10] for coating suspensions. If
the deformation is too large, particles move away from the static position and a
non-linear viscoelastic result is obtained.
The position of different sized pigments and particles in the coating layer has
been predicted by Stokesian dynamics calculations. Chang and Powell [17] show
how small particles can disrupt the formation of clusters and, thereby, reduce the
837

shear viscosity. Toivakka and Eklund [18] confirm this mechanism and show how
small particles can end up closer to the blade surface during shear. Fine particles
are able to be closer to the solid boundaries because of their size. In addition,
when particles are exposed to a flow field, larger particles must move faster than
the finer particles near the walls. This velocity difference causes the larger
particles to push the smaller particles even closer to the walls. After a number of
interactions, the larger particles migrate away from the walls. This would create a
separation of particles according to size, with the finer particles near the solid
boundaries. The coating layer could have gradients of particles from its surface to
the bulk. This mechanism is illustrated in Figure 9. Bousfield et al. [24] show that
the free surface after the blade may cause a similar separation of particles based
on size, with the smaller particles being located at the top surface. These results
may explain the high concentration of latex binder at the top layer of the coating
observed industrially.

4. M O D E L I N G OF BLADE COATING
Initial attempts to model the blade coating operation are based on Newtonian
fluids. Follette and Fowells [25], Bliesner [26], Turai [27], and others give simple
analyses to relate shear rates, viscosity, geometry, and web velocity to the force
generated on a blade during coating. Lubrication theory is used to reduce the
complexity of the equations. The pressure distribution under the blade is found to
be a strong function of the blade operating angle. Bliesner [26] and Saita and
Scriven [28] used lubrication theory to relate the details of the blade geometry or
blade deflection to the coat weight and blade loading.
The force on the blade required to reject coating due to inertial forces was
given by Eklund and Kahila [29]. This force can be obtained from a mechanical
energy balance on the flow upstream of the blade. The resulting force F on the
blade, for a roll applicator system, is estimated to be

mV(1 + eosa)
Y= (1)
sin a

where m is the incoming mass flow rate of the coating to the blade, Vis the web
velocity, and cx is the operating angle of the blade. This force seems to be
important for roll applicator systems, but it is not clear how to use this type of
equation for a short-dwell coater. This expression indicates that the blade loading
is not a function of the coat weight metered by the blade nor the viscosity level of
the coating. Therefore, some other viscous forces must be important, in addition
to this force, in the coating operation.
838

Pranckh and Scriven [30] used fimte element method to solve the two
dimensional Navier-Stokes equation for the flow near the blade. The solution is in
conjunction with a nonlinear beam equation that describes the deflection of the
blade and a spring model that describes the deformation of the rubber backing roll
and the web. Surface tension forces upstream from the blade and after the blade
were included in the analysis. In the flow field, there is a stagnation line that
separates the coating that is retm-ned and the coating that is metered onto the
web. The pressure distribution near the heal of the blade and under the blade tip
shows little pressure gradients in the vertical direction: this result indicates that
lubrication theory is valid in this region. Again, the operating angle of the blade is
found to be a critical issue in these flows. The pressure distribution shows a
stagnation pressure upstream of the blade, which is close to 89 p V 2 , and a
pressure increase near the blade heel due to viscous forces. This result links the
pressure pulse due to the inertial terms with that of lubrication theory.
The influence of the absorption of the coating into the porous web is first
accounted for by Chen and Scriven [31]. The pressure distribution of a roll
applicator and by a blade metering element from Pranckh and Scriven [30] is
used to calculate the penetration depth of the coating. The influence of air being
trapped into the web and the compression of the web are taken into account.
Assuming a specific pressure distribution under the blade, Letzelter and Eklund
[32] predict the filter cake thickness and the amount of dewatering. Bousfield [3]
proposed a model to describe the formation of a filter cake on the web during
blade coating. Lubrication theory was found to duplicate the pressure distribution
of Pranckh and Scriven [30] and was used to describe the dewatering and
subsequent filter cake buildup on the paper web. The model by Bousfield [3] is
shown to predict the onset of operational difficulties of literature data, if
dewatering parameters are known. These difficulties seem to be related to the
growth of a filter cake which had a thickness near the blade-paper gap. In Figure
2, the data of Taylor [2] are compared with the predictions of the model. This
show that high solids content increases the coating viscosity and reduces the
amount of water needed to be removed to generate a significant filter cake.
The influence of the shear thinning nature of coating was first modeled by
Modrak [33]. A power-law model was used in conjunction with lubrication
theory to describe the flow in a "rounded" blade. Viscosity data from a capillary
viscometer was fitted for a Newtonian-like fluid, a shear-thinning coating, and a
shear-thickening fluid. The power-law exponent varied from 0.675 to 1.3. The
stagnation point moved downstream under the blade tip as the coating became
shear thickening. The pressure distribution and the hydrodynamic litt on the blade
were an order of magnitude higher for the shear thickening fluid, even though at a
839

shear rate of 3 104 S-1, the shear viscosities were similar. However, the shear
rate under the blade tip must reach 106 s-1. For the three coatings, the shear-
thinning model did a reasonable job at predicting pilot scale data.
Finite element methods were used by Triantafillopoulos et al. [34], Roper and
Attal [9], and Isaksson and Rigdahl [35] to calculate the flow field in a short
dwell coater pond and near the blade tip for a roll-applicator system. Free
surfaces and inertial terms were included in the later analyses. The coating was
described as a power-law model with various values of the power-law index. The
circulation within the pond of the short dwell coater was calculated to change
from a single vortex at high viscosity values to two vortices at low viscosity
values. The shear thinning nature of the fluid was fotmd to reduce the pressure
distribution under the blade and decrease the total blade loading. In addition,
shear thinning fluids are found to produce a "plug flow" nature to the flow field,
where most of the shear can occur in a thin layer near the moving web. The
influence of blade angle and "slip" at the blade-coating interface are also
discussed in Isaksson and Rigdahl [35].
The importance of viscoelastic properties of coating was first brought up by
Windle and Beazley [12]. The discussion on viscoelastic properties was focussed
on the extra normal forces generated during shear: these normal forces could
contribute to the hydrodynamic lift forces during operation. An expression for the
extra force on the blade due to normal forces was proposed.
Two-dimensional solutions for a viscoelastic fluid are described in Sullivan et
al. [36], Olsson [37], and Olsson and Isaksson [38]. The results by Olsson use a
geometry that closely resembles the blade geometry of paper coating. An upper
convected Maxwell fluid, an Oldroyd-B fluid, and a Giesekus fluid are used as
constitutive equations. The pressure distributions and streamlines are significantly
influenced by the presence of viscoelasticity. A recirculation region is predicted
upstream from the blade tip and near the exit of the blade, as depicted in Figure
10. The pressure is predicted to decrease due to the extensional components of
the flow field. This result would produce a lower blade force to obtain the same
coat weight, as compared to a Newtonian fluid. The entrance region seems to be
"clogged" with a vortex, resulting in less fluid that is able to pass under the blade.
This result resembles the outcome of Sullivan et al. [36] where the force on the
blade is reduced because of the influence of viscoelasticity.
The K-BKZ constitutive equation was used by Mitsoulis and Triantafillopoulos
[39] to describe the flow under a blade with the influence of a free surface. The
K-BKZ constitutive equation has multiple relaxation times and predicts normal
forces. The pressure distributions and the net flow rate under the blade are
influenced by the viscoelastic nature of the constitutive equation.
840

Figure 10. The formation of recirculation regions for an Oldroyd-B fluid, as


described by Olsson and Isaksson [38]. One vortex forms upstream from the
heel of the blade. Another vortex forms near the tip of the blade. Both act to
constrict the even flow of fluid under the blade.
The importance and our current understanding of viscoelasticity in coating
flows are reviewed in detail by Triantafillopolous [40]. The influence of various
additives on the viscoelastic nature of the coatings are discussed. The importance
of the viscoelastic nature of these coatings on the operation of blade coaters and
the formation of defects are summarized. A clear direct link between the
viscoelastic nature of the coatings and specific operational difficulties is not yet
established. This direct link is difficult to establish because of the experimental
difficulties that exist in terms of measuring blade forces and comparing one
coating against another for identical operating conditions: changes in coating
composition to change he viscoelastic nature of the coating also changes the
water holding behavior.

5. CONCLUSIONS
With the recem advances in theological testing methods and computational
tools, our understanding of the blade coating of paper has improved considerably
in the last ten years. The importance of "water retention" on the ability of a
coating to operate at specific conditions has been one significant step. However,
paper coating still offers a number of challenges, as high speeds and high solids
are desired. Still the small scale between the blade and the web requires a better
understanding of the small-scale microstructures that can form under the blade.
The role of viscoelasticity is still open to debate. A direct comparison of a
complex fluid dynamics calculation and pilot scale results has yet to be
841

accomplished. These issues should become clear as effort is spent to understand


the role of rheology in the blade coating process.

REFERENCES
1. R. Van Gilder, D.I. Lee, and R. Purfeerst, TAPPI J., 66 (1983) 49.
2. J.A. Taylor, Proceedings of the 1987 TAPPI Coating Conference, TAPPI
Press, Atlanta GA (1987) 1.
3. D.W. Bousfield, TAPPI J., 77 (1994) 161.
4. N. Triantafillopoulos and T. Grankvist, Proceedings of the 1992 TAPPI
Coating Conference, TAPPI Press, Atlanta GA (1992) 23.
5. P.J. Carreau and P.-A. Lavoie, Proceedings of the 1993 TAPPI Advanced
Coating Fundamentals Symposium, TAPPI Press, Atlanta GA (1993) 1.
. S. Purkayastha and M.E. Oja, Proceedings of the 1993 TAPPI Advanced
Coating Fundamentals Symposium, TAPPI Press, Atlanta GA (1993) 31.
7. T. Ghosh, P.J. Carreau, and P.-A. Lavoie, Proceedings of the 1996 TAPPI
Coating Conference, TAPPI Press, Atlanta GA (1996) 303.
8. T. Ghosh, Proceedings of the 1997 TAPPI Advanced Coating Fundamentals
Symposium, TAPPI Press, Atlanta GA (1997) 43.
9. J.A. Roper III and J.F. Attal, Proceedings of the1993 TAPPI Coating
Conference, TAPPI Press, Atlanta GA (1993) 107.
10. G. Fadat, G. Engstrom, and M. Rigdahl, Rheol. Acta, 27 (1988) 289.
11. T.S. Young and E. Fu, Proceedings of the 1991 TAPPI Coating Conference,
TAPPI Press, Atlanta GA (1991 ) 61.
12. W. Windle and K.M. Beazley, TAPPI J., 51 (1968) 340.
13. J.F. Brady and G. Bossis, J. Fluid Mech., 180 (1985) 105.
14. J.F. Brady and G. Bossis, Annual Rev. of Fluid Mech., 20 (1987) 111.
15. J.L. Durlofsky and J.F. Brady, J. of Fluid Mech., 200 (1989) 39.
16. M. Toivakka and D. Eklund, Nordic Pulp and Paper Res. J., 9 (1994) 143.
17. C. Chang and R.L. Powell, J. of Rheol., 38 (1994) 165.
18. M. Toivakka and D. Eklund, TAPPI J., 79 (1996) 211.
19. W.H. Boersma, J. Laven, and H.N. Stein, J. Rheol., 39 (1995) 841.
20. T.N. Phung, J.F. Brady, and G. Bossis, J. Fluid Mech. 313 (1996) 181.
842

21. D.W. Bousfield, Nordic Pulp and Paper Res. J., 8 (1993) 176.
22. D.W. Bousfield, Proceedings of the 1990 TAPPI Coating Conference,
TAPPI Press, Atlanta, GA. (1990) 325.
23. M. Toivakka, D. Eklund, and D.W. Bousfield, J. Non-Newt. Fluid Mech.,
56 (1995) 49.
24. D.W. Bousfield, P. Isaksson, and M. Rigdahl, J. Pulp and Paper Sci., 23
(1997) J293.
25. W.J. Follette and R.W. Fowells, TAPPI J., 43 (1960) 953.
26. W.C. Bliesner, TAPPI J., 54 (1971) 1673.
27. L.L. Turai, TAPPI J., 54 (1971) 1315.
28. F.A. Saita and L.E. Scriven, Proceedings of the 1988 TAPPI Coating
Conference, TAPPI Press, Atlanta, GA (1988) 13.
29. D. Eklund and S.J. Kahila, Wochbl. Papierfabr., 106 (1978) 661.
30. F.R. Pranckh and L.E. Scriven, TAPPI J., 73 (1990) 163.
31. K.S.A. Chen and L.E. Scriven, TAPPI J., 73 (1990) 151.
32. P. Letzelter and D. Eklund, TAPPI J., 76 (1993) 63
33. J.P. Modrak, TAPPI J., 56 (1973) 70.
34. N. Triantafillopoulos, G.R. Rudemiller, and T. Fanfngton, Proceedings of
the 1988 TAPPI Engineering Conference, TAPPI Press, Atlanta, GA (1988)
209.
35. P. Isaksson and M. Rigdahl, Rheol. Acta, 33 (1994) 454.
36. T. Sullivan, S. Middleman, R. Keunings, AIChE J., 33 (1987) 2047.
37. F. Olsson, J. Non-Newt. Fluid Mech., 51 (1994) 309.
38. F. Olsson and P. Isaksson, Nordic Pulp and Paper Res. J., 4 (1995) 234.
39. E. Mitsoulis and N. Triantafillopoulos, Proceedings of the 1997 TAPPI
Advanced Coating Fundamentals Symposium, TAPPI Press, Atlanta GA
(1997), 27.
40. N. Triantafillopoulos, Paper Coating Viscoelasticity, TAPPI Press, Atlanta,
GA (1996).
843

RHEOLOGY OF LONG DISCONTINUOUS FIBER


THERMOPLASTIC COMPOSITES

S.G. Advani* and T.S. Creasy +

*University of Delaware, Department of Mechanical Engineering, Spencer


Laboratory, Newark, DE 19716
+University of Southern California, Centerfor Composite Materials, VIlE 602
MC0241, Los Angeles, CA 90089-0241
1. INTRODUCTION

The total cost of installed composite components is high enough to keep these
materials out of many products that could benefit from them. This introduction
discusses the motivation for using discontinuous fibers to reduce this cost, the
status of research into the manufacturing behavior of these materials and the
scope of this chapter in presenting these properties.

1.1 Motivation
High performance composite materials are beneficial when compared with
other candidate materials by specific strength. But applying advanced composite
materials to mass market industrial products has a tremendous limiting factor: the
traditional high cost of producing parts. The benefits of high strength composites
emerge when controlled fiber placement optimizes the directional strength of the
material. Traditional construction techniques involve significant labor and time
costs. These costs must be reduced to increase the market for designed materials.
Burdensome hand-layup is the most expensive fiber placement method m,.d
must be avoided [ 1]. Tape laying machines, although automated, are too slow for
rapid production cycles [2]. Many industries use injection molding, transfer
molding or sheet molding of low to medium strength composites with particle or
short fiber reinforcement. These techniques rapidly produce items such as
electronic connectors or automotive panels. If high strength composites can be
formed in a manner analogous to sheet metal stamping new markets may benefit
from their use.
844

Thermoforming may allow rapid stamping of high strength composites [3]. In


thermoforming a sheet of fibers impregnated with polymer heats up until the
polymer melts. Two general methods form the sheet: stamping between dies [4]
or applied differential pressure [5], both shown in Figure 1. Initially at rest, the
sheet flows until it contacts the mold. A distribution of strain rates and total strain
occurs over the area of the sheet. Sections of the sheet receive different
displacement conditions as new features of die contact the sheet. Clamps hold the
edges of the sheet; thus, the primary mode of deformation is transient biaxial
elongation. Otherwise, the fonning may occur by other mechanisms common in
continuous fiber systems, [6, 7]. The transients are important as the process time
is short. Steady flow may only be achieved in limited regions if at all.

Figure 1. Sheet forming processes: a) matched die stamping employs precision


tools on either side of the fluid sheet that ensure surface dimensions and quality,
b) differential pressure uses pressurized gas on one side of the sheet to push the
melt onto a tool that controls the quality of one surface.

But strong materials require high volume fractions of either continuous fibers,
which cannot extend in the fiber direction [5], or well aligned arrays of long-
discontinuous fibers [4]. Continuous fiber materials have limitations in
thermofonning [5]. If the fibers remain unbroken--necessary to avoid an area with
a concentration of damaged fibers--the sheet conforms to the mold by interply slip
and by drawing extra fiber length from the edge of the panel toward the center of
the sheet. The sheet accommodates the entire distributed total strain in the
direction of a given fiber by drawing in that fiber. For large deformations a
845

suitably large flange region provides the needed fiber length. This fiber drawing
process produces buckling and wrinkling of the sheet [8]. If severe, buckling can
intrude upon the region of the finished part and decrease the quality of the
component.
Discontinuous fiber systems allow the fibers to move past one another during
forming. Since the fibers can move independently, different regions conform to
their total strain.
Creating a high strength discontinuous composite is challenging. Fibers must
be long enough to provide strength close to that obtained with continuous fibers
[9]. The discontinuous arrays must be well aligned to maintain the necessary high
fiber volume fraction. Fibers conforming to these requirements interact in typical
processes such as injection molding. Fibers will break and alignment will be
limited from fiber entanglement and flow field conditions. Innovative methods are
needed to create a high performance sheet product.
An example long discontinuous fiber thermoplastic melt system (LDFMS) is
LDF/PEKK. Here the matrix is poly-etherketoneketone (PEKK) [ 10, 11]. These 7
micron diameter, 5.6 cm long carbon fibers fall between typical short fiber fillers
and continuous fibers in length and performance. LDFMS material is an attempt
to approach high solid strength typical of highly aligned continuous fiber
composites with melt formability similar to short-fiber filled polymers by allowing
relative motion between adjacent discontinuous fibers. This material contains a
high volume fraction (60 %) of aligned fibers with an average aspect ratio (a r) of
8000, [1 ]. The transient and steady state rheological properties of this fluid are
expected to be different from the unfilled melt.

1.2 Status of the Field


A detailed review of relevant research in particle and fiber filled fluids follows
in Section 2.1.2. In summary, the studies of filled fluids show that fillers increase
the shear and elongation viscosity of the system when compared to the matrix
polymer. Increasing particle aspect ratio magnifies this effect with viscosity
increases of several orders of magnitude. In elongation the fillers--including
particulates--induce a reaction that many investigators believe involves "local"
shear of the fluid between the particles.
However, the specific physics of a system and flow condition must be analyzed
carefully. Similar transient response curve shapes may be caused by unrelated
phenomena.
846

1.3 Scope of the Chapter


This chapter covers the investigation of the elongational flow behavior of two
LDFMS systems. It includes fabrication of a model system from nylon fibers and
a low density polyethylene (PE) matrix with fiber aspect ratios of 25 and 100.
Tensile extension of each LDFMS is compared to a micromechanics model of
relative fiber motion with appropriate constitutive relations considered. The
results are discussed with regard to the measured rheology of the polymers and
implications for formability and future research. Section 2.1 reviews the literature
of the shear and elongational properties of neat and filled polymer melts to show
the generally understood properties of these systems. Section 2.2 discusses the
material preparation steps for making long fiber reinforced melts. In section 3 the
experimental procedures are listed for each device applied to the investigation.
Section 4 contains the results of the experiments. Possible models of extension
behavior of filled systems are presented in section 5 with a check of the data from
section 4. Finally, the conclusions are summarized in section 6 with some
suggested avenues for further study.

2. MATERIALS

Polymer science has grown rapidly since the 1940s as new synthesis methods,
experimental appliances and structural theories were introduced [12]. This section
stmunarizes the outcome of research into the response of polymer melts in shear
and elongation flows.

2.1 Traditional Fluid Systems


It is good practice to study each constituent of a system independently if they
are available as separate compounds. This is certainly the case for systems
containing polymer melts. The polymer melt by itself has an array of properties
that must first be understood.

2.1.1 Neat Polymers


The fluid dynamics of polymeric melts can be described by a set of rheological
properties called material functions [13]. These functions describe the manner in
which the fluid response changes as a function of the flow conditions imposed
upon the melt. The flow conditions can be the applied strain rate (~'yxfor shear,/~
for elongation) or the applied stress (xyx for shear, x E for elongation).
847

Shear Flows
Material functions in shearing flow can be defined under both steady state
conditions and as transient functions with time as an additional parameter. The
transient functions will approach the steady functions for large times trader steady
applied strain rate or stress. We begin with Newtonian material functions under
steady flow conditions. Modified, these functions handle polymeric fluids under
steady and also under transient conditions.

Steady Flow
The fluids of interest in sheet forming are usually non-Newtonian polymer
melts. For these complex fluids, the entanglement of the polymer macromolecules
produces a rate dependence of the steady viscosity [12, 14]. We define the rate
dependent steady viscosity, rls (Tyx), such that
Xyx - rlS(Tyx)Tyx (1)

As shown by Figure 2, isotropic polymer melts behave as Newtonian fluids in


creeping flows, ~'y~ << 1.0 sl; but, they then demonstrate "shear thinning" for
large strain rate. Less common is "shear thickening" at high strain rate. At
decreasing strain rate viscosity approaches a constant value called rl0, the "zero
shear rate viscosity." The viscosity function of the fluid is defined in
nondimensional form as rls (~'yx)/1"10.
The Carreau relation provides a good description of the steady viscosity from
Newtonian creeping flows to high strain rates (power-law region) [15]:
- rl0[1 + O#y )2 ] (2)
The ~ (with units of time) adjusts the position of the "knee" in the TIs--Tyx
curve. Here n controls the power law slope of the curve [13].
Polymer melts also exhibit elasticity which introduces a new stress in shear
flow; this stress acts normal to the direction of shear flow. A cone and plate
rheometer measures the normal force during a shear flow. This normal force is the
total thrust parallel to the rotation axis of the cone and plate. Bird et al. [13]
define the first and second normal stress coefficients in the form:
-2
'lTyy-- '17XX -- l~/rl('il)~,X and "r, z z - X y y - gt2(~)i, 2 . For these coefficients Yyx is
used since a reversal of the direction of the shear flow will not change the sign of
the normal stress. The steady state first normal stress coefficient behavior is
similar to the steady viscosity behavior.
848

Shear
Thickening

Newtonlan
_c
Shear
Thinning
LCP

In~
Figure 2. Steady viscosity of Newtonian and non-Newtonian fluids as a function
of shear strain rate. Newtonian fluids have a constant viscosity. Shear thickening
fluids increase in viscosity as strain rate grows. Shear thinning fluids and liquid
crystal polymers (LCP) have regions of decreasing viscosity for an increase in
shear rate. Some LCPs have two regions of thinning connected by a Newtonian
plateau between them as displayed here.

Transient Flow
The above discussion assumed that steady-state behavior applied to the
measured shear viscosity. Polymer fluids show time dependent properties when
flow starts from rest. The way that transient stress (or equivalently transient
viscosity by rl~ =xy~/'~y~) grows to its steady-state value as the flow progresses
+

delineates the transient material functions of the fluid.


When a Newtonian fluid at rest goes into motion in shear, the shear stress
moves from zero to "Cy~in a step function as in Figure 3. When the flow stops the
stress immediately drops to zero. A viscoelastic fluid set into shear flow from rest
takes finite time to reach steady-state. Stress and rls are now fimctions of Jfyx
and time, x ~ (~(yx,t) and rl~ (J~yx,t) respectively. The plus sign represents a flow
+ +
begun from rest at time "zero" with Xyx or rls increasing into positive time. As
+x ~ Xyx and rls+ --~ rls. We call x +
the flow continues Xy ~ the "stress growth"
[131.
849

Figure 3 demonstrates stress growth for both linear and nonlinear viscoelastic
fluids. For linear viscoelastic fluids the stress grows up to the steady stress level
monotonically. A constant relevant to the flow, usually denoted by ~,l and called
the "relaxation time," is a measure of the time required to reach the steady state
response of the system, e.g., the Maxwell fluid model [16]. The relaxation time
changes the arithmetic equation 1 into a differential equation on stress. Nonlinear
models, such as the White-Metzner fluid, incorporate rate-dependent viscosity
[17]:
+ .+
"l:yx(~'yx , t ) + G "l:Yx(~/yx,t) - rl s (~'yx )~/yx (3)
G is a constant stiffness term that models the elasticity of the fluid. The
solution to equation 3 defines the transient viscosity as
+
rl~(~y~,t) = xy~(5/yx,t)/~y~. Monotonic stress growth function still confines this
model to linear viscoelastic material functions.

--- Newtonlan Fluid


--- Linear Viscoelastic
Non-linear Viscoelastic
. . . . . . . . .

"1 Stress Relaxation


|
~yx

~---- Flow Start-up


, II

o t1
Time
Figure 3. Transient stress in starting-up and stopping of simple shear for
Newtonian, linear viscoelastic and nonlinear viscoelastic fluids. At time t - 0 a
step value of Yyx is applied; at time t - tl, 5[yx - O .

More challenging to model is a material that exhibits a nonlinear viscoelastic


response with stress overshoot. The material functions still carry the same
(~/y~ ,t); but, Xy
notation, e.g., Xy~ +x approaches Xy~ with oscillations. At small
strain rate this nonlinear viscoelastic fluid will act just like a linear one. When
850

+
~(y~ grows large enough it stimulates stress overshoot. As Figure 3 shows, Xyx
grows larger than Xy~ and then falls to reach the steady value. As ~yx increases
the ratio of the peak Xy+x to the steady Xyx increases. The peak value of Xy+x is the
yield stress x y . A set of rl~/TIo fimctions for a nonlinear viscoelastic fluid is
presented in Figure 4. The slowest ~/yx data form an "envelope" curve that is a
maximum TI~ fimetion for the fluid. As ~yx grows, each rl~ curve falls under
this envelope. The steady rls values reached at each strain rate are equivalent to
the steady-state shear thinning property shown in Figure 2.

1.0

~.0.1 ~ 1000~0
+W

o.01 10,ooo~o

0.01

Time
Figure 4. Transient viscosity function for a nonlinear viscoelastic fluid in shear.
This equation includes both shear thinning and stress overshoot with increasing
strain rate. The first strain rate is within the Newtonian plateau since rl~ / rio - 1.
The two highest rates fall within the power-law region of the Carreau curve so
that rl~ / 110 < 1.

The appearance of nonlinear stress growth follows the change in the


conformation of the polymer melt. At rest and without applied stress the long
chain molecules reach a random entangled network, their preferred rest
conformation, through diffusion processes that move the melt to a minimum
energy state [14]. This rest conformation is depicted by the maximum valued
characteristic relaxation parameter ~,l- When exposed to small ~/yx -- small
851

enough to produce linear viscoelastic response -- stress growth and relaxation are
linear with relevant ~,1 9 That is, the strain rate is small enough that the melt can
adjust to the flow by the same diffusion processes that allow attainment and
maintenance of the rest conformation.
Higher ~/yx stimulates a rearrangement of the conformation. The fluid cannot
accommodate the greater strain rate with the diffusion processes. The stress
grows until sufficiently large to change the mechanism by which the molecules
adjust to the stress. The initial rest structure has a potential steady stress
9 + *
comparable to xyx = rl0 ~/yx which is large since rl0>>rls(~/y~). As xyx ~ Xyx
the stress level rises and the diffusion processes cannot adjust to the rapid flow.
The high stress level supplies the energy necessary to break up the network and
align the macromolecules to the flow. As the network tears apart the polymer
becomes less resistant to the applied strain rate. A catastrophic realignment
occurs. Stress falls to Xyx and becomes stable [14].

Elongational Flows
Shearing flows are not the only deformation mode possible in a fluid. Shear
free flows involve extension of the fluid without shearing. Before considering
flows that may contain both extension and shearing components, it is worthwhile
to study the sheerer flows separately when possible. Material functions
definitions for elongational flow follow by analogy to the shear functions. Reports
of steady elongational viscosity data are rare [18]. Many polymers seem never to
attain steady elongation. With this and the transient nature of thermoforming in
mind, we turn to the transient material functions.

Elongation Flow Defined


The volume of fluid remains constant during deformation under the assumption
of incompressibility. The incompressibility effect renders elongation a three
dimensional flow. Uniaxial elongation in the x direction makes the velocity
components of V - v x ~ + Vy ] + v z k equal to:
V x "- ~(,

1.
~'y = -~-ey (4)

Vy-- 1.
--'~.Z
2
852

This produces the strain rate tensor:

0 0
1.
+ (w)r) _ _ 0 --E 0 (5)
2
1.
0 0 --E
2 _
With only terms along the diagonal nonzero, elongational flows are shear-free
flows. We define the shear free elongational flow function [13] as:
+ -- 'l;yy
~xx + - rl~; (6)
where each term with a superscript "'+" is a function of strain rate and time.
This equation defines rl~(/z,t) for measured stress difference under applied
strain rate.
A Newtonian fluid has the single-valued shear viscosity It. It can be shown
that there is a single-valued Newtonian elongational viscosity given by the
Trouton relation
tie = 31.t (7)
This equation also holds for unfilled polymer melts when t~ is small enough
(creeping flow) to keep the response within the linear viscoelastic limit. In this
+
case TIE -- 3 rl~. Building upon the discussion of transient shear behavior above,
one could presume that a shear thinning fluid in elongation shows increasing
deviation from Newtonian behavior as ~ increases. These differences are
discussed next.

Measured Elongation Behavior


Different shear thinning melts may show strain hardening or softening in
elongation [19,20,21]. Figure 5 shows strain hardening for LDPE at 150 ~
[22land for a polystyrene melt at 170 ~ [23]. For the lowest strain rate in each
chart rl~ follows the Newtonian curve up to 1000 and 100 s elapsed time,
respectively. As k rises the non-Newtonian behavior occurs earlier in the flow.
The slowest ~ test forms an envelope curve; but, it is a lower bound envelope for
extensional viscosity.
The shear and elongational functioning of unfilled polymer melts is complex
enough on its own. When particles and fibers enter as well the results become
convoluted and require extra care in interpreting the data.
853

2.1.2 Filled Polymers


Analyzing the flow of filled materials requires giving careful attention to the
relevant physics. With solid constituents come a plethora of plausible
interactions. Impact, rotation, relative motion and non-hydrodynamic (friction)
events may appear in varying degrees. Similar stress strain shapes may have
different effects at their root.
We select the particle aspect ratio as a starting point for classifying filled
systems. For an ideal regular cylinder the cylinder length divided by the diameter,
L / D , reasonably defines aspect ratio. For other shapes, e.g., grotmd minerals,
the major particle dimension divided by the minor length determine L / D .
Generally L / D varies from 1 to infinity (continuous fibers). Aspect ratio
determines the maximum filler volume fraction based on the particle geometry
and orientation. L / D and volume fraction together establish the types of
interactions possible. Tightly packed spheres may occupy up to 74 percent of the
volume; highly aligned cylinder arrays admit 78 and 91 percent for uniform
square and hexagonal packing. Random three-dimensional arrays of low aspect
ratio filler may have few particle/particle interactions at sizable volume fractions.
But large L / D particles easily can reach each other and have significant
interactions in random orientations at small volume fractions. The effect of
increasing particle aspect ratio is discussed next for shear and elongation of filled
melts.

Measured Shear Behavior


Parallel plate and capillary rheometers can evaluate filled systems with particle
or short fiber fillers [24,25, 26, 27, 28, 29, 30, 31, 32, 33]. In Figure 6 glass
fillers in a polyamide substantiate the effect of aspect ratio on shear flow. The
stress obtained with the neat polymer melt in shear contrasts with the stress
ensuing as filler aspect ratio increases at a fixed volume fraction of 30 percent
[34]. Glass spheres raise the steady stress by 3.5 times, also, the time to reach
steady state increases considerably. The fiber/polymer systems each introduce
stress overshoot to the reaction. For aspect ratio 7.3 the stress falls to the same
level obtained with spheres. At L / D 27 however, the stress overshoot is larger
still and the steady stress is 11 times the neat polymer level.
Laun showed that the stress overshoots were the result of the orientation of
fibers initially perpendicular to the plate of the rheometer. When a sample of neat
fluid passes through the overshoot and relaxes completely it can regain its original
structure and again undergo the overshoot. When a sample of the filled material
duplicated the test it did not return to an overshoot. Stress rose only to the level
developed just prior to the relaxation. The physics effective here was fiber
854

rotation, not conformation realignment. Experiments with all fibers prealigned


either parallel or perpendicular to the plate verified this [34, 35]. Barbosa et al.
[36] produced an apparatus for studying fiber interactions and the resulting
transients as short fibers orient to the flow.

lOs
O.lO~/,~o.o3
1.0 / / '~.01
A

!
gO 1S"1
I1. los_

104 , ,,J : ,,J i ''J ' ''d ' ''J - ' ''

10.2 10"1 100 101 102 103 104


Time(s)
a)
10 10
10 9

i~" 10 8
g ,
10

+~10 6

10 5
10 "1 100 101 102 103
Time (s)
b)

Figure 5. Elongational stress growth function at strain rates between 103 and 1.0
s1 for a) PE [22], b) PS [23]. The linear viscoelastic curve in extension forms a
"floor" for the transient viscosity at increasing strain rate.
855

The shortest fibers elevated the steady stress to the same value as the spheres.
These fibers were so short that they aligned to the imposed flow and had the
same impact as the spheres. That is both spheres and fibers with L / D = 7.3
interrupt the streamlines of the flow by nearly the same amount [34]. The longer
fibers could not align completely as they interfered with each other. Their
imperfect alignment kept more of them interfering with the flow. The total stress
level can modify this entanglement effect by providing more energy to align the
fibers.

150 AspectRatio
~il 9 27
~, 100~ ~/ He 7.31
,

so

0
0 50 100 150

Figure 6. Effect of aspect ratio on transient shear response of a polyamide melt


filled with 30 % vol random orientation solids [34]. ~yx =0.1 s1 for each sample.
Particles raise the apparent viscosity from the neat level. Elongated particles
introduce an overshoot in stress in addition to the higher steady stress.

Measured Elongation Behavior


Figure 7 shows the impact of particles and volume fraction on elongation of PS
presented in Figure 5. Both loadings of carbon black generate a large increase in
the extensional viscosity. This increase scales with the particle loading. The trend
of the transient viscosity growth also changed. The neat polymer showed strain
hardening with increasing strain rate. The filled system behaves like a strain
sottening (or shear thinning) material. That is, the elongational viscosity reaches
lower values with increasing strain rate.
Creating short fiber filled systems with controlled fiber orientation employs
injection or transfer molding or even capillary extrusion [37, 38]. Variation in
856

fiber interactions and rate effects modify the quality of the composite [39]. The
elongation of short-glass-fiber filled polypropylene demonstrated the effects of
moderately aligned approximately 150a r fibers on rl~ [40]. The stresses
fluctuated from + 5% to + 15% due to the fiber/fiber interactions from the
distributed fiber alignment within the transfer molded specimens.

10 lo
10 9

~ '10 8

10 7 9 0.(0063
+KI II 0.02
9 0.O63
106 20% Carbon black m o~
10 5 I I IIIIIII I I I llllll I I IIIIIII I I IIIII1
10-1 100 101 102 103
Time (s)
a)

1010
10

"~10 8

10 7 0~0063
+t~ 0.02
0.063
lOS ~ 25% Carbon black l.n_ 0.2
10 5 I I1111111 I IIIIIIII I I!111111 ! III!111
10-1 100 101 102 103
b) Time (s)
Figure 7. Elongation of polystyrene with carbon black particles, aspect ratio = 1,
at 170 ~ at particle volume fractions of a) 20 and b) 25 percent [23]. With a
filler the linear viscoelastic limit curve again forms an envelope under which the
curves at higher strain rate fall. The higher volume fraction shows this effect to a
greater degree.
857

An interesting set of experiments were performed on glass mat/thermoplastic


matrix (GMT) systems at 20 vol. percent fiber [41 ]. Squeezed in biaxial extension
to 1/2 their initial thickness, GMT elongational viscosity showed no correlation to
the zero shear rate viscosity of the various thermoplastic matrices (polycarbonate,
polypropylene and polybutylene terephthalate). Where test temperature controlled
the zero shear rate viscosity of two matrix polymers to match their zero shear rate
viscosity the GMT samples produced differing elongational viscosity! When the
authors considered a shear rate magnification effect on a local level, they
obtained good correlation between the elongational data and the shear viscosity
of each matrix fluid. They tried several magnifications by plotting elongational
viscosity versus matrix shear viscosity. At a strain rate of 1000 times the biaxial
extension rate the data from all GMT materials fell on a common line. All GMT
samples contained the same fiber mat and should have the same shear rate
increase at a fixed elongation rate. They conclude that local shear flow dominates
the bulk flow condition. These results show that from particles to fibers, the
fillers increase extensional viscosity and create a local flow that adds shearing to
the deformation.
In the present work we seek to use a basic micromechanics model [42] to
obtain the relevant shear strain rate induced by the relative motion of the
discontinuous fibers in elongation. From the observation of shear dominated flow
in the elongation of the filled polymer a continuum approach will be used to
model the system as nonlinear viscoelastic fluid subjected to shear.
In order to test this approach we need to produce the LDFMS material. The
next section describes two processes for making highly aligned long
discontinuous fiber composites. The first is a model system that incorporates well
behaved components; the second is a commercial product using a new high
temperature polymer.

2.2 Novel Fluid Systems


Making a highly aligned long fiber system at a substantial fiber volume fraction
with control of fiber placement is a challenge. Early extension experiments with
highly aligned fibers employed capillary rheometers and filler volume fraction of
only 0.0078 % [43]. Production of highly aligned discontinuous fiber arrays is
interesting for both academic study and industrial applications. In academic use
they increase understanding of the basic principles behind the forming process.
Improved strength plus easy processing with properties tailored for specific
applications intrigues industry.
Two materials were used. The commercial product, long-discontinuous-
fiber/poly-ether-ketone-ketone (LDF/PEKK), takes continuous fiber tows and
858

creates a highly aligned system with a distribution of fiber lengths and overlap.
The matrix material is not that well characterized and the fiber interactions with
the matrix are complex. However the material is available in usable form. The
model material can be made with a simple matrix and a selected L / D ratio.
However the price to pay for this control is the labor required to construct the
fiber arrays. For the model system we required a simple polymer with a high
degree of fiber control at 50% volume fraction.
The next sections detail the idea of a novel technique for manufacturing these
highly aligned arrays for academic research and several areas in which the
technique may be applied to produce tailored composite sheets for rapid forming
of panels.

2.2.1 Nylon/PE Model Material System


Rheological models that describe the motion of filled systems usually are ideal
cases easily considered but not often accomplished. In an attempt to control fiber
volume fraction and L / D ratio, we added nylon fibers to PE to make a
composite. The nylon monofilament, which was 0.51 mm diameter, came from
910 rn spools of 134 N test strength fishing line. PE film, 1 mil thick, acted as the
matrix. The details of the process that merged the two components follow.

Control of Fiber Aspect Ratio and Position


A custom built loom allowed fabrication of discontinuous fiber preforms. The
loom bed provided a workspace where preconsolidation with polymer stabilized
the fiber array before the consolidation steps. Fiber mats were 108 by 311 mm.

Consolidation
Two consolidation steps formed the panel. The first pressing combined the
fiber mat with 8 plys of the 0.025 mm thick PE film. A picture frame mold held
the sandwich of bottom release film (Kapton), bottom four film plys, fiber mat,
top four film plys and top release film. A Wabash press pressed the mold and its
contents at 795 kPa, 125 ~ for 45 min. Pressure remained on the mold during the
45 min water cooling cycle that returned it to room temperature.
Wherever the wett lifted some fibers above the others a razor blade cut the
raised fibers. The bridging filaments below each cut held the array together.
Pressing six of the preforms a second time under the same conditions produced
six-ply test panels.
859

Sectioning and Bonding


All L / D 100 tests and some L / D 25 tests at 0.01 S -1 used 22 by 311 mm
specimens cut from the panels with the fibers parallel with the long dimension.
These specimens employed a phase change from solid to melt and back to solid
that allowed the grips to hold the specimen ends.
An attempt at achieving a fully melted specimen used bonded specimens for
the L / D 25 tests at 0.001 s1. These N/PE specimens were 75 mm long with 118
mm nylon extensions bonded to each end with high use temperature epoxy, which
had a 100 ~ cure temperature. The nylon extensions provided a solid gripping
surface outside the fiamace. These coupons would have no solid/liquid transition
zones.

2.2.2 LDF/PEKK System


The commercial material used in this study is a composite of a PEKK matrix
with a high volume fraction loading of discontinuous carbon fibers. This polymer
is a high temperature thermoplastic with a recommended processing temperature
of 370 ~ [11]. The advantages of this system are high strength fibers at high
volume fractions. The manufacturing process keeps the fiber alignment very high.
Final structures made of this material will approach the performance of
continuous fiber composite. In terms of this research the material has the
disadvantage of high processing temperature and a sensitivity to degradation
when exposed to air as a melt. These needs require extra attention during the
experiments.

Fiber Morphology
A breaking strategy converts the continuous fiber tow to discontinuous fibers
and creates a distribution of fiber lengths [44]. A lognormal distribution produced
an average fiber length of 5.07 cm with the longest measured fiber at 16 cm. The
fiber tow is Hercules AS-4 graphite with an average fibril diameter of 7 lam.

Assembly of the Composite


Fibers and matrix combine to produce a preimpregnated tape. Hand layup of
the preimpregnated tape produced panels 42 cm square and 8 plies thick. The
panels were consolidated under pressure at the processing temperature. This step
formed the panels without shearing deformation and allowed some stress
relaxation in the melt. Specimens cut from the panels measured 2.5 by 42 cm with
all fibers aligned along the long dimension.
860

At this point we have two material systems to characterize. The model system
has well known matrix properties and a simple fiber arrangement. The
commercial system contains a new polymer and distributed discontinuous fibers.
In the next section details of the test methods and equipment are discussed. Each
system is characterized by rheometry of both the neat matrix fluid and the filled
composite.

3. TECHNIQUES TO C H A R A C T E R I Z E LDFMS

Rheological measurements of neat polymers by capillary and rotational devices


are covered well in the literature [13, 45]. Many devices for measuring
elongational viscosity exist [18, 21, 41, 46]. None work as well or are as versatile
as shear rheometers. Most designs are optimized specifically to the fluid of
immediate interest. The device described below was developed from the need to
study elongation in long-discontinuous-fiber/poly-ether-ketone-ketone
(LDF/PEKK) at 370 ~ The same device stretched the nylon/polyethylene
(N/PE) so that data from both materials could be compared.

Constant elongation rate tests provide transient tensile stress growth coefficient
+
data, rl~ ( t , ~) = XE ( t , ~)/k. These data describe the difficulty that will be
encountered in forming a useful component at the high elongation rates desired
for mass production. Controlled strain rate forming occurs in matched dies [3].
The fluid sheet becomes trapped at two or more points between the dies and
stretches at a rate proportional to the closure speed of the dies.

3.1 Rheological Measurements


Getting rheological functions from tensile experiments requires analysis of the
specific specimen geometry and the displacement or load control scheme. Each
method is discussed and analyzed in the next sections.

3.1.1 Constant Velocity Flow


Standard test practice for solids involves moving one end of the specimen at a
constant velocity v. This is a reasonable approximation of constant strain rate for
the small displacement of most solid specimens. The execution of the test is
simple to do; electronic controllers always have a linear ramp function available.
Screw drive test machines do this extension mode smoothly. Why is this style of
control insufficient for rheology? We start with the strain rate relation [47]"
ldL v
= - - - - = -- (8)
Ldt L
861

where the specimen length L = L ( t ) = L o + vtsince v is constant.


Continuing:

v{l +
~(t) = -~0
-I -~:o {1 + ~:ot l_l (9)
L0J
with g(0)=~o=V/L 0, which is the coveted constant strain rate. Equation 9
establishes strain rate as a decreasing function under a constant velocity scheme.
An example from linear viscoelastic theory discloses the impact of this difference.
For LDPE the tensile stress growth at 0.001 sl can be obtained from [13, 22]"
"on- i 3rl~ }e(t')dt' (10)

With ~1 = 1000 s and t reaching a maximum of 7000 s for constant velocity


extension, the strain rate decreases by 88%. The drop in tensile stress is slightly
less, 85%, due to the memory effect of the relaxation modulus.

3.1.2 Constant Strain Rate Flow


Obtaining extensional viscosity data from a polymer melt is a complex task
with three major problems for the experimentalist. When the melt contains long
fibers the difficulty increases and requires additional steps in producing values of
elongational viscosity. The next two sections discuss the problems and their
solutions that allowed calculation of apparent viscosity.

Achieving Constant Strain Rate


The first problem comes from the definition of rl~ and the displacement
needed to measure it. We require the transient load data to solve the relation
+(t,~) / ~ - F(t k) / A(t,k) / ~ (11)
where the strain rate ~ is constant throughout the experiment. Equation 8 with
v - L may be integrated to obtain
~t = ln(L / Lo) (12)
Keeping ~ constant requires that the moving end of the specimen follows the
displacement
L ( t , ~ ) - Lo e~t (13)
to account for the increase in gage length [47]. Thus, for constant k, both L
and v are exponentially increasing functions of time. Since the specimen length
increases exponentially a sophisticated control system is needed to accomplish
the experiment. With the growing number of computerized function generators for
driving test machines this is less of a problem than it was 10 or more years ago.
862

The second problem is the conversion of the transient load data F(t,/;) into
tensile stress "c+
E . Engineering stress cannot be used for displacements of interest
to thermoforming. Instead, assuming constant volume of a melt the cross sectional
area decreases with extension as A ( t , g ) - Aoe-~t .
Sectioning the sample after an experiment and weighing it as described below
verifies this assumption. Thus equation 11 above determines the transient stress
and viscosity.
Third, the ends of the melt must be securely held during the test. Several
methods do this with neat fluids and particle or short fiber filled systems [48]. For
some L / D 25 tests detailed here we could use a short specimen with nylon
extensions bonded to its ends with high temperature epoxy. This specimen type
could only be used at the lowest strain rates. Larger L / D or faster testing rates
need another technique.
Adding long fibers to the melt makes each of these problems more challenging.
Long fibers greatly increase the elongational viscosity and the force on the
specimen is larger. In this study a phase change method held these specimens
securely. The extreme ends of the specimen remained solid so that standard
Instron grips could hold them. The specimen, when heated in the center, formed a
melt/fiber system in a gage length of approximately 200 mm within a total
specimen length of 305 mm.
This solution to the third problem affected the first two as well. With a
transition from solid to melt and back there is a distribution of strain rates from
zero to a maximum and back to zero again along the major axis of the specimen.
This distribution broadens with the presence of long fibers. Some fibers at each
end of the melt zone will have some part of their length stuck in the solid or semi-
solid polymer. With a distributed strain rate, the change in cross sectional area is
distributed also and must be determined before performing the conversion of load
to stress and displacement to strain rate. Since no simple assumption of the area
distribution is possible for these systems we determine the resulting strain rates
and cross section from a post-experiment analysis of the specimens. Since there is
a transition in strain rates from one end of the specimen to the other, this viscosity
must be designated as an apparent value [49, 50].

3.1.3 Interrupted Flow


Interrupted flow experiments provide a means of determining the source of
nonlinear response (stress peaks) and the change in relaxation response as flow
progresses from rest to steady-state. A variation of the constant /~ test, the
interrupted flow experiments run at a constant strain rate for a time and then the
863

displacement abruptly is held fixed for a controlled amount of time. Upon


resumption of the constant g flow, the transient response of the fluid depends
upon the complete history of the specimen, that is, prior flow, hold period and
resumption of flow.
The objective of this experiment is to show that the stress growth is similar to
the behavior of a nonlinear viscoelastic fluid. That is, the presence of the peak
stress comes from the fluid dynamics of the melt alone and not from fiber-fiber
interactions or changes in fiber orientation--as for random oriented short fiber
+
systems [34]. If the melt is solely responsible for the peak in XE, then, given
sufficient stress relaxation, the LDFMS will again display stress growth to a
transient peak [51]. The opposite also must apply, i.e., insufficient stress
relaxation will inhibit the peak stress. It is possible to select a single interruption
period that will provide sufficient stress relaxation at one strain rate and
insufficient reduction at another rate.

3.2 Apparatus and Procedure


Petrie [47] reviewed the factors that complicate construction of an apparatus to
achieve constant rate extension of specimens. LDF/PEKK complicates the
problem ft~her by requiting both high processing temperature and long specimen
length. The melted zone must be at least three times the average fiber length ( i.e.,
totaling 16.8 cm) and ideally should be 1.0 to 1.2 times the largest known fiber
length ( i.e., totaling 16.25 to 19.5 cm). This length avoids capturing a significant
population of fibers that would otherwise bridge the melt zone.
An electric tube furnace combined with a computer-controlled Instron model
1321 hydraulic test machine produced the rheological measurements collected
here. This apparatus allowed elongation rates, ~, from 10-5 to 10-3 s"l for
LDF/PEKK and 10-3 to 10"1 s"1 for N/PE . Processing requirements (exposure
limit of the material at 370 ~ restricted the total test time during the lowest
and the Instron hydraulic power supply's maximum flow rate limited the fastest
attainable.
Each test began with the furnace at room temperature. A consolidation device
held the specimens under pressure during heating and testing. Both types of N/PE
specimen used the same frame and insert. Four spring clips held the monolithic
samples in place between the steel insert and frame. A single clip over the center
of the 76 mm bonded coupon kept the system together. Bonded specimens used
nylon shims in the comers of the frame to keep the insert from touching the nylon
extensions.
A high temperature film material (Upilex R) sealed and consolidated the
LDF/PEKK specimens during the experiment. A vacuum drawn on the bag
864

protected the melt from degradation during the long heating times required with
the fitmace used. With the heaters switched on, the specimen temperature rose to
the test temperature. The ends remained well below the Tg of the matrix so that
they provided a finn grip surface. After reaching the test temperature--following
20 min of heating--the sample rests for 5 min before the extension starts.
With the techniques developed here, we evaluated the filled fluid systems
proposed in section 2. In the next section the results of each experiment are
shown. First the neat polymers are characterized, then the filled systems are
extended and the data are discussed.

4. BEHAVIOR OF LONG DISCONTINUOUS FIBER SYSTEMS

This section presents the experimental data for nylon/polyethylene (N/PE) and
long-discontinuous-fiber/poly-ether-ketone-ketone (LDF/PEKK) in series. First,
for each system, results for the neat polymer appear progressing to the filled
system in elongation.

4.1 Nylon/PE Model System


This system contains two well developed materials. The rheological testing of
the LDPE confirms the specific properties of the batch of film used. Extension of
the nylon fiber/PE melt demonstrates the impact of fixed length aligned fibers on
viscosity of the composite.

4.1.1 Rheology of PE Film


The general character of a LDPE melt is well known. The melt is isotropic,
shear thinning and stable in air when close to its melting temperature. The first
task of this research with PE was finding a reasonable melt temperature that does
not degrade the melt during the extensional tests. Differential scanning
calorimetry (DSC) [12] of the PE film shows gradual softening until 109 ~
Melting is complete at 121 ~ with no additional thermal effects found to 200 ~
Temperatures of 125, 140 and 200 ~ were considered for the role of standard
temperature for all PE and N/PE experiments. The shear viscosity at 125 ~ was
too high for easy use of the cone and plate rheometer. At 200 ~ PE degrades
severely within an hour. PE viscosity is stable for 2600 s at 140 ~ with viscosity
averaging 30970 + 650 Pa-s during the test.

Steady Shear Viscosity


The rotational and capillary viscometers discussed above produced the steady
shear viscosity data presented in Figure 8. Also shown are the data of Meissner
865

[52]. The LDPE film corresponds to a Carreau curve with parameters 11o =
43,850 Pa-s, ~ = 5.40 s, n = 0.489.
2x10 5
osl_ :3o'c l

I
104 150~ "~"~"0 PE 140~
n
ffl
~' 103

102
10-4 10-3 10"2 10-1 100 101 102 103
t (s")
Figure 8. Steady shear viscosity of PE at 140 ~ in air as a fimction of strain rate.
Data are from rotational and capillary rheometers. Curve is a Carreau curve with
parameters: rl0 = 43,850; ;~ = 5.40; n = 0.489. Data at 130 and 150 ~ from
Meissner [52].

4.1.2 Rheology of N/PE


Measured rheology of the N/PE system is limited to extensional deformation.
With 60 % vol fibers that are 13 and 50 mm long, N/PE is incompatible with
standard rheometers.

Controlled g Flow
Controlled g flow experiments on two types of N/PE produced the following
set of data. The extensional viscosity shows the impact of aligned fiber filler on
an isotropic melt in extension.

Elongational Viscosity
Nylon fibers with an aspect ratio of 25 raise the elongational viscosity of the
system to 57 times 311o for PE. Figure 9 displays the measured transient
elongational viscosity. For a total strain of 0.15 the average viscosity is 7.47
MPa-s. For strain rates covering one order of magnitude the viscosities are close.
The lowest viscosity was reached by the panel with the curved fibers.
866

Increasing the aspect ratio to 1O0 raises the extensional viscosity at the lowest
strain rate by another order of magnitude to 192 MPa-s. This is three orders of
magnitude above the 31"10 value of 0.132 MPa-s. Figure 10 shows this and it also
shows the drop in viscosity with increasing strain rate. Each rate increase by
about 10 drops the viscosity by a factor of 5 to 6. These changes are similar to
the function of a shear thinning fluid in the power-law region.

2x107

107
n
A n
W
!
m


0 0.0457s-1
I I 0.0257
El 0.00261
106
0.15 0.30
~t
Figure 9. Transient elongational viscosity of L / D 25 N/PE. The prior stress data
are divided by the extension strain rate to obtain viscosity. The data show that the
elongational viscosity is fairly consistent among the three tests. The wavy fiber
plys in the 0.0257 s"1 strain rate test make the viscosity increase less and the peak
occur later. The smooth curves are average curves for three repetitions; only one
set of data points is shown for each strain rate to clarify the chart.

Figure 11 shows "steady" type viscosity data. The error bars at each point
show the range of peak viscosity to the last viscosity attained at the end of the
experiment. Points are connected with straight lines to group them by aspect
ratio. The L / D 25 samples are rather fiat with strain rate--discounting the
sample with wavy fibers. The 100 L / D specimens look like a power-law region
response with decreasing viscosity.

4.2 LDF/PEKK System


Experiments with LDF/PEKK follow the reporting sequence of the model
material above. For this system, interrupted flow data follow the steady extension
867

results. These experiments explore the interesting nonlinear aspects of the


response. We begin with the neat PEKK.
109
0.00437 s "1
108
- -1
A
....... 0.04-~7 s
m 107

106
tll

31"1o PE
+~ 105

104 [I m I m
0 0.15 0.30
~t
Figure 10. Transient elongational viscosity of L / D 100 N/PE at three strain
rates. The slowest strain rate shows an increase of viscosity of approximately 3
orders of magnitude from 3r10 of PE.

109

A
Io
t~
IL
I
108

107
0 100
I-! 25
~176
II
o. 106
m

3rio PE
10 5 m m m m m m mmmm mlmm mmm m

104 , , ,,I , , , ,I . . . . 1
0.001 0.01 0.1
(s)
Figure 11. Extensional viscosity of N/PE for two L / D ratios compared with
neat PE; points are connected with straight lines to group the aspect ratios. The
L / D 25 material, accounting for the wavy fibers in the middle sample, remains
fiat in terms of the viscosity/strain rate ratio. L / D 100 material shows a power-
law response to increasing strain rate.
868

4.2.1 Rheology of PEKK Polymer


High temperature melts may degrade with long exposures to their processing
temperatures. This would not be of concern in a rapid industrial process; but,
slower processes and characterization experiments may be carried out over an
extended time in order to improve the understanding of the material's behavior.
First aging of PEKK is discussed, then the measured viscosity results are shown.
DSC of PEKK illustrates the extra complexity of this matrix thermoplastic.
There is a glass transition temperature at 148 ~ Soon after this the solid
crystallization peak arrives at 201 ~ Further heating brings melting at 322 ~
which is complete at 360 ~ The recommended processing temperature is 370
~ This temperature, with a few noted exceptions, was the thermal condition for
rheometry and extension.
In order to differentiate the effect of aging from other viscosity changes, PEKK
melt was sheared at 1 s~ for 1 h in order to track the change in 11 with time for
two environments: N2 gas and air. The PEKK aged in N2 had an increase in
viscosity of 60 % during the hour with viscosity increasing from 470 to 749 Pa-s.
The air exposure experiment was a severe test of the effect of air on the PEKK
melt. The rotational rheometer was heated with air prior to insertion of the PEKK
disk. After re-heating to 370 ~ the PEKK melt was allowed continued exposure
to air for the 5 min stabilization time plus an additional 15 min. This exposed over
one-half the surface of the 1.5 mm thick melt to air. After this conditioning the
initial viscosity of the melt was 915.5 Pa-s, almost twice the viscosity of PEKK in
an N2 environment. After 2000 s the experiment was halted because the viscosity
rise could lead to damage to the rheometer. Viscosity at 2000 s was 2310 Pa-s.

Steady Shear Viscosity


The shear viscosity of PEKK measured from 2 x 10- 3 to 20 S-1 is in Figure 12.
For strain rates from about 0.1 s-1 and higher, PEKK behaves as a shear thinning
polymer melt. Below this rate PEKK demonstrates aspects of the behavior of a
thermotropic liquid crystal polymer (LCP) [53, 54]. For strain rates less than 0.1
s-1 the viscosity rises from the one-decade-wide plateau and increases with
decreasing shear rate. At the lowest rates, 2 x 10-3 and 2 x 10-2 s-~, the melt did
not attain a steady shear rate within the processing time limit adopted for this
study. The viscosity shown for the slowest rate is the average; the vertical bar
indicates the range from 1/3 of the maximum to the maximum viscosity measured.
The previously reported experiments with polyethylene at strain rates from
4 x 10-4 to 4 x 10-1 s1 verified the performance of the RMS-800 at low sheafing
rate. As ~ decreased the noise level increased; but, the steady behavior is as
869

expected for PE. The slowest shear rate was selected as 1/5th the slowest rate
used with PEKK. Here the PE did not reach steady-state within 3800 s; but, the
viscosity grew towards the steady value without any unusual viscosity change.

10 s Melt
- O - PEKK 3 7 0 ~
10 s PE 140 ~ C
I

0,, 10 4

10 s

10 =
10 4 10 -a 10-= 10-1 10 0 101 10 =

(S "1)
Figure 12. Steady shear viscosity of PEKK at 370 ~ PEKK has power-law
behavior at low shear rates up to 0.2 s -~ and at high shear rates above 2.0 s -]. A
plateau viscosity connects the two power-law regions at about 470 Pa-s. The PE
data shown were collected with the same rheometer to test its performance at low
shear rates.

The steady shear viscosity of PEKK measured from 2 x 10 . 3 to 20 S-1 is shown


in Figure 12.

Transient Shear Viscosity


Figure 13 shows the change in transient viscosity growth as ~ is reduced from
20 to 0.02 s-]. At 0.02 s-] the viscosity grew from the expected plateau value of
about 470 to 600 Pa-s to reach a peak of 5610 Pa-s. After the peak, viscosity fell
to 2260 Pa-s, which is more than four times greater than the steady viscosity at
0.2 s-~. Note that at the highest shear rate some undershoot of ~1~ appears.
Similar viscosity growth has been noted for thermotropic liquid crystalline
polymers [55].
After a pre-shear at 1 s-~, a PEKK sample was allowed to relax for 30 rain.
Then it was sheared at 2 10-3 s-~ for over 3500 s. The viscosity rose to a
maximum of 136,000 Pa-s. Figure 14 shows the experiment from pre-shear
870

through two post-shear runs ('i, = 1.0 sl). The post-shear experiments returned the
measured viscosity to the levels measured in the pre-shear plus an increase
expected from the aging results above. Thus the significant rise in viscosity at
2 x 10 -3 sl cannot be attributed to aging because the increase is reversible and
the exposure time in N2 was limited to the time in which viscosity would be
expected to increase by 60 % at the most.

($-1)
lO 4 9 0.02
? D 1.00
A 20.0
v

+
~. l O =

10 =
0 6 12 18

7
Figure 13. Transient viscosity of PEKK at 370 ~ N2 environment. The tests at 1
and 20 s~ strain rates respond as expected for shear thinning isotropic melts. The
drop of shear rate to 0.02 sl produced a dramatic rise in viscosity.

4.2.2 Rheology of LDF/PEKK


LDF/PEKK is the material system that started this research into long fiber
systems. The results obtained with LDF/PEKK guided the sequence of
experiments toward additional cases of deformation control. With the basic
elongational tests finished, we looked interrupted flow as a means of separating
fiber interaction and rheological components of the tensile response.

Controlled ~ Flow
Following the basic format used with N/PE, the next sections show the
reactions of LDF/PEKK to similar experiments and additional tests.
871

,,

_ ,,, ~ (,-1)
10 s
A 1.0 Pre-Shear
,~ 10 s D 0.002
m
el 10 4 9 1.0 1st Post-Shear
II 1.0 2nd Post-Shear
+t=-I 0 s

10 2

101

10 0 ~ , , I , , , ! . , , , ! , , , ,

0.0 2,5 5.0 7.5 10.0

Figure 14. Transient viscosity of PEKK at 2 x 10 -3 s"l compared with pre- and
post-shear transients of the same specimen, 370 ~ N2 environment. This shows
that the viscosity increase at the low shear rate was a reversible function of the
applied shear strain rate.

Elongational Viscosity of LDF/PEKK


For each elongation rate, Figure 15 shows the normalized rl~-t behavior of the
LDF/PEKK. At the slowest rate, T1E reached its maximum value. This value was
used to normalize the rl~-t curves for all three rates. The -+
"rE - ~t data presented
are remarkably like the response of a nonlinear viscoelastic fluid in shear. This
becomes more apparent in the r l ~ - t curves shown here. The elongational
viscosity is checked against neat PEKK data at the end of the controlled stress
tests which immediately follow.

Interrupted Flow
In these experiments LDF/PEKK extended at a constant strain rate for a time
and then the extension stopped with the total displacement held constant.
Following a fixed rest period the flow resumed and the transient response
recorded. Repeated appearance of the peak stress is a function of the rest time
and applied strain rate. Also, the effect of slight fiber misalignment on the initial
stress growth was noted. Panels are produced by hand layup of prepreg sheets.
During the first extension step the fibers align to the drawing direction and
872

subsequent elongation started with an increased modulus of 41.3 GPa, up from


the initial 14.7 GPa slope.

I
Q.
0.1
~o
x
oo
04 0.01
.3 -1
Ip 0 1.49x 10.4 s
+ UJ I I 1.66 x lOs
9 1.39x 10
0.001 9 I , I . I 9
0 100 200 300 400
Time (s)
+
Figure 15. Normalized elongational viscosity growth, rl~. , for three elongation
rates of LDF/PEKK. The curve for the 10s rate reaches a value of 1 at 994 sec.

The data in Figure 16 show the result for three extensions of a single specimen
at a target ~ of 0.001 s-1. For the first extension the specimen elongated 8% and
the stress relaxed for 165 s. This time allows over 90% of x E to dissipate. After
the 165 s relaxation period, x~ fell 92.8% to 12.4 MPa. Atter restarting the
extension the specimen again passed over a yield peak, although, at 231 MPa, it
is 5.0% lower than the peak in the first flow. Also, in the second flow the sample
did not attain the desired near-fiat quasi-steady plateau after passing x y . Instead,
4-
a near-linear decreasing x~.-e results. The lower x y is probably due to the
incomplete relaxation of x~. atter the first flow. The declining x~. atter the
second peak is due to an error in the assumed gage length of the sample during
the second flow.
The three elongation test was repeated at the next slowest rate, 10-4 s1. The
stress relaxation data above demonstrate that stress decay is much slower after
drawing at this rate. Thus, for a fixed rest period, less stress will dissipate before
the extension resumes and this should significantly influence the size and shape of
the peak during the second and third extension. Figure 17 shows that the peak is
less in the second and third flows.
873

250 i I

Extension
200
i~ ~ First
13.
v
150 '~ Second
Third

I.U
100

50

0 ii . . I,
0.00 0.05 0.10 0.15 0.20 0.25

Strain
+
Figure 16. 1; E for three step elongation of LDF/PEKK with g of 10"3. Each stress
relaxation period is 165 s. Peak stress returns after the second and third restart of
extension.After the second extension, x E decays from 54.1 to 5.38 MPa, a 90 %
+
decrease in 165 s. The third elongation step produces another peak in x E . By this
time the error in the apparatus is significant throughout the elongation. The
significance of the third extension is that a yield stress occurs again.

Since this is a new material with unique properties, two questions about the
+
test results come to mind: (i) What is the effect of the fiber alignment on the x E -
e behavior? (ii) To what extent might bridging fibers contribute to the peak stress
or the magnitude of the stress growth slope?
The effects of fiber alignment and fibers that bridge the gage length can be
studied with the multiple extension test data by shifting the second pull curve so
that it starts at zero relative e. After the first extension the stress growth slope
grows from 14.7 GPa in the first extension to 41.3 GPa in the second. This shows
that the minor misalignment that results from hand layup of the plies is significant
in its effect upon the initial slope of the stress growth. No further increase in
slope occurred in the third extension.
On the prospect of bridging fibers adding significantly to the slope of the stress
growth or the value of x y , the population of bridging fibers is much less than 2.6
% of the fibers in the cross section based upon the distribution data of Chang and
Pratte [1 ]. The work of Bums [44] shows that at most 0.5 % of the fibers have a
874

statistical chance of avoiding fracture and bridging a 30.5 cm melt zone. So the
bridging fibers alone could account for at the most 1/22 of the slope during the
first pull based upon the modulus of AS-4 carbon fiber. Then, during the first pull,
any bridging fibers break. The second and third pull stiffness and x~ are
therefore entirely due to the properties of the fluid and the shear strain rate
magnification within the material.

200 Extension
- 0 - First
150 A
! , ~ Second
Third
13.
v
IO0 I
,, ,

+~
J .

50 f ' ' 6 "

O' . . . . 9 i i

0.00 0.05 0.10 0.15 0.20 0.25


Strain

Figure 17. XE for three step elongation of LDF/PEKK with ~ of 10-4 The stress
"~ ~

relaxation period is the same as for the higher strain rate test. Less stress relaxes
and the peak stress is diminished.

The multiple flow test at 10-4 s-I ~, when combined with the prior discussion,
supports the view that x y is a nonlinear viscoelastic effect under constant ~.
Since x E had only decayed by 68.1% atter the 165 s pause the melt could not
move to its preferred rest conformation before the next pull started. Thus, x y
diminished.

4.3 Summary of LDFMS Results


The data presented above show that LDFMS have the following properties in
extension:
Extensional viscosity is orders of magnitude greater than the zero shear
rate viscosity of the matrix.
875

The increase in extensional viscosity generally increases as the aspect ratio


of the aligned fibers increases.

At L / D 25 N/PE had a "flat" extensional viscosity and at L / D 100


N/PE had a power law extensional viscosity.

At an average L / D of 8000 LDF/PEKK had a power law extensional


viscosity.

At the highest two strain rates, LDF/PEKK in extension demonstrated a


peak stress that could be recovered with sufficient relaxation of stress and
that was diminished with insufficient stress relaxation.
The next section proposes an analysis of the data that attempts to explain these
properties by relating the shear and extensional flows through applied
micromechanics.

5. MODELS OF R H E O L O G I C A L PERFORMANCE

Fibers placed in a matrix can dramatically shift the deformation mode of the
system. For the filled fluids discussed here the total data set of flow, relaxation
with cessation of flow and behavior with restart of flow must be considered.
Figure 18 shows micromechanics and constitutive domains for the aligned fiber
systems of interest. The left figure shows a fiber/polymer cell in scale for a 60 vol
~ fiber loading. The cell is globally deformed in extension. A fluid element from
the cell is displayed on the right to illustrate the applied global elongation and the
generated local shear deformation defined by the development that follows.
Micromechanics provide the local deformation. The constitutive relation must be
appropriate to the superimposed deformations of the fluid element.

5.1 Micromechanics: Shear Cell Analysis


The fiber/polymer cell in Figure 18 is the "shear cell" used in solid and fluid
mechanics models. Next we develop the basic properties of the regular cell.

The first analysis of the effect of long fibers on extension of Newtonian fluids
applied to dilute suspensions [42]. The shear strain rate is concentrated in the
near field of the fiber. This effect increases the extensional viscosity of the
system, which was verified by Mewis and Metzner [56]. A derivation for shear
thinning fluids finds that the effect of the localized shear rate removes the fibers
876

as a factor in extensional viscosity [57]. That is, the viscosity drop in the fiber
region compensates for the presence of the fibers. However, the systems
employed here are not dilute. There is a substantial difference in shear rate
distribution in a concentrated suspension. The shear rate is significant and of the
same magnitude along the entire cell radius Ri < r < R 0.

Figure 18. The combination of micromechanics and constitutive relations provide


the model for filled polymers. Micromechanics relate the strain rates and portions
of the stress obtained from each mode. Constitutive relations relate the stress to
the strain rate of each flow mode.

The relevant equations for the micromechanics start with the fiber pulling
speed. Figure 19 illustrates the relative motion of the top fiber with respect to the
lower neighbor fibers. Each fiber remains whole and therefore moves with the
velocity of its centroid within the global elongation flow [58]. The relative
velocity for ideally spaced fibers is VreI - ( x 2 - x l) - L / 2 . The Navier-Stokes
equations yield the x component of the velocity vector:
#-L ln(r / R~ (14)
vx (r) - 2 In k
where
vx (R i ) is the inner boundary condition vx = ~ / 2

vx (R 0) is the outer boundary condition v x - 0


877

is the strain rate in s -~

L is the fiber length

is the radial coordinate such that R i < r < R o

& is the cell radius, Ro > R i = fiber radius

k is the ratio fiber to cell radii Ri / Ro , this is equivalent to


where f is the fiber volume fraction
Then the shear rate is
~Vx_ kL 1
(15)
r ~rr - 21nk r

Figure 19. Relative fiber motion forms a shear cell in the Batchelor model.

The magnification scales directly with L / D.


Continuing with the micromechanics, we compare the portions of the tensile
stress each deformation contributes. This shows the effect of aligned fibers on the
elongational viscosity for a Newtonian fluid. Pipes took a similar approach [59]
but immediately eliminated the extensional load of the polymer. Here both effects
are included. The total stress to extend the system is x E - ( F m + F f ) / 2 A; this is
the sum of the force needed to extend the melt annulus (Fro) and the force
required to pull one fiber from the annulus ( F f ) . The forces are divided by the
878

area of two shear cells (2A) as only half the fibers carry the tensile load with a
good choice of the free body diagram [60]. For any volume fraction
Fm = 2AfqE~. The fiber pulling force comes from the shear stress over the fiber
surface within the cell [58] or F f = x , . x ( 2 ~ i ) L / 2 = rl'~rx/~.L. Using equation 15
the tensile stress is:

XE "-
t,}
3(1- f)rl~ + Ink tie -~ (16)

with tiE = 311. The first term shows the elongation component and the second
term shows that the shear cell effect has a factor of (L / D)2. The portion of the
total stress due to the sheafing within the cell quickly becomes orders of
magnitude larger than the portion provided by elongation of the annulus. For
L / D greater than 5 the annulus stress is less than one percent of the total
measured stress.

5.2 Constitutive Relation: Giesekus Fluid Model


Many constitutive relations exist for polymer melts. The objective here was
finding a suitable relation for the combined flow of the filled system. If a relation
does a reasonable job with both shear and elongational flows of neat systems it
would be useful with a variable L / D as one or the other mode becomes
dominant. The relation must incorporate the effects of conformation changes that
accompany the restructuring of the polymer network under flow. The Giesekus
model [61] provides realistic behavior for nonlinear viscoelastic fluids; it is based
upon a network breakdown of interacting polymer molecules. The form of the
model used is from Bird et al. [13]"

+ ~1~(1) - - 0 ~ { ~ 9 ~} --O~ 2 {~(1) " 17 + 1; " T(1)} ----


110 (17)
+ - ~

where ~ is the stress tensor, ~ is the strain tensor, 11o is the zero shear rate
viscosity, ~l is the relaxation parameter, ~,2 ~ 1 / 1000 is the retardation
parameter, and tx is the parameter that determines the degree of nonlinear
behavior. This set of equations was solved with a 4th order Runge-Kutta
program.
879

The strain rate tensor acts as the driving element of the solution to equation 17.
These tensors arise from the micromechanics of the shear cell. The strain rate
tensor for the elongational flow of the polymer annulus comes from the velocity
vector: Y = Vr~.r + vo~o + Vx~x with v x = kx , v o = 0 and the boundary conditions
on radial velocity of v~(r= R i ) = 0 and v~(r= Ro)< 0. These BCs allow the
polymer to flow toward the center to correspond to the extension in X with the
fiber acting as a solid inner boundary. To satisfy continuity

V ' V - 1 r ---5__ ~(R2


- ,--7-
r/9 (rVr)+ k = 0. And ultimately V - 2zk. - r ) ~', + kXex. The strain

rate tensor is:


- _R.2
0 0

R?
- Vv + {Vv} r 0 e(-~-l) 0 (18)
/

0 0 2~

The sum of the diagonal terms is zero, which satisfies continuity.


In shear flow the only velocity term is as defined above in equation 14. Thus:
0 0 ~L
2rlnk
= Vv + {Vv} r _ 0 0 0 (19)
~L 0 0
2rlnk
These tensors show that the magnitude of the strain rates are bounded by 2e
in elongation and ( 1 / l n k ) ( L / D ) ~ in shear. For L~ D over 5 we rewrite
equation 16 as
XE f _fL] 2

5.3 Application to the Systems


Next the micromechanics and constitutive relations are applied to each
material. The N/PE and LDF/PEKK use the micromechanics relations to verify
the steady state data and the constitutive equation is applied to the LDF/PEKK
system.
880

5.3.1 N/PE Analysis


In the section of results we presented the apparent extensional viscosity of the
N/PE in Figure 11. This chart placed the values of viscosity with respect to the
average tensile strain rate. With the micromechanics discussed just above, we
now determine both the shear strain rate and the equivalent matrix shear viscosity
by equations 15 and 20. Figure 20 displays the resulting matrix viscosity from the
micromechanics model compared with the neat shear viscosity of the PE film.
The correlation is good overall with the larger L / D showing the best following
of the PE curve with the power law behavior of the neat fluid followed by the
N/PE result. The viscosity calculated is lower than the neat PE however. Two
possible causes for this are insufficient consolidation force and less applicability
of the model as L / D approaches 1.

5. 3.2 LDF/PEKK Analysis


The LDF/PEKK system contains fibers of the same length as the L / D 100
N/PE but the aspect ratio is 8000. We find that the viscosity ratio TIE/TI is
expected to range from 2.2 to 7.4 xl 0-". This is a reasonable range of values for
the two high strain rate experiments; but, it is up to 2 orders of magnitude too low
for the slower tests.
An alternative viewpoint is provided by rearranging equation 20 and predicting
the neat melt viscosity at the shear rate obtained with equation 15 for each
experiment. For the two highest strain rates the estimated viscosity from
LDF/PEKK covers the possible range of neat viscosity from PEKK data. As the
strain rate decreases the over-estimate rises. At the slowest rate the viscosity
estimate is about 68 times larger than the PEKK viscosity, TII,EXr, in the LCP
region.
At this point we can present two possible scenarios for this deviation at low
shear rates. The sliding plate work of Ericsson et al. [62] suggests the first. Their
work was with L / D 1000 and 2000 glass fibers in polypropylene (PP)
thermoplastic at 14 and 28 % vol fraction. They found that at high shear rates the
increased viscosity scaled with the square of the fiber concentration which
indicates the response is dominated by hydrodynamic effects (shear of the fluid in
narrow fiber regions at rates much greater than the bulk). At low shear rates they
argue that non-hydrodynamic, i.e., fiber to fiber, interactions dominate the stress.
They make no physical argtmaent for the change in behavior with strain rate; but,
they note the increase in viscosity scaling at lower rates. The behavior of neat
PEKK suggests the second possibility.
881

105

A
104
0
u) Source
103
- PE
O N/PE L/D 100
I"! N/PE L/D 25 r
102
10 .3 10 .2 10 "1 1 10 102 103

(s "1)
Figure 20. Matrix viscosity of N/PE derived from the extensional viscosity data
for the system. Solid line is the steady viscosity of the PE film.

This system could be responding to the thermotropic liquid crystal polymer


(TLCP) character of the PEKK. We know the matrix material starts with TLCP
structure and moves to isotropic as shear rate increases. There are at least two
possible mechanisms for the transition. In the first, the TLCP structure "fails"
first at the fiber surface where the shear strain rate is largest. In the second, the
TLCP is "bound" to the fiber surface and becomes isotropic last.
Bhama and Stupp [63] showed that carbon fibers in TLCP made a tremendous
change in the formation of LCP structure. The fiber surface nucleated the crystal
structure and increased the temperature of the nematic/isotropic transition. The
LCP phase remained for several microns from the fibers surface after the rest of
the fluid became isotropic above the transition temperature. LDF/PEKK has an
average 1 lam thick PEKK layer around each fiber. Also, LCP capillary flow data
have shown that a liquid crystal structure may preferentially form at a treated
solid surface [64]. LCP layers up to 7 ~m thick formed a high viscosity layer
with a lower viscosity orientation present in the flowing core. Carbon fibers are
treated to control the interface with the polymer. The fiber surface morphology
and surfactants may influence the rheology at shear rates where the LCP structure
is stable. (One way to check this would be to add ground carbon fibers to PEKK
and test it in a cone and plate rheometer. The small particles at low volume
fraction could be run on the standard device.)
882

Under either mechanism this could create a "two fluid" system in the shear
cell. A viscosity difference of 10 or higher quickly concentrates the deformation
to the outer layer. This increases the shear rate for the outer fluid and increases
the effective fiber volume fraction. Changes in this direction would reduce the
difference in PEKK and system matrix viscosity as the strain rate decreases.
Thus the interaction of the TLCP melt with the carbon fiber surface may make
the shear cell model inapplicable at low shear rates. Some innovative techniques
would be needed to assess this effect. More of this is discussed in the modeling
that follows.

Giesekus Fit to Steady Extension Data


Since the elongation of the LDFMS engenders shear dominated response, a
constitutive model that describes stress overshoot in simple shear might be
applied to this material. Solution of the Giesekus equation by Runge-Kutta
method allowed f'mding the parameters that reach x y at the required time with
x y/'r E equal to the same ratio found by experiment. Figure 21 shows the best
Giesekus model fit at all three strain rates. From the figures one notes that near
quantitative fits of the data are possible at each k.

Giesekus Applied to Predict Interrupted Flow


Previous experiments and analysis show that a single parameter Giesekus
model suitably describes the behavior of LDF/PEKK in elongation at a fixed
strain rate [65]. But, the relaxation response of the single parameter model was
insufficient to describe the complex shitt of the fluid to larger characteristic times.
That is, the model would always relax the stress faster than the actual fluid. For
the model, the relaxation constant is less than ~,1 during the flow but can only
return to ~'1 as its maximum value during the relaxation period. To predict the
response to an interrupted flow experiment one must either model the relaxation
more precisely--so that the model reaches the same residual stress as the fluid in
the same time--or shorten the model's period of rest.

Allowing the Giesekus model to relax for 165 s would relax away most of the
stress at either strain rate. Instead, the relaxation period can be reduced to that
required to get the same residual stress prior to restarting the flow. Figure 22 and
Figure 23 show the predicted stress response to the resumption of flow when the
hold period drops to 30 and 35 s for the 10.3 and 104 models respectively.
Besides the error in the rheometer, in which steady strain rate flow was not
883

attained on the second extension, the model shows a good prediction of the
degree of stress overshoot upon resumption of flow for both strain rates.
10 o .

0
tO, 10 4

Ii
Source
lu
- Model
~ 10"= 9 10-5
E! lo-4
0 10-3

10 "=
0 100 200 300 400
Time(s)

Figure 21. Best Giesekus model fit with single parameter sets at all three strain
rates.

250

200 0 l~

000,~
13.. 150

I,M 100 Source

50 Model
0 LDFMS
i ... i J
o

0.00 0.05 0.10 0.15


Strain

Figure 22. Giesekus model applied to predict response to interrupted flow, 1.49
10-3 s-1. The single parameter fit covered the initial flow. The model relaxed to the
same stress level as the experiment and the flow rate resumed using the same
Giesekus parameters fit to the first extension.
884

200

150

C)O00 D O 0 0 0
v lOO
u.I
Source
50 Model
O ,,
LDFMS
I ,
0 (~ = 9 i

0.00 0.04 0.08 0.12 0.16


Strain

Figure 23. Giesekus model applied to predict response to interrupted flow, 1.66
10-4 s"~. The single parameter fit covered the initial flow. The model relaxed to the
same stress level as the experiment and the flow rate resumed using the same
Giesekus parameters fit to the first extension.

6. OUTLOOK

This section first summarizes the test and model results for both the
nylon/polyethylene (N/PE) and long-discontinuous-fiber/polyetherketoneketone
(LDF/PEKK) systems. Then the implications of applying these results to process
models are discussed. The prospect of timber experiments with new fiber
patterns is presented. Finally, a method of investigating microrheology and
LCP/fiber interaction is presented.

6.1 Conclusions
The evidence discussed above leads to the following conclusions about the
model micromechanics, the suitability of the constitutive equation and the
LDFMS properties.

6.1.1 Micromechanics Scaling


The N/PE model system shows that high volume fraction highly aligned fibers
in a specific overlap pattern generate shear-dominant flow when extended. This
effect makes the system's extensional viscosity greater than the neat polymer's by
885

changing the dominant strain rate from stretching to shearing. The induced shear
strain rate, which rises directly as a factor of the fiber aspect ratio, can stimulate
conformation changes (shear thinning, nematic/isotropic transition) in the
polymer. For a fixed volume fraction of fibers, extensional viscosity increases as
the square of the fiber aspect ratio. Finally, the extensional data follow from the
shear properties of the matrix fluid by micromechanics.

6.1.2 Constitutive Model


A shear-thinning constitutive model applicable to sheared polymers was fit to
the transient stress data from LDF/PEKK experiments. At the highest extension
rates, which had the greatest chance of making the PEKK matrix predominantly
isotropic, the nonlinear constitutive equation successfully predicted the size of the
stress overshoot after stress relaxation that followed an initial extension at two
different strain rates. The combination of micromechanics and constitutive
relation can provide a method for predicting the forming loads-- especially for
better understood polymers.

6.2 Implications for Process Application


The results of these experiments bear upon the planning of finite element
prediction of a forming process and upon the design of fiber filled systems for
thermoforming. The models needed to predict forming of these systems are
complex and computationally extensive. They must also address fiber rotation
and deconsolidation.

6.2.1 Complex Models


The nonlinear behavior of systems with large L / D fillers in extension raises a
computational challenge for f'mite element modeling of a forming process. Prior
researchers have opted for a "solid equivalent" constitutive model technique that
allows codes existing to predict stresses under applied deformation [66, 67, 68].
This strategy's shortfall is the over simplified stress to strain relationship, which is
suited to Newtonian fluids under small deformations. These methods employ a
basic relation of
[a] =[c][E] (21)
where the CO9 compose the stiffness matrix with reduced terms by virtue of the
symmetry of the material. For the solid equivalent method the right side of the
equation becomes
[a(t)] = In]I?] (22)
The stiffness matrix is now a viscosity matrix. Terms like Tlll and 1"122
represent the extensional and shear viscosities respectively. This expression
886

misses the polymer's nonlinear response and the effects of large deformations in
contrast to the small deformations of solids. The Giesekus equation in a similar
matrix form is
f ([x],[~:],['i'],t)= g([~'],[~?],t) (23)
which is more involved. Numerical techniques must include the transient
effects since stress overshoots can be significant and we need the total stress
history to deal with changes in deformation rates in simulated tool contact.

6.2.2 Fiber Rotation


A short series of tests checked the effect of fibers oriented at an angle to the
extension direction for LDF/PEKK. Fiber angles of five and 10 degrees and +45
laminates showed that rapid fiber rotation must be quantified if final fiber position
in multiple ply laminates is to be predicted. The "hyperanisotropy" [69] of
LDF/PEKK makes the material effectively unextendable in the fiber direction
until the fibers rotate parallel to the applied deformation [70]. Then the loads rise
rapidly as the shear cell effect occurs. Therefore, the off-axis fiber behavior is
identical to continuous fiber systems. Numerical software must handle the
transition from an interlaminar shear mode to shear cell approach as the fibers
orient.

6. 2.3 Deconsolidation
Two deconsolidation mechanisms occur in sheet forming. When melted in
open air without any applied deformation, N/PE and LDF/PEKK deconsolidate.
Residual stresses in the sheet separate the plys and lott the material. Tensile
experiments show that forming enhances deconsolidation without an applied
pressure. In the shear cell model, the relative motion of the fibers generates
substantial normal forces. This normal force would push fibers apart. Raising the
strain rate increases the effect.
At the low strain rates used here the specimens remained consolidated within
either the metal fixture or vacuum bag. Consolidation maintained the continumn
so that the analysis could be completed. In some industrial forming processes no
consolidation force keeps the sheet together. The fmal forming step
reconsolidates the panels. Although this lowers the effective viscosity of the sheet
with the addition of voids [71], the void content changes the fiber motions in
ways that are very complex to model.

6. 2.4 Parameters that Affect the Process


This is a summary of some parameters in applying long fiber reinforcement and
goals in the control of the parameters.
887

Fiber Length
The material should contain the shortest possible average fiber length for the
application. This keeps the viscosity increase to the smallest level needed to
make a good part.

Fiber Length Distribution


A random distribution of fiber lengths should improve the forming of the
material. The strain rate achieved in the LDF/PEKK was much closer to the target
rate than that in the N/PE model. Distributed fiber length raises the effective
average fiber length [69] but some longer fibers can bridge gaps between the
shorter fibers.

Strain Rate Limitations


Two effects restrict the applied strain rate. The first is deconsolidation. Normal
forces must not overpower the consolidation force needed to maintain a
continuum. The second is inertial effect in the polymer. Two experiments with
LDF/PEKK at a target strain rate of 0.01 s~ and a target stress of 400 MPa
provided a glimpse of this limitation. At 0.01 s"1 the material reached a peak
stress of 380 MPa followed by a precipitous drop in the tensile load. When the
moving grip stopped the specimen did not. It became larger than the grip to grip
distance and went into compression. Similarly the 400 MPa test obtained a great
strain rate that caused the computer to halt the test at the safe extension limit of
the test frame. Again the specimen continued to flow although the test machine
had stopped.

6. 2. 5 Microrheology
Recent research into the microrheology of polymers shows that thin film and
bulk properties may be very different [72, 73, 74, 75, 76, 77]. Since the flow
condition of interest to this research is annulus flow, we suggest a
microrheometer that simulates this condition. A capillary rheometer with a fiber
inserted through the hole of the capillary. With the proper selection of the fiber
diameter the device would span the range of bulk (no fiber inserted) to
microrheology. Two methods of measuring the viscosity of the polymer are
possible with this device. The first method would be to rtm the instrument as a
capillary rheometer with the analysis of the data accotmting for annular flow. The
fiber insert may be either stationary or moving. The difference in cross sectional
area between the reservoir and the annulus assures that the polymer flows at peak
velocities much greater than that of the fiber surface.
888

A second method for collecting the data is to use the filled capillary/polymer
system in a "pull-out" test. The filled capillary/polymer system would be placed
in a test machine and heated to the appropriate temperature. Then the fiber would
be withdrawn from the capillary at various speeds as the data acquisition system
records the load on the fiber. This device could also determine the effect of LCP
structure on microrheology as discussed in the next section.

6.2. 6 LCP/Fiber Interaction


The effect of solid surface characteristics on LCP structure formation and
stability is largely unknown although some experiments show dramatic effects
[63]. The microrheometer proposed above would allow systematic study of these
interactions. Fibers would be selected to change the surface from smooth and
nonadhering (an untreated glass fiber) to smooth and adhering (a coated glass
fiber), to rough (an etched glass or untreated carbon fiber), and finally to rough
and adhering (a coated carbon fiber). Either method could measure any viscosity
change that the fiber surface induces.

REFERENCES

1. I.Y. Chang and J.F. Pratte, Journal of Thermoplastic Composite Materials, 4


(1991), 227.
2. R.M. Jones, Mechanics of Composite Materials, (1975), 1.
3. J.L. Throne, Thermoforming, (1987), 1.
4. R.K. Okine, D.H. Edison and N.K. Little, 32nd International SAMPE
Symposium, 32 (1987), 1413.
5. P.J. Mallon, C.M. O'Bradaigh and R.B. Pipes, Composites, 20 (1989), 48.
6. F.N. Cogswell, International Polymer Processing, 1 (1987), 157.
7. A.S. Tam and T.G. Gutowski, Journal of Composite Materials, 23 (1989),
587.
8. D.W. Coffin, Flange Wrinkling and the Deep-Drawing of Thermoplastic
Composite Sheets, (1993), 1.
9. D. Hull, An Introduction to Composite Materials, (1981), 1.
10. I.Y. Chang and B.S. Hsiao, 36th International SAMPE Symposium, 36
(1991), 1587.
11. I.Y. Chang, 37th International SAMPE Conference, 37 (1992), 1276.
12. F. Rodriguez, Principles of Polymer Systems, (1982), 1.
13. R.B. Bird, R.C. Armstrong and O. Hassager, Dynamics of Polymeric
Liquids, Volume 1 Fluid Mechanics, (1987), 1.
14. S. Matsuoka, Relaxation Phenomena in Polymers, (1992), 1.
889

15. P.J. Carreau, D. De Kee and M. Daroux, Canadian Journal of Chemical


Engineering, 57 (1979), 135.
16. J.C. Maxwell, Phil. Trans. Roy. Sot., A157 (1867), 49.
17. J.L. White and A.B. Metzner, J. Appl. Polym. Sci., 7 (1963), 1867.
18. H.J. Mthastedt, Rheol., 24 (1980), 847.
19. F.P. La-Mantia, A. Valenza and D. Aciemo, Polym Eng Sci, 28 (1988), 90.
20. H.M. Laun and H. Schuch, J Rheol, 33 (1989), 119.
21. J. Meissner, Polym Eng Sci, 27 (1987), 537.
22. J. Meissner, Chem. Engr. Commun., 33 (1985), 159.
23. V.M. Lobe and J.L. White, Polym. Eng. Sci., 19 (1979), 617.
24. S. Barbosa, M.A. Bibbo and A. Miguel, Applied Polymer Symposium, 49
(1990), 127.
25. M.A. Bibbo and R.C. Armstrong, Proceedings of Manufacturing
International '88, 105 (1988), 123.
26. A. Jamil, M. S. Hameed and A. Stephan, Polymer-Plastics Technology and
Engineering, 33 (1994), 659.
27. O. Lepez, L. Choplin and P. Tanguy, Polym Eng Sci, 30 (1990), 821.
28. A.T. Mutel and M.R. Kamal, Two-Phase Polymer Systems, (1991), 305.
29. L.A. Utracki, Polym. Compos., 7 (1986), 274.
30. A. Vaxman et al., Polym. Compos., 10 (1989), 78.
31. S N. Maiti and P K. Mahapatro, Polym. Compos., 9 (1988), 291.
32. R. Hingmann and B.L. Marczinke, J. Rheol., 38 (1994), 573.
33. T.M. Malik et al., Polymer Composites, 9 (1988), 412.
34. H.M. Laun, Colloid & Polymer Science, 262 (1984), 257.
35. A.T. Mutel and M.R. Kamal, Polym Compos, 7 (1986), 283.
36. S.E. Barbosa et al., Composite Structures, 27 (1994), 83.
37. M.L. Becrafl, The Rheology of Concentrated Fiber Suspensions, (1989), 1.
38. S. Toll and P.O. Andersson, Polymer Composites, 14 (1993), 116.
39. N. Dontula et al., J. Reinforced Plastics and Composites, 13 (1994), 98.
40. M.R. Kamal, A.T. Mutel and L.A. Utracki, Polymer Composites, 5 (1984),
289.
41. S.M. Davis and K.P. McAlea, Polymer Composites, 11 (1990), 368.
42. G.K. Batchelor, Journal of Fluid Mechanics, 46 (1971), 813.
43. J. Mewis and A.B. Metzner, Journal of Fluid Mechanics, 62 (1974), 593.
44. J.S. Bums, The Influence of Aligned, Long-Fiber Array (ALFA)
Reinforcements on Composite Manufacturing and Structural Performance,
(1995), 1.
45. F.N. Cogswell, Polymer Melt Rheology, (1981), 1.
46. R.M. Patel and D.C.Bogue, J. Rheol., 33 (1989), 607.
890

47. C.J.S. Petrie, Elongational Flows: Aspects of the Behavior of Model


Elasticoviscous Fluids, (1979), 1.
48. M.R. Kamal and A.T. Mutel, Polymer Composites, 5 (1984), 289.
49. R.K. Gupta, Flow and Rheology in Polymer Composites Manufacturing, 10
(1994), 89.
50. Secor, R.B. et al., J. Rheol., 33 (1989), 1329.
51. R.M. Christensen, Theory of Viscoelasticity, (1982), 1.
52. J. Meissner, Kunststoffe, 61 (1971), 576.
53. K.F. Wissbrun, Journal of Rheology, 25 (1981), 619.
54. A. Ciferri, Liquid Crystallinity in Polymers: Principles and Fundamentals,
(1991), 1.
55. S.M. Guskey and H.H. Winter, J. Rheol, 35 (1991), 1191.
56. A.B. Metzner, Rheol. Acta, 10 (1971), 434.
57. J.D. Goddard, Journal of Fluid Mechanics, 78 (1976), 177.
58. A.J. Beaussart, J.W.S. Hearle and R.B. Pipes, Composites Science and
Technology, 49 (1993), 335.
59. R.B.Pipes et al., Proceedings of the American Society for Composites, 159
(1992), 123.
60. R.B. Pipes et al., Journal of Composite Materials, 28 (1994), 343.
61. H. Giesekus, Journal of Non-Newtonian Fluid Mechanics, 11 (1982), 69.
62. K.A. Ericsson, S. Toll and J.E. M~tnson, Rheol. Acta preprint, (1996).
63. S. Bhama and S.I. Stupp, Polym Eng Sci, 30 (1990), 228.
64. J. Fisher and A.G. Frederickson, Mol. Cryst. Liq. Cryst., 8 (1969), 267.
65. T.S. Creasy and S.G. Advani, Developments and Applications of Non-
Newtonian Flows, 66 (1995), 123.
66. T.G. Rogers, Composites, 20 (1989), 21.
67. C.M. O'Bradaigh and R.B. Pipes, Composites Manufacturing, 2 (1991),
161.
68. D.W. Cot~n and R.B. Pipes, Composites Manufacturing, 2 (1991), 141.
69. R.B. Pipes et al., Journal of Composite Materials, 25 (1991), 1379.
70. N.J. Pagano and J.C. Halpin, J. Composite Materials, 2 (1968), 18.
71. S.F. Shuler et al., Polymer Composites, 15 (1994), 427.
72. B.A. Costello and P.F. Luckham, Materials Research Society, 289
(1993), 7.
73. I. Hersht and Y. Rabin, Journal of Non-Crystalline Solids, 172 (1994),
857.
74. E. Pelletier, J.P. Montfort and F. Lapique, Journal of Rheology, 38
(1994), 1151.
75. K.D. Danov et al., Chemical Engineering Science, 50 (1995), 263.
891

76. J.A. Tichy, Tribology Transactions, 38 (1995), 577.


77. M. Urbakh, J. Klafier and L. Daikhin, Materials Research Society, 366
(1995), 129.

NOMENCLATURE

English
ar Aspect ratio of a particle, the major length divided by the minor
length.
H(~.) The height of the relaxation spectnun as relaxation parameter
increases from zero to infinity.
L/D Aspect ratio of a particle, the major length divided by the minor
length.

Greek
O~ Parameter controlling the non-linear response of the Giesekus
model. Varies from zero to one.
Axial density distribution with units of g/cm.
13(x) Axial density distribution as a function of position along the
specimen.
Shear strain rate in sec ~.
rx(r) Shear strain rate in cylindrical coordinates for the shear cell
model.
~yx Shear strain rate in cartesian coordinates.
s Extension strain.
Extension strain rate in secf1.
rio Zero shear rate viscosity; viscosity of a polymer melt in the linear
viscoelastic strain rate range.
I"IE Extension viscosity of a fluid in tension.
TlEapp Apparent extension viscosity measured in a region with a strain
rate gradient.
Transient extension viscosity measured from the startup of flow
from the rest state.
I"1S Shear viscosity in steady shear flow.
Transient shear viscosity measured during startup of flow from
the rest state.
892

rls Shear viscosity decay measured during the time following steady
shear flow.
Carreau equation parameter that controls the onset of shear
thinning.
Relaxation parameter in fluid constitutive equations.
Z'max Relaxation parameter value at which 99.9 % of the area udner the
relaxation distribution function H(E)is attained.
~p Relaxation parameter value at which H(~) attains its peak value.
"~yx Shear stress in units of Pa.
+
1;yx Transient shear stress measured from the start up of flow until
steady flow is obtained.
1;yx Shear stress decay following flow. Deformation held constant
during the decay.
The peak value of the transient shear stress. A 'yield' value for
the conformation change in the polymer.
'[E Extension stress in tension.
+
17E Transient extension stress measured from the start of flow from
rest.
Peak value of tensile stress.
Stress relaxation measured atter extension.
893

THERMOMECHANICAL MODELING OF POLYMER


PROCESSING

J.F. Agassant, T. Coupez, Y. Demay, B. Vergnes, M. Vincent

Centre de Mise en Forme des Mat~riaux, Ecole des Mines de Paris,


URA CNRS 1374
BP 207, 06904, Sophia Antipolis , France

1. INTRODUCTION
Polymer processing involves complex flow geometries, in the plasticating
units (single or twin screw extrusion, injection) as well as in the shaping tools
(die, mold). Thermoplastic polymer processes started only around sixty years
ago and these processes have been firstly developed by trial and error.
Polymer processing modeling is a recent story and appears as a useful tool, not
only to limit the number of trials, but also to master all the thermomechanical
parameters which will induce, for example, crystallization, macromolecule
orientation and, as a consequence, end-use properties of the produced part.
Modelling polymer processes appears as a challenge for several reasons :
9 molten polymers are highly viscous materials, which leads to important
viscous dissipation ;
9 polymer processes may rarely be considered as isothermal;
9 molten polymers are non-Newtonian viscoelastic fluids, which means for
example:
- shear-thinning behavior for the viscosity ;
- transient effects ;
- normal stress differences in pure shearing flow.
9 the viscosity of molten polymers is temperature dependent, which implies
that mechanical and heat transfer equations have to be solved simultaneously.
9 polymers present a low thermal diffusivity, which may lead to important
temperature gradients, even when the polymer is flowing in very thin gaps;
9 most of the polymers are semi-crystalline materials and very specific
structure developments (spherulites, for example) may be observed
depending on the cooling rate and stress field encountered during the process.
In this chapter, we will only consider the flow of an homogeneous molten
polymer. Neither plastication mechanism nor crystallization kinetics will be
taken into account. In a continuum mechanics approach, the different equations
894

to solve are the mass, momentum and thermal balances, linked by a


constitutive equation and appropriate boundary conditions. These governing
equations are now described in details.

1.1 Mass balance


Molten polymers can generally be considered as incompressible materials,
which leads to 9

V.u : 0 (1)

where u is the velocity field. When very high pressures are encountered (as
for example in injection molding), the variation of the material density p in
time and space has to be taken into account :

@
0t + V.(pu) = 0 (2)

1.2 Stress balance


The stress tensor o" is symmetrical and only force balances have to be
considered:

V.(r+ F - p?'= 0 (3)

where F represents the gravity forces and p?'the inertia forces.


The Reynolds number Re compares inertia and viscous terms 9

Re= p Uh
7/ (4)

where U is the characteristic velocity of the flow, h the flow gap and 7/the
viscosity. Generally, Re is negligible in polymer processing or, at a maximum,
of the order of magnitude of several units (this is the case around gates and
runners in injection molding or at high speed fiber spinning).
To compare gravity and viscous forces, an equivalent Stokes number is
defined:

St- pgLh
rl U (5t

where g is the gravity and L the vertical dimension of the flow. In horizontal
processes (extrusion, injection molding), gravity forces are negligible. This is
no more the case when large vertical stretching distances are considered (fiber
895

spinning, film blowing). When both gravity and inertia forces are neglected,
the momentum equation reduces to 9

V.o-= 0 (6)

1.3 Constitutive equations


Molten polymers are generally viscoelastic, but several constitutive
equations may be used, depending on the polymer, the flow geometry and the
level of approximation one wants to use.

1.3.1. The Newtonian behavior


a=-pl+27/ e (7)
where p is the pressure, I the identity tensor and e the rate of strain tensor,
defined as :

= ~ (Vu + Vu t) (8)

The Newtonian behavior is a crude approximation, but it may provide


reasonable results when the rate of strain remains quite uniform within the
flow geometry. It allows analytical calculations in simple flow geometries,
which is very useful to test the validity of numerical methods.

1.3.2 The generalized Newtonian behavior :


The viscosity 7/is a spatial function which depends on the temperatureT and
on the second invariant of the rate of strain tensor"

a = - p l + 2 rl(T, ~) [~ (9)

where the second invariant is expressed as 9

42 '""'
92
(10)
ld

Several functions may be proposed for the viscosity :


9 the so-called power-law :
11= K(T) ~ n-1 (11)

where n is the power-law index and K the consistency, which is only a function
of temperature, following for example an Arrhenius law 9
896

E 1 1
K ( T ) = Ko exp ~ (~ - ~00) (12)

E is the activation energy, which may vary significantly from one polymer to
another, R is the ideal gas constant and Ko is the value of the consistency at the
reference temperature To. The advantage of the power-law is to provide
analytical solutions in a wide range of flow geometries, but its main drawbacks
are an infinite viscosity at zero shear rate and the absence of Newtonian
plateau at low shear rate.

9 the Carreau law [ 1] :

77 = r/0(T) [1 + (A 2]

may be considered as a relaxation time. n is, as previously, the power-law


index and 770 is the temperature dependent viscosity of the Newtonian plateau.
This expression is also used under the following form, called Carreau-Yasuda
law [2], in which the parameter a describes the transition between the power-
law region and the Newtonian plateau:

(n-1)/a
= 00(73 [1 + (~ ~) a] (14)

These laws provide a precise description of the shear viscosity as a function


of the shear rate. However, even the solution of a simple Poiseuille flow
requires a numerical approach.

1.3.3 The viscoelastic behavior


Viscoelastic constitutive equations are numerous and, at that time, it is still a
difficult task to select one which accounts for a large number of viscoelastic
phenomena : existence of normal stress differences in pure shearing flows,
transient phenomena in strain or in stress steps, strain hardening in
elongational situations, extrudate swell... Two families of constitutive
equations are encountered 9 the differential and the integral models.
9 The simplest differential model is the Maxwell model :

cr = - p ' l + s (15.1)

as
s+2~=2r/ e (15.2)

where p ' is the isotropic part of the stress tensor, s the extra-stress tensor and
5 s / & the upper-convected derivative. It is to notice that this Maxwell model is
897

a generalization of the crude dashpot-spring linear model, but it may be also


derived from the elastic dumbbell model (Rouse [3], Zimm [4]).
The Oldroyd-B model [5] is derived from the previous one by adding a
Newtonian contribution r/s"
cy--p'l + 2 rls e + s (16)

It allows to account for the two normal stress differences in simple shear
flow (a first normal stress difference, as for the Maxwell model, and a second
one), but also for strain transition after a stress step. Jeffreys models [6]
consist in introducing an additional function of the extra-stress tensor in
equation (15.2) :

&
f(s) s + A-~= 2 r/ e (17)

For example, for the well known Phan Thien-Tanner model [7], we have :

e2~
f ( s ) = (1 + - ~ tr s ) I (18)

where e is a material function, and tr s is the trace of the extra-stress tensor s.


Other functions may be introduced, for example in the Giesekus model [8] :

a~
f(s) = I + s (19)
77

These models may be generalized by using a spectrum of relaxation times


(A,i ,r/i) instead of a single one. Multimode Maxwell or Phan Thien-Tanner
models are expressed as [9] :

~si
S - ~. S i , f(si) Si + / ] , i - - ~ - 2 17i ~ (20)
1

They account simultaneously for shear viscosity, elongational viscosity and


first normal stress difference in simple shear.
9 The simplest integral model is the Lodge model [ 10] 9

s- f t m (t, t ,) C i 1 (t, t ,) dt' (21)


- C>O

where m(t, t') is a memory function and C t I the Finger tensor. This model is
equivalent to the Maxwell model by choosing :
898

m (t, t') = exp( t t) (22)

Wagner [11 ] improved the Lodge model by introducing a damping function


h of the two invariants I l and 12 of the Finger tensor 9

s = f t m (t, t ,) h (1 i, 12 ) C i I (t, t ,) dt' (23)


-00

Different forms of the damping function may be found in the literature.


For example, Papanastasiou et al. [ 12] proposed for h the following equation 9

h (I l, I2) = 1
1 + a [fl 11 + ( 1 - [ 3 ) I 2 - 3] b/2 (24)

where a,/3 and b are material parameters.

1.4 Energy balance equation


The most general form of the energy balance equation is 9

de
p = - v.q + (25)

e is the mass density of internal energy ; for an incompressible material, e


is proportional to the temperature-
de dT
dt = Cp dt (26)

where Cp is the heat capacity. More complex equations are proposed for
compressible materials with phase transition [13].
q is the heat flux, which is proportional to the temperature gradient
following the Fourier law 9

q =- k V T (27)

where k is the heat conductivity.


I~ = o'" ~" is the viscous dissipation, expressed as 9
.

O'" 8" = 7/ y 2 for N e w t o n i a n materials, (28.1)


899

a " e" = K ~ n + 1 for power-law fluids. (28.2)

It is to notice that for a viscoelastic constitutive equation, all the energy is


not dissipated and equation (25) has to be modified [14].

1.5 Boundary conditions


In order to solve mass, stress and thermal balance equations, we need to
define relevant boundary conditions.

1.5.1 Mechanics and kinematics


A zero velocity at the wall (sticking contact) is generally assumed in most
polymer flows. This is reasonable for thermoplastics at low or intermediate
flow rates. At high flow rates, flow instabilities are encountered, which may
correspond to stick-slip transition [see Chapter 7, section 7.3]. When
processing PVC, rubber compounds, or highly filled suspensions, slip at the
wall may occur even at low flow rates. The determination of accurate slip
velocity measurement methods and of relevant slip constitutive equations
remain an open problem [see Chapter 4, section 4.2].
Depending on the problem to solve, a velocity profile, a pressure or a
pressure gradient have to be prescribed in the inlet section. In the particular
case of viscoelastic constitutive equation, the extra-stress components at the
entry surface of the flow domain have also to be defined. At the die outlet, a
zero pressure is generally assumed.
In stationary free surface flows, a velocity vector parallel to the free
surface, as well as a zero stress component perpendicular to the free surface,
are considered.

1.5.2 Temperature and heat transfer


A temperature profile has to be known at the entry surface of the flow
domain. The more difficult problem is to determine accurate boundary
conditions along the processing tools (extruder, die, mold) or along free
surfaces.
It is customary to fix the temperature at the wall (T = Tw) ; this is the case
when precise and powerful thermal regulation systems are used. This could
also be used as a first approximation to check a preliminary value of the
temperature field. Generally, one imposes a heat flux q or a heat transfer
coefficient hr"

q - hr (Tw - Te) (29)

where T e is the controlled or measured temperature of the tool. In a flat mold,


for example :
900

km
hr--- l (30)

where km is the heat conductivity of the metal and l is the average distance
between the cooling channels and the mold wall. When the geometry of the
mold (or the die) is complex, l may be difficult to determine and it is
preferable to develop a global computation in the polymer domain and in the
tool (with, for example, an iterative loop between polymer flow and tool).
For free surface problems (fiber spinning - cast film - film blowing ...),
the determination of a realistic heat transfer coefficient remains a challenge
because coupled and complex phenomena may be encountered : free
convection, forced convection, radiation ... [13]. Very often, the heat transfer
coefficient (or the heat transfer function, because it can vary for example
between the die and the winding system in fiber spinning) will be considered as
an adjustable parameter.

1.6 Scope of the chapter


Mass balance, stress balance, energy balance and constitutive equations with
appropriate boundary conditions have now to be solved in the non trivial flow
geometries encountered in processing equipments. Finite elements methods are
generally used and the accuracy of the results will significantly depend on the
precision of the mesh. In the next section, several complex 3D flow geometries
will be considered. However, for some processing geometries, approximation
methods may lead to simplified solutions with a reasonable accuracy. This will
be presented in section 3 for confined flow situations and in section 4 for free
surface flows.

0 DIRECT SOLUTION FOR POLYMER FLOWS USING


THE F I N I T E E L E M E N T M E T H O D

2.1 Viscous flow problem


As shown in the first part of this chapter, the molten polymer can be
considered as incompressible and the inertia and mass forces are neglected. A
purely viscous isothermal polymer flow is then described by combining
equations (1), (6) and (9), giving the following mixed velocity-pressure
problem:

Find (u, p) solution of"


V.[2r/( ~ ) ~: (u)- Vp ]= 0 (31.1)

V. u = 0 (31.2)
901

+ boundary conditions

The bounded domain f2 of boundary o~ is the region occupied by the fluid or


more precisely the domain of calculation as shown in Figure 1. The above
problem will be well posed if adequate boundary conditions are prescribed.
For instance, in the extrusion die of Figure 1 (in fact, Figure 1 presents the
internal volume of the die), the pressure is imposed at the inlet and outlet of
the flow and a zero velocity is prescribed along the walls.
The Newtonian law leads to a classical Stokes problem, considered here as a
model problem to point out one of the difficulty in solving such a viscous flow
problem : the treatment of the incompressibility condition. The Stokes
problem has been early studied as a subset of the Navier-Stokes equations [15]
and gave rise to the mixed theory, which general framework is now well
established [ 16].

Figure 1" Boundary conditions associated with a profile die extrusion


flow problem

2.2 Variational formulation


The finite element approximations are based on a weaker form of equations
(31), also known as the virtual work principle. In order to simplify this
presentation, we suppose that the velocity is prescribed everywhere along the
boundary. Let Vand P (the Sobolev space V= (HI(~)) 3 and P = L2(~) for the
Newtonian case) be respectively the velocity and pressure spaces. The problem
to solve can be rewritten as 9
902

Find (u,p) ~ V x P , u = u 0 on bfl, u0 being given, such as :

~ 2r I i: (u ) " e (u *) dr2 - ~pV.u* df2 = O (32.1)

rj p *V.u dr2 = 0 (32.2)


f2

V(u*, p*) ~ V 0 x P, where" V 0 = {u* ~ V, u * l ~ = 0}

The velocity and the pressure must be computed simultaneously. Most of


the numerical schemes are based on a particular choice of the pressure
interpolation with different solution techniques. The penalty method associated
with the reduced integration technique [17] corresponds to a discontinuous
interpolation of the pressure, theoretically well understood in the robust
augmented Lagrangian technique [18,19] used in 3D [20].
The mixed finite element formulation presented here is based on a simple
and almost natural continuous interpolation of the pressure, entering in the
mini-element family [21 ].

2.3 Finite element discretization


The fundamental idea of the finite element method is to approximate spaces
V and P by discrete spaces V h and Ph. The domain of calculation f2 is
decomposed in a finite family of simple geometrical elements f2e. The
numerical method will depend on the choice of the geometrical element. In
3D, hexahedral elements are often used, but restricted to relatively simple
geometries. The general way to solve the mesh generation problems and to use
automatic meshing methods is to use tetrahedral elements [22]. Moreover, it is
possible to mesh complex three-dimensional geometries, such as extrusion dies
for example, and to control the local mesh size, particularly at the very thin
exit [23]. Figure 2 shows the surface of a mesh of the extrusion die example,
composed of more than 50 000 tetrahedra.
The accuracy of the numerical solution will depend on the mesh size
parameter h which is defined as the maximum of the elements diameters. It
will depend also on the polynomial interpolation order, defined as follows.
Assuming a continuous velocity approximation, a pressure approximation
which can be continuous or discontinuous, and tetrahedral elements (triangular
elements in 2D), the discrete spaces are defined by 9

V h = {u h ~-. (C 0 (~))n, u h I ~ e E (pk (f~e))n} (33)


903

Figure 2" External view of the mesh of a die extrusion geometry and
detail of the mesh in the exit section

ph = {ph E (C i (~), ph I ~e ~ (pl (~e)} (34)

where pk(f~e) and pl(f~e) are the set (or sub-set) of polynomials of degree k
and l on element f~e. n =2, 3 is the space dimension. If i = 0, the functions are
continuous, if i = -1, they are discontinuous.
V h and p h cannot be chosen independently. The inf-sup Brezzi-Babuska
condition must be checked (see [16] for a complete discussion of this
compatibility condition). A well established technique to ensure this condition
is to enrich the velocity interpolation with a bubble part [19, 24] (Figure 3).
The discrete velocity space is augmented (in the hierarchical form) by the
discrete bubble space denoted by :

B h = {bh ' bhlf~e ~ (H~(~e))3 } (35)

The velocity takes the following form"

v h = u h + b h ~ (V h + B h) (36)
904

The bubble has the advantage to be local and it can be condensed at element
level without changing the bandwidth of the global system. This technique gave
rise to the simplest mixed finite element : the mini-element [21 ]. In theory, the
shape of the bubble function is free and it has only to vanish at the element
boundary. We use a pyramidal bubble function (Figure 3) preserving the exact
integration property of a first order tetrahedral element. The bubble can be
chosen optimally [25] and it can be related to the stabilization technique [26].
The mini-element is used for the calculation of the three-dimensional examples
presented in this paragraph. It has four unknowns per node.
In two-dimensional calculations (viscoelastic examples presented later), we
used the Crouzeix-Raviart element (also known as the P2+/P1). When coupled
with an augmented Lagrangian technique [19], the pressure unknown can be
locally eliminated, which leads to two unknowns per node.
The main difference between mini and Crouzeix-Raviart elements is that the
pressure is continuous for the first one and discontinuous for the second one.
Despite the increase of degrees of freedom with the mini-element, the
continuous interpolation of the pressure is, in 3D, largely less expensive in
term of computational resources.

.;. . . . . .
.Q"

Figure 3 9 3D mini-element with a pyramidal bubble function

Under this stability condition, the numerical solution converges with respect
to the mesh refinement to the real solution. The approximation in velocity and
pressure are then dependent through the a priori following estimate :

Ilu - uhll 1,f~ + lip - ph IIL2(f~) < C h i (37)

For the mini-element, i - min(k, l + 1) = 1, k and 1 being respectively the


interpolation order of the velocity and the pressure. It is then a first order
element, since i = 2 for the second order Crouzeix-Raviart element.
905

Equation (37) can be rewritten by introducing the strain rate numerical


error 9

[~~z (71 I ~(u ) - "C(uh) I 2 + I p - p h I 2)dE').,]1/2 <__C'h i (38)

and finally this can also be related to the error in stress by using the
constitutive relation"

l-j I I 2dC ]1/2 _< C"hi (39)

For the example of extrusion die, Figure 4 shows how the flow rate evolves
with the mesh refinement for a fixed pressure drop.

700

~. 600
O~

500
E
400 ~
300
0 0

u. 200

100 9 ! ' T ' I

0 5000 10000 15000 20000


Number of mesh nodes

Figure 4 9 Evolution of the calculated flow rate with the mesh refinement.

When the mesh is composed of more than 10 000 nodes (for instance the
mesh of Figure 2), it reaches an asymptotic value, independent of the mesh. In
this example we used a linear interpolation of the velocity (plus a pyramidal
bubble term for stability purpose) and a linear interpolation of the pressure,
giving a first order element.
Under these conditions, we obtain the results presented in Figure 5, which
are almost free of numerical approximation. The pressure remains quite
constant in the upstream die region and a high pressure gradient appears at the
die exit. The goal of such a calculation is to compute the velocity profile at the
die exit in order to predict the balance of the polymer flow and then to
validate or modify the proposed geometry.
906

Figure 5 : Pressure field in the whole die (a) and velocity field at the die
exit (b)

2.4 l t e r a t i v e solver
An attractive feature of the mini-element interpolation is the possibility to
solve the large linear discrete systems by an iterative method [27]. This is
crucial for 3D applications for which the use of direct solvers is too much
consuming, both in term of memory and in term of CPU.
By using a static condensation technique [19], the internal velocity bubble
terms vanish from the system. This elimination provides a new diagonal block
matrix C. The remaining linear parts of the velocity and the pressure are then
solution of the following system :

B
H _BC ) ( pV) = ( F ) (40)

where H and C are symmetric definite positive. The unknowns are thus the
three components of the velocity and the pressure for each vertex. The size of
907

the system is then 4 N, N being the number of mesh nodes. The new block -C
can be seen as the optimal stabilization matrix in the context of a stabilized
P1/P1 method [28]. The form of the blocks remains almost the same in the
non-linear case, with H = H(V) and F = F(V), V being fixed at each Newton
iteration. The global matrix is symmetric indefinite and an iterative approach
must be based on a residual method. The preconditioned conjugate residual
method (PCR) for the Stokes problem has been introduced in [29]. The PCR
method can be seen as a particular case of the generalized minimal residual
method (GMRES) [30], which can be used with indefinite possibly non-
symmetric matrices. However, from the practical point of view, it can be seen
as a preconditioned conjugate gradient extended to saddle point problems. The
method used in [27] for 3D forming applications is based on the hybrid
Orthomin/Orthodir form of the algorithm, as described in [31] for instance.

2.5 Thermal resolution


2.5.1 Convective diffusive equation
Using equations (25), (26), (27), the heat transfer equation has the
following form"

dT
pep = k AT+ (41)

The time derivative of the temperature can be decomposed in a heat


variation at fixed point and a convective term 9

d T OT
dt = oat + V T.u (42)

From the mathematical point of view, equation (41) can be of different


types, depending on the relative weight of each of its terms. When the flow
velocity is zero or almost zero, the equation is dominated by the diffusive
term. The problem is then elliptic for stationary problems or almost elliptic
when the time variation is relatively small (cooling for instance). A standard
Galerkin method will be appropriate for numerical solution. When time
variation of the temperature is very fast, the equation tends to a parabolic type
and its solution may exhibit shock. That is the case in injection molding at the
contact between the cold wall and the hot polymer : the flow near the mold
wall is slow due to the sticking contact since the temperature gradients become
important. A relation must be respected between the time step and the mesh
size, leading finally to adopt an adaptive anisotropic meshing technique [32].
Figure 6 shows an example of refinement of the mesh close to the wall in the
case of the filling of a cavity. Convection is dominating in the main stream and
the heat transport solution must be accurate. The standard finite element
908

methods often fail when the temperature gradients are important. Among the
possible alternative method we can mention the streamline upwind technique
[33], which consists to add diffusive terms in the standard Galerkin finite
element method. These terms must be smaller of one order of magnitude to
preserve the consistency of the equations. The standard discontinuous Galerkin
method is well suited for hyperbolic equation, but restricted to purely
convective term. The method of characteristics [34] is well suited both for
convective and dissipative dominated problem.

Figure 6 : Anisotropic mesh of the filled area in injection molding

2.5.2 Finite element solution of the thermal equation


The weak form of the equation is obtained by a standard Galerkin
technique. The equation is multiplied by a test function, T*, chosen in an
appropriate functional space T 0. T* is vanishing at the boundary and the
equation is integrated by part. The problem to solve is thus 9
Find T ~ T(= HI(E~)) such as :

f~ pCp -~
dT T* dE~ =- ~ kVr.Vr* dE~ + f W T* dE~, VT* ~ T O (43)

We assume that T is prescribed on the boundary. In the standard finite


element method applied to these equations, we introduce one discrete space T h
for the temperature field and the test functions.

pCp --~T*hd~ =-~ kVT h VT*hdE~ + WhT,hdE2, VT,h ~ T0h (44)


909

dTh
The evaluation of--d~ will differ from one scheme to another and will lead
to symmetric or non-symmetric systems. For instance"

dTh Th(x(O)- Th(x(t-at))


= lim (45)
dt at~o At

will lead to the method of characteristics [34], which requires to compute at


each integration point the upstream trajectory on a length determined by the
time interval At. The advantage of such a method is to introduce only a mass
matrix for the discretization of the convective term and then to maintain a
global symmetric system. The drawback is to be non-local. It has been
intensively exploited to perform the thermomechanical coupling in injection
molding simulation [32].
On the other hand, a standard Galerkin method leads to a non-symmetric
system:

f~pCp--~T*hdE~
0Th + ~ VTh.uhT*hd~ + ~EkVTh.VT*hdE~- f I~hT*h dE~
(46)

This method is quite natural but can show some instabilities and it is restricted
to small temperature gradients. The other difficulty is to solve such a non-
symmetric system by an iterative solver (direct 3D solutions by means of
Choleski factorization are too expensive to be really usable). The minimal
residual method (which can be adapted to non-symmetric system) works when
the gradient along the flow remains small enough. Afterwards, one can
observe loss of convergence, even with the GMRES method [30]. The direct
Galerkin method using a minimal residual method as iterative solution was
used for the profile die extrusion application [35].
Very recently, we have proposed an explicit thermal solution algorithm.
The heat flux is introduced in the general heat transfer equation as a new
unknown variable, leading to a two fields problem"

cgT
pCp Ot = - pCp V T.u - V . q + W (47.1)

q =- k V T (47.2)

These equations are solved explicitly by using a local Taylor-Galerkin


scheme. The interpolation is performed by a piecewise constant per element
for both the temperature field and the heat flux field. The advantage of this
910

discontinuous interpolation is to control the thermal shock between the cold


wall and the hot polymer in injection molding application and to avoid any
numerical instability.

2.6 Application to 3D mold filling


Let us introduce f2 as the cavity of the mold geometry. The sub-domains C2f
= C2Xt) and ~2e = I2e(t ) are respectively the fluid domain and the empty space.
We define the characteristic function of fluid domain ~lf2fby 9

~l~f(x't)=f 1 i f x ~ ~2f
0 if x ~ f2 e (48)

As a consequence of the mass conservation applied to the fluid material, and


whatever the density of the empty part, the fluid movement is described by the
following transport equation 9
d
(x, t) = 0, V x ~ ~2
(49)

Figure 7-
3D finite element simulation of the filling of a grid cavity
911

The whole cavity is meshed. At each time step, the velocity and pressure
fields are calculated everywhere using the 3D Stokes solver previously
described. The empty part of the mold is represented by a zero velocity field
and a zero pressure field. The interpolated fluid position (one constant per
element) evolution is performed at element level by using a transport equation
finite element solver, based on an explicit Taylor discontinuous Galerkin
method. The basic idea of this scheme is that the successive time derivatives in
the Taylor development of the characteristic function 1 can be calculated
recursively by"

at ll n f = - V l o f .u
032 O
03/2 t[nf "-- V ~-~ a n f .U

69 3 02
0t 3 1 h i =_ 7 - & _ f ~ a / . u ... (50)

We have only to enforce the successive derivatives to be in the same


approximate space, for instance the piecewise constant, the convective term
being treated by a discontinuous interpolation technique.

Figure 8 9 3D double jets of viscous polymer


912

Figure 7 shows different steps of the filling of a thick 3D grid mold [36]. It
has been run on a single middle size workstation. There are 9 827 nodes,
which means 39 308 degrees of freedom and 50 739 elements. We can clearly
see the evolution of the free surface and the formation of the weld lines. This
method is not restricted to the flow front movement in confined mold filling
problems. It can be used to compute accurately transient free surfaces as
shown in Figure 8 where two jets of polymer collide each other, showing that
large moving free boundaries can be well rendered by this approach.

2.7 Viscoelastic computation


The direct calculation of viscoelastic flow remains one of the difficult topics
in numerical analysis [see Chapter 3]. At that time, 3D computations are just
emerging. For that reason, we restrict this presentation to the 2D numerical
treatment of differential models, as the Oldroyd-B or the Phan Tien - Tanner
model. The viscoelastic flow leads to a three-fields problem where the
velocity, the pressure and the stress tensor are solutions of coupled equations.
Let us consider the following problem derived from the Oldroyd-B model :
Find u, p, s solutions of"
V.[2r/s e (u) + s] - V p ' = 0 (51.1)

V.u = 0 (51.2)

s + ~ S~--20 e (u) (51.3)

with boundary conditions 9u - u 0 on o ~ ; s = s o on c~'2i (inlet boundary).


= 0 leads to a classical Stokes problem. In that case the three-fields
formulation must be equivalent to the two-fields velocity/pressure formulation.
For this reason, the unknown extra-stress tensor s belongs to the functional
space S defined by:

S = {s E ( L 2 ( ~ ) ) nxn} (52)

The space 5 must contain the rate of strain space, denoted e (V). The finite
element approximate space 5h must be chosen in order to be compatible with
V h and p h previously introduced. An easy way to check these compatibility

condition is to ensure that 5 h contains the discrete space e (Vh), for example
by choosing S h = k( V h).
The discontinuous Galerkin method is a consequence of the above
construction. Indeed, if the velocity discrete space is based on C Oelements, the
913

gradient of the discrete velocity is then discontinuous. Consequently, the


simplest discrete space containing the strain rate tensor is made of piecewise
polynomial interpolation. In [13], the Crouzeix-Raviart element for the
velocity/pressure discretization is associated with a second order interpolation
of the extra-stress :

W h = {u h E (C~ 2, u h l ~ e E (P2+(f2e))2}

p h _ {ph E (C-1(~2)), ph I~e ~ p l ( ~ e ) } (53.2)

S h - {S h e (C-l(~)) 2x2, s h I f~e E (p2(~e))2x2} (53.3)

In this case, the standard notation p2+ means that the velocity is enriched
with a bubble term. The bubble being made of a polynomial of degree three,
the gradient of the velocity is of degree two. The variational problem can be
written as :
Find (u h, ph, s h ) E V h x p h x s h , u h = Uo on o3~, u 0 being given, such as:

Sf~ 2rls ~(uh)" ~(uh*)df2 - S~P V'uh*d~ + fE~ sh ~(uh*)dE2 = 0 (54.1)

Sf~ ph V.u h d~ - 0 (54.2)

2uh.ne[sh]. .Ch do3~ _ f f2 2rls ~uh)"chd~2


e

V (u h,ph) 6 V h x Ph (54.3)

Figure 9" Comparison between calculated (top) and experimental (bottom)


birefringence patterns in an abrupt contraction
914

where [s h] is the jump of s between two successive elements, n e is the normal


to the element side and o~- is the reentrant side.
Figure 9 compares, in an abrupt contraction geometry, the computed
principal stress difference pattern to the experimental one, deduced from flow
birefringence experiments [9]. The agreement is fair and allows to foresee the
development of inverse methods for the characterization of molten polymer
rheological behavior in non-trivial flow configuration, starting from flow
birefringence patterns.

2.8 Conclusions
Despite interesting and promising results, the use of direct 3D numerical
simulation remains today generally limited to purely viscous fluids and
requires high numerical resources. To solve industrial problems encountered
in polymer processes, the use of approximation methods remains very often
necessary. These methods, which will be presented in the next sections, require
less computer resources and corresponding software packages are now well
implanted in polymer processing industry. They may be roughly divided into
two categories 9
- the confined flow approximation, which is adapted to flows into narrow
channels or between nearly parallel plates, dominated by shear;
- the thin film (or fiber) approximation, devoted to free surfaces, for
which elongational flows are dominant.

3. CONFINED FLOW APPROXIMATION

3.1 Kinematics approximation


The confined flow approximation is also well known as the lubrication
approximation [37-39], first applied to 2D flow simplification. It may be
applied to flow situations were one dimension is small compared to the others.
Let us consider for sake of clarity the schema of Figure 10, presenting the
flow of a polymer between two nearly parallel plates, defined by the local
distance h(x, y).
The following assumptions are made :
9 the flow is steady state, laminar and isothermal;
~?h ~?h
9 the distance h(x, y) varies slowly in x and y directions 9 ~xx << 1, ~yy << 1 ;
9 the curvature of the surface is negligible compared to the local thickness.
If these assumptions are verified, one can consider the flow as locally
between two parallel plates, separated by the distance h(x, y). Thus, the full 3D
velocity field can be simplified into a local 2D field :

u = (u(x, y, z), v(x, y, z), w(x, y, z)) u' = (u(z), v(z), O)


915

\ \ uT~ "%u t.k.j \\


iX ,.. \\

Xl ~ . . . . . . -

Figure 10 : Typical example of confined flow geometry

It means that elongational components are neglected compared to the


gapwise shear components and that hydrostatic pressure is constant throughout
the thickness. The local continuity equation (equation (1)) is replaced by its
integrated form :

3w
+ ~y + ~ z dz = o (55)

or" -~x u dz + -~y v dz + [w] =0 (56)

Assuming no slip conditions on upper and lower walls, equation (56) may
be written as"

~q~
Ox + ~Oy = 0 (57~

where qx and qy are the local flow rates per unit width in x and y directions,
defined by :

qx = u dz, qy = v dz (58, 59)


916

Remark 9 the use of a viscoelastic behavior in confined flow geometries does


not alter greatly the velocity, nor the stress fields. Thus, we will just consider
viscous behaviors in the following.

3.2 Restriction to isothermal Newtonian or power-law behavior 9


the Hele-Shaw equation
If we consider the confined flow of a Newtonian fluid of viscosity 7/in the
geometry of Figure 10, assuming the simplified velocity field, the Stokes
equations reduce to"

_~ O~2U 0p a21:
0X "- ~ 0Z2 ' Oy = 77 OZ2 (60, 61)

Integrating twice and using appropriate boundary conditions (no slip on


upper and lower walls) lead to the expressions of the local flow rates per unit
width:

1 ae 1 aph3 (62, 63)


qx = - 12 0 cgx h3, qY = - 12 0 0 y

Introducing equations (62) and (63) in equation (57) leads to the well
known Hele-Shaw equation :

a
~xx [o3x h3] + [0y ] = 0 or 7.(h 3 V p ) = 0 (64)

This equation can be generalized for non-Newtonian fluids under the form :

V.(/~ V p ) = 0 (65)

where/3 may have different expressions depending on the selected viscous law.
For example, for a power-law fluid, we obtain [40-41] :

+ 1)/n (1-n)/n
/ 3 = 2 n + 1( ) Vp (66)

3.3 Restriction to 2D flow : the Reynolds equation


If we consider the same type of flow, but in a 2D situation (calender gap or
journal bearing, for example), the integration of Stokes equation leads to 9

dp 12r/q
dx = - h(x)3 (67)
917

where q is the flow rate per unit width in x-direction. For a power-law fluid, a
similar expression is obtained"

dx = 2n q [h~) (68)

If the change in gap h(x) is known, the above expression allows to calculate
the pressure evolution along the flow direction and the pressure drop/flow rate
relationship. For example, for a dihedron with initial gap ho, final gap hi,
length L, and a linear variation of h between ho and hi, the pressure drop Ap
is expressed as :

ho+ hi
Ap - 6 71 q L ho 2 h l 2 (69)

3.4 Heat transfer approximation


The general temperature field T(x, y, z) corresponding to the flow
presented in Figure l0 is defined by the heat transfer equation (see equations
(41)). In a 2D flow where the temperature gradient across the flow thickness h
is limited (low shear rate, wall temperature close to polymer temperature), it
is possible to define a bulk temperature by :

T(x) = 1 u(y) T(x, y ) d y (70)


h(x) Y

where ~ is the mean velocity in x-direction. Using this expression, the energy
equation can be written under the form [13] 9

_ dT 1 ~2k Nu Tw- T (71)


u h(x) ~ + pep = pCp h(x)

where Tw is the wall temperature and Nu the Nusselt number, which


characterizes the heat transfer between the polymer and the wall 9

h(x)
Nu - hy k - (72)
918

where hr is the heat transfer coefficient previously introduced in section 1.5.


Equation (71) can be easily generalized to the 3D confined flow geometry
presented in Figure 10 under the form :

OT 1 Op c?T 10p 2k Tw - T
-~h (-O-x-x + ~pcp Ox) + ~h (-~-y-y + ~pcp Oy) = pcp
~ Nu h (73)

It is clear that these simplifications will give pertinent results only if the
heat transfer is correctly estimated and the transverse temperature gradient is
limited. They are generally useful for computing the flows in extrusion dies or
in calender gaps, but inappropriate to the strong conditions encountered in
injection.

3.5 First example : coat-hanger flat die geometry


Coat-hanger flat dies are very often used for the production of films and
sheets until a few meters width. In this geometry, the molten polymer is
distributed through a coat-hanger, which design is supposed to ensure a perfect
homogeneity of the flow rate at the die exit [42]. The flow in such a geometry
may be described using the confined flow approximation. For a power-law
fluid depending on temperature through an Arrhenius law, the set of equations
to solve is :
2n+l Oh Op h OK Op
h Ap+~ n ~x
+ ~?y ~ - -~n Ox+c)y

+ 2 OP 0P 002x~y
1-n Ox2 OxOy +
+ h = 0 (74)

- 1 2n (1)l/n(2)(2n+l)/n[lo~x~ ( O~y~ l (1-n)/'2n tT~ox (75)


u(x,y) = - h 2n+l

v_ (x,y) - - h12.2n+ 1 1 l, l 12.+l I( y + Oy (76)

2k Nu Tw- T (77)
u h (-ff~-x + ~pcp )+ v h ( + ~pcp ) = pcp h
919

E 1 1
K- Ko exp ~ ( ~ - ~-~o) (78)

This set of equations may be solved iteratively, using for example a finite
volume method [43-45].
Figure 11 presents an example of computed result, corresponding to the
extrusion of a polymer with a power-law index equal to 0.35. Flow lines and
isobars are drawn on Figure 1 l a. The pressure drop is mainly consumed in the
straining bar, located after the coat-hanger. The polymer flows preferably
along the coat-hanger and is progressively distributed throughout the lips. The
homogeneity of the exit flow rate may be evaluated by plotting the ratio of
local flow rate qx to mean flow rate q as a function of the position (Figure
lib). In the present case, the distribution is correct (+ 10 %), but it is
important to point out the strong dependence on the power-law index.
The temperature plays an important role in the flat die extrusion. Isovalues
of bulk temperature for an initial temperature of 200 ~ and for a wall
temperature of 200 ~ are shown in Figure 12a. Figure 12b indicates that the
exit flow rate distribution is very sensitive to the wall temperature. It proves
that it is possible to correct a flow rate heterogeneity just by changing the
regulation temperature.

100 bars
90.1bars . 70.3 bars

5o.s~
30.7"i ]
.zo.81"---T
10.9J--- 4
Po=l.o [ . . . .
bar

t 1.2
C~
~" 1.1------, ' i
d
a) n-0.2
t.
1,0 b) n:= 0.35
o c) n=0.5
~- 0.9
,~_.
C
~' 0.8
A B

Figure 11 9 Flow lines and isobars (a) and flow rate exit distribution (b) for
the flow in the coat-hanger fiat die
920

200.0 *C

t
"M~L

i 202.9

1 1.4
I -
203.3

,-~
1.3
I IB ;,
l
O
1.2
AF sS
Wail femperafure:
O a) T : 180 ~
c~ bl T = 200 ~
. 1.1 c) T : 220 ~
i /

1.0
Q,P
bx,......t , J/
"- 09
~u

0.8 a%,,"
/
s
..,,.,,"

O.7
A

Figure 12" Isovalues of bulk temperature (a) and exit flow rate distribution
(b) for the flow in the coat-hanger flat die

3.7 Second example : mold filling of a box geometry


For a large number of molded components, the confined flow
approximation can be used. In injection molding, gapwise temperature
gradients are very important, so that, contrary to extrusion, an average
temperature cannot be used. Moreover, the rate of strain varies from zero to
very large values (several thousand of s-l), and precise viscous model such as
the Carreau equation must be used. Therefore, equations (60) and (61) are
integrated twice without specifying a form for the viscosity 77, and /~
appearing in equation (65) is defined as :

'z (z - z*)
13- ~ d z (79)

z* is the value of z for which the velocity is maximum. It is equal to h/2 with a
symmetric thermal regulation. In more general situations, it is defined by 9
921

~Z-Z*
o dz - o (8o)

The boundary conditions are 9


9 on the lateral walls, the velocity along n, normal to the wall, is set to zero,
which means that 3p/On - O. The velocity component tangent to this lateral
wall is not equal to zero. This is one of the limitation of Hele-Shaw models ;
9 the flow rate is given at the entrance, that is to say the pressure gradient is
imposed;
9 at the flow front, the pressure is equal to zero, which means that the air is
supposed to escape the cavity freely.
When the cavity is not in a plane, the same kind of resolution is carried out
in a tangent plane.
As defined in paragraph 1.4, the energy equation is"

pcp U
OT
+ V --~ + W
c)__~zI c) 2T 9
= k t~Z2 + W (81)

Due to high transverse temperature gradient, the term w cgT/cgz cannot be


neglected a priori, even if w is small in the confined flow approximation, w is
not explicitly taken into account, but it can be evaluated by considering the
local continuity equation. Boundary conditions are indicated in paragraph
1.5.2.
The Hele-Shaw equation can be solved at each time step by finite difference
[46-48] or finite element methods [49-52]. The position of the flow front can
be evaluated with a moving mesh method [51], but a more simple technique
consists in using control volumes built with the finite element description of
the mid-surface of the cavity (Figure 13a).

Figure 13 9 Triangular mesh of the mid-surface and control volumes (a),


and extension in the thickness within an element (b)
922

A fill factor a is associated to each control volume. If a = 0, the volume is


empty. If a = 1, the control volume is full. If 0 < a < 1, the control volume is
currently being filled, and the flow front is somewhere inside. The evolution
of ct for each control volume ~ is obtained by'solving equation (82), which is
deduced from the continuity equation for an incompressible material 9

~9
& Jt~k h a d s =-kf~a h u.n ot dl (82)

The iterative resolution is usually the following 9


9 The pressure is linearly discretized on the finite elements. Equation (82) is
written on each control volume, u being expressed as a function of the
pressure gradient. A linear system with the nodal values of the pressure is
thus obtained.
9 The temperature field is obtained by writing heat flux balance in control sub-
volumes (Figure 13b), and a temperature for each layer and each control
volume is obtained.
9 Equation (82) permits to update the value of ct, and to obtain the new value
of the flow front position.

Figure 14 9 Isovalues of the pressure at different time steps during the


filling of a box shaped cavity (a-d) and fill factor at time 0.09 s
923

Figure 14 shows the isovalues of the pressure and the fill factor ~ for the
filling of a box cavity. This kind of computation gives accurate results for the
pressure, the temperature, and the flow front. The location of weld lines is
usually correct. The equations can be generalized to compressible materials,
for the packing phase, or for a better resolution of the filling of large cavities
at high pressure.

3.7 Comparison between direct and approximated solutions


3. 7.1 Flow of a Newtonian fluid in a dihedron
In order to check the validity of the lubrication approximations, we can
consider for example the flow of a Newtonian fluid in a dihedron. An exact
solution can be obtained using cylindrical coordinates [13]. The relationship
between flow rate and pressure drop is 9

W Ap (si220 - 0 cos20) Ro 2 R12


Qexact = 17 (Ro2 _ R12) (83)

where 0 is the dihedron angle and Ro and R1 the radius defining the entry and
exit, respectively. The approximated solution using the lubrication
approximation is given by equation (68). So, the ratio of flow rate between
exact and approximate solutions is only a function of the angle 0"

Qapprox 4 tg30
(84)
Qexact = 3 sin2 0
- 0 cos20
2

1,3

t,,.}
m
X
1,2
o
x
0 C,one
~ 1,1
0 9 dron

1,0 ~ " I " I " I ' '

0 5 10 15 20
Angle (degree)

Figure 15 9 Comparison between approximate and exact solution for the


flow in a cone and in a dihedron
924

Figure 15 presents the evolution of the flow rate ratio as a function of the
angle. It can be seen that the error made in using the lubrication
approximation remains less than 5 % for an angle inferior to 10 ~ For an
axysymmetrical conical geometry, this limit is roughly equivalent but further
deviation is enhanced. These particular examples are representative of the
global validity of the confined flow approximation. Whatever the real
geometry and the polymer behavior, the upper limit of the relative angle
between the walls must not exceed 10 ~ to ensure a correct description of the
flow [53-54].

3.7.2 Particle tracking in injection molding


We have seen in section 3.6 that the Hele-Shaw equation is now widely used
to predict pressure, temperature fields, and flow front progression during
cavity filling. Nevertheless, a direct resolution (see section 2) is the only
technique which can be used when the geometry is complex. But even with
simple cavities, more precise kinematics description can be necessary to
predict accurately the flow, for example at the junction between the runner,
the gate and the cavity, or at the flow front where the fountain flow takes
place. Moreover, crystallinity ratio for semi-crystalline thermoplastics, cross-
linking for thermosetting polymers or rubber compounds, reinforcing
particles displacement and orientation need a precise knowledge of the thermal
and kinematics history of each material point starting at the mold entrance.
Particle tracking at the flow front has been specially studied. The pivot point
method has been introduced by Manas-Zloczower et al. [55] and Dupret and
Vanderschuren [56]. Heat diffusion and viscous heating in the frontal zone are
neglected. The neutral line is defined as the gapwise position for which the
actual velocity is equal to the mean velocity. A particle "below" the neutral
line (that is between the mid-plane and the neutral line) will reach the flow
front and jump instantaneously "above" the neutral line (between the neutral
line and the mold wall). The position is determined by writing a local mass
balance. A more precise description can be obtained in a reference frame
linked to the flow front in isothermal conditions. Then, a steady direct two
dimensional resolution is possible, either analytically [57] or with a finite
element method [58]. Finally, the direct resolution in the laboratory reference
frame described in section 2 can be used in isothermal and non-isothermal
conditions. Figure 16 shows the evolution of a line of tracers obtained with the
three methods [32, 59]. Direct 2D models permit to obtain the V-shape which
has been experimentally observed by Schmidt [60]. After a certain time, when
the tracer is stretched, the three approaches give similar results.
925

~ 0,8 F l o w front position 98

0,6

~.,
,-, 0,4 ----------- 2,~r~
D #,, F i n i t e e. l e m e n t
~ 2.D ~ A n a _ l y t i c a l - " ~ . .
0,2 ......... Pivot "kN
0 , , , , i , , , , i , ~ , , i , , , ,.i , , ,. , , i , , , ~ ' ' ' ' I ~ ' ' I

0 l 2 3 4 5 6 7 8
Flow length

~' 0,8 F l o w front position 9 2


0a

0,6
~

0,4
....... ~D # A n a l y t i c a l "'.~
l~ 0,2 ......... Pivot I",
0 , , ~ i , , , i , , , i ,_i , i , , , i , , ~I i ,. , , i , ,..

0 2 4 6 8 10 12 14
Flow length

1 ....

N 0,2
~
0,6

0,4 1 ~
. . . . .
.........
2D #
2D # A n a l y t i c a l
Pivot
Finite element

F l o w front position 9 16
"-.,,.

""-

" 1 0
0
' ' ' ' I

5
' ' ' '

10
I ' ' ' '

15
I I I I
....

Flow length

] ~
0, 8 ~ ..........
_ ..-......., -
, \

0,6

0,4 2D Finite e l e m e n t
0,2 . . . . . 2D A n a l y t i c a l F l o w front position 9 40
......... Pivot
0 ' ' .... ' I ' ' ' I ' ' ' I ' ' ' I ' ' ' I ' ' '

0 8 16 24 32 40 48
Flow length

Figure 16 9 Comparison of three methods for tracking a line of particles at


the flow front

0 THIN FILM (OR FIBER) APPROXIMATION : EXAMPLE


OF THE CAST FILM PROCESS

Confined flows considered in the previous section are mainly shear flows.
As the fluid is sticking at the wall, the flow is locally a plane Poiseuille flow
926

and this is the key point of the confined flow approximation. On the contrary,
for stretching flows, the normal stress is zero at the free surface and hence the
velocity is quite constant throughout the thickness. In the following, we
describe one and two dimensional kinematics and mechanical approximations
for the cast film process, considered as an example of free surface stretching
flow. The same basic ideas can be developed for melt spinning and blown film
processes.
In the cast film process, the molten polymer is stretched between the flat die
and the chill roll (Figure 17). The stretching process may be considered as
isothermal, and cooling is achieved very rapidly on the chill roll. A shrinkage
(neck-in) is observed in the transverse direction. The thickness is more
important on the lateral edge of the film (dog bone defect), so that it is
necessary to cut the edges after solidification. Challenge of modeling is to
predict width and thickness of the solidified film. We first describe a two-
dimensional model for the isothermal Newtonian cast-film. Coupled equations
for thickness and mean velocity in the thickness are derived from these
approximations. Then, a one-dimensional model is presented by introducing
additional approximations. This 1D model allows easy extension to viscoelastic
constitutive equations, non-isothermal conditions or multilayer film
processing. Finally, a stability analysis will be performed using different
techniques both for the 1D and 2D models.
Wo
A
v

exit

W(x)

Chill roll

Figure 17 9A schematic view of the cast film process

4.1 G e n e r a l m e c h a n i c a l a p p r o x i m a t i o n s for thin film 9 a 2D m o d e l


Figure 17 gives a schematic view of the cast film process. The molten
polymer is outgoing from the flat die at x = 0 and is solidified on the chill roll
927

at x = L. The mean surface of the film is, for simplicity, assumed to be the
plane z = 0 .

4.1.1 Kinematics
Let us note h(x,y) and W(x) the film thickness and width. The polymer
flows in the volume defined by :

1 1
{- -~ h(x,y) <_ z <- -~ h(x,y), 0 <_x <_L, -W(x) ~_y ~_ W(x) } (85)

We assume, as for the confined flow approximations, that the thickness


h(x,y) is slowly varying in x and y-directions 9

Oh Oh
<< 1,--~-<< 1 (86)
Ox oy

Let us note U ( x , y ) a n d V(x,y) the velocity components on the mean


surface:

u(x, y,z) = U(x, y) + o(z) (87)


v(x, y,z) = V(x, y) + o(z) (88)
where o(z) is a term of the order of magnitude of z.
As w(x,y,O) = 0 on the mean surface, we have :

~w
w(x,y,z) = z ~ (x,y,O) + o(z) (89)

Using the incompressibility condition, we have"

c)w c)u Ov OU 8V
c?z (x,y,O) =- -ffs (x,y,O)- -~y (x,y,O)=- -~x (x,y)- -~y (x,y) (90)

4.1.2 Mass conservation and rate of strain tensor


The continuity equation, naturally verified by the three dimensional
velocity field, is for stationary 2D free surface flows replaced by an integrated
form (this is to compare to equation (56) of section 3) :

a ~_~ a ~_~ ah ah hi2


a~ h/2
u dz +
~ ha
v dz + [ w - -~x u - -~yV ]-h/2 =0 (91)
928

Using free surface conditions on upper and lower boundaries and estimating
integrals by a mean value formula, the previous equation becomes 9

Ox (hU) + (hV) - 0 (92)

As the flow is mainly extensional, the rate of strain tensor is (neglecting


shearing in the thickness) :

9 ~

Exx Exy 0
1~ - g.xy Eyy 0 (93)

0 0 ezz

9 8U 9 9 I 8U OV 8V 9 8U 8V
with" 8xx = o3x ; gXy = 8yx = -~ ( - ~ +-~-x ) ; ~ y = ay ; eZZ = - a x - ay

4.1.3 Stress balance for a Newtonian constitutive equation


Assuming a Newtonian behavior, we have"

8U 8V
f i z z - - 20 (-~x + ffyy ) - P (94)

Using the zero stress condition normal to the free surface, and identifying
the normal to the free surface to z-direction, we have :

8U 8V
p = - 2r/(fiX-
x + -~-y) (95)

Gxx r 0
and hence" ff =
I I
axy ayy 0
0 0 0

8U 8U 8V
(96)

with" axx = 271 -~x + 271 (ff-ff + -~v ) ;


i 1

8V 8U 8V
Gyy- 2~ Uy + 2~ (gx + 7y~ ;
8U 8V
~ y = 77 ( -~y + ~ ).
929

The stress tensor can be considered as a 2D tensor. Integrating the


equilibrium equation (gravity forces are neglected) through the thickness and
using the free surface condition, we obtain"

V.(hff) = 0 (97)

4.1.3 Free surface and boundary conditions


Equations (92), (96), (97) have to be completed with appropriate boundary
conditions on the edge of the film of normal n 9

U nx + V ny = 0 (98)

(Txx nx + ffxy ny = Gxy nx + O'yy ny = 0 (99)

c)W o3W 2 -1/2 0 W 2 - 1/2


with" nx = - Ox (1 + (-~x ) ) ; ny = (1 + (--~x ) )

We add classical boundary conditions at the die exit and on the chill roll (at
distance x = L):

U(O, y ) = UO; U(L, y) = UL ; W(O) = WO ; h(O) = h 0 ( 00)

UL
The draw ratio is defined as Dr =
u0
4.1.4 Results o f the two dimensional model
For a given value of the thickness distribution h ( x , y ) , the velocity
components U and V and the width W are solutions of a free surface elliptic
problem (equation (97)), associated with boundary conditions (98), (99) and
(100). For a given velocity field, thickness h is solution of a transport equation
(92). This coupled system of equations can be solved iteratively by using finite
element solvers for velocity and thickness, and a fixed point method to reach
convergence [61, 62].
The preceding analysis may be easily extended to a viscoelastic Maxwell
equation. The equations and the numerical resolution are detailed in [62]
(another viscoelastic 2D model has been recently proposed by Debbaut et al.
[63]). Figure 18 presents the final thickness distribution of the film. Figure 19
shows the shape of the film between the flat die and the chill roll [62]. When
introducing a viscoelastic equation, the neck-in phenomenon is less pronounced
and the dog bone defect remains located at the periphery of the film.
930

0,15
Newtonian
gl
(1)
e-
0,10
c" c-

O r

0,05
rr' Viscoelastic i.

0,00 9 I ' I ' I ' I " !

0,0 0,1 0,2 0,3 0,4 0,5 0,6


y/W o

Figure 18 9 Final film section for a Newtonian and for a Maxwell fluid

1 ~0 ' '

0,8

..~ 0,6
i

0,4

0,2

0,0
0,7 0,8 0,9 1,0
Reduced half-width (W(x)/W O)

Figure 19 9 Half-width of the film for a Newtonian fluid and for a


Maxwell fluid

4.2 A d d i t i o n a l
mechanical approximations 9 a ID model
A one-dimensional model may be obtained if the thickness and velocity are
assumed to be varying with x only. In this case the model predicts neck-in and
mean thickness of the solidified film.
931

4.2.1 Mass conservation


Let us assume that h= h(x) and U = U(x). Equation (92) becomes 9

0
~x (hWU) = 0 (101)

4.2.2 Equilibrium equation


The stretching force F is defined by 9

F = h(x) W(x) tYxx (102)

As we are neglecting mass and inertia forces, the balance equation (97)
becomes 9

OF
Ox = 0 (~03)

and hence F is a constant. The differential system of equations obtained by


adding these kinematics hypothesis can be solved for Newtonian or viscoelastic
fluids by using a shooting method on the value of the force F. The
dimensionless parameter defining the aspect ratio of the film (A - W/W0)
appears in this set of equations. Figure 20 compares the shape of the film with
2D and 1D Newtonian models.

1,0

0,8

_J 0,6
!

"" 0,4
1D model
0,2

0,0 9 I 9 I 9

0,4 0,6 0,8 1,0


Reduced half-width (W(x)/W 0 )

Figure 20 9 Shape of the film surface for Newtonian 1D and 2D models


932

For that particular free surface flow, the kinematics assumption (1D or 2D)
and the constitutive equation (Newtonian or Maxwell) have a marked influence
on the shape of the film between the die and the chill roll and on the final
dimension of the film (thickness and width).

4.3 The stability of the cast film process


4.3.1 The draw resonance instability
Draw resonance instability appears at large value of the draw ratio for
isothermal cast film and melt spinning processes. If the draw ratio is greater
than a critical one Drc, a periodic flow is observed. This critical draw ratio is
close to 20 for isothermal melt spinning [64]. Onset of draw resonance
instability for cast film induces periodical variations of both thickness and
width [65].

4.3.2 Computation of the critical draw ratio


The stability of the one-dimensional solution is studied by using a linear
stability method, identical to those developed for fiber spinning [66, 67]. It
means that the time dependent set of equations is linearized and eigenvalues are
computed. For a Newtonian fluid, the critical draw ratio is a function of the
aspect ratio A (Figure 21). Increasing this aspect ratio leads to stabilize the
process.
Direct time dependent simulation is used to study the stability of the 2D
solution. Figure 21 points out that the shape of the stability curve is not too
different between the 1D and the 2D models. The 1D model is able to capture
the influence of the film aspect ratio on the stability.

1,3

< 1,1 STABLE


~
o

x__
0,9
to
0,7
o0
t~
E 0,5
.m
ii

0,3
/ UNSTABLE,
0,1 9 I I I I I

15 20 25 30 35 40 45
Draw ratio

Figure 21 9 Cast film stability curve 9 influence of the film aspect ratio.
933

5. C O N C L U S I O N

As shown in this chapter, it is nowadays possible to compute the flow of


molten polymers in most of the complex geometries encountered in polymer
forming processes, in both stationary (extrusion) and unstationary (injection
molding - blow molding) conditions. This requires first precise volume
meshing methods, starting for example from a CAD surface meshing of the
tools. This necessitates also robust numerical finite element methods 9 mixed
velocity/pressure Galerkin method for the mechanical problem, Taylor-
Galerkin method for the thermal problem, discontinuous Galerkin method for
convection problems (viscoelasticity, for example). Iterative solvers and
parallel computing reduce storage requirement and computation time, which
allows to perform numerical simulations in complex industrial geometries,
with refined meshing. These direct numerical methods do not necessitate any
geometrical or kinematics assumptions, but they remain limited, at the present
time, to purely viscous constitutive equations (in 3D). In addition, computation
time is important and die or mold optimization, which necessitates a lot of
successive numerical simulations, remains a difficult task. In that sense,
approximation methods, which require however sophisticated mechanical and
thermal treatments, based on kinematics, heat transfer and geometry
assumptions, are still useful.
In confined flow geometries, approximations consist generally in neglecting
elongational components compared to the shear components throughout the
thickness. As a consequence, viscoelasticity is a second order phenomenon in
these shear dominant flows. The so-called Hele-Shaw approximations result in
a single mechanical differential equation with the pressure as unknown. For
example, most of the commercial injection molding software packages are
based on this kind of approximation.
In free surface stretching flows, on the contrary, shear components
throughout the thickness are neglected compared to elongational terms. The
so-called thin film approximation leads to differential equations with only the
mean velocity components throughout the thickness as unknowns. Stationary,
as well as time dependent solutions, may be obtained. In these elongational
dominant flows, viscoelasticity has a first order influence.
One cannot conclude at that time that polymer processing modeling is now
fully completed. Several problems remain open : thermomechanical coupling
with severe temperature gradients, 3D viscoelastic computations, adaptative
meshing based on error estimation ... But the methods we presented can now
be considered as practical tools to calculate flow rate and temperature
distributions in polymer processing devices, to optimize processing geometries
and, soon, to predict end-use properties of the produced polymer parts. This
will require to incorporate new equations (crystallisation kinetics depending on
temperature, temperature gradient and stress field, macromolecule or fiber
934

orientation...) and so to investigate in details the corresponding new scientific


fields.

NOMENCLATURE

cp heat capacity
-/
C t Finger tensor
C matrix
Dr draw ratio
e mass density of internal energy
E activation energy of the viscosity
F,F force
g gravity
h flow gap or thickness, damping function, mesh size
hT heat transfer coefficient
I identity tensor
11,12 invariants of the Finger tensor
k, km heat conductivity
K,K 0 consistency
l distance
L length
m memory function
n power law index
n normal vector
Nu Nusselt number
P pressure
p' isotropic part of the stress tensor
P pressure space
q heat flux
q, qx, qy flow rates per unit width
R ideal gas constant
Ro, R 1 radius of a dihedron
Re Reynolds number
S extra stress tensor
T temperature
T* test function
Te, Tw controlled temperature, wall temperature
U, V, W velocity components
U velocity vector
U* test function
935
m

U,V mean velocity


U,V velocities
V velocity space
W width
f~ viscous dissipation
IIa characteristic function of the fluid domain

o~ fill factor
".L.
second invariant of the rate of strain tensor
acceleration
e rate of strain tensor
77, 7/0 shear viscosity
F/e elongational viscosity
r/s Newtonian viscosity
A, relaxation time
P density
o stress tensor
0 angle of dihedron
f~ domain
~-'~e element
~e empty part of the cavity
fluid domain
boundary of the domain

V~ divergence
V gradient
tr trace
d
material derivative
dt
6
& upper convective derivative
0
partial derivative
Ot

Hl(~'2) = {u E L2(~')), ]Vu I EL2(~)}


L2(~) - {u, I [u 12d~ < oo}
II. IIl,a = ( I lu 12da + I lV. 12da)l/2
936

REFERENCES

~ P.J. Carreau, Trans. Soc. Rheol., 16 (1972) 99


2. K.Y. Yasuda, R.C. Armstrong, R.E. Cohen, Rheol. Acta, 20 (1981) 163
3. P.E. Rouse, J. Chem. Phys., 21, 7 (1953) 1272
4. B. Zimm, J. Chem. Phys., 24, 2 (1956) 269
5. J.G. Oldroyd, Proc. Roy. Soc. London, A345 (1958) 278
6. R.G. Larson, Constitutive Equations for Polymer Melts and Solutions,
Butterworths Eds., Stronham (1982)
, N. Phan-Thien, R.I. Tanner, J. Non Newt. Fluid Mech., 2(1977) 353
8. H. Giesekus, J. Non Newt. Fluid Mech., 11 (1982) 69
9. C. B6raudo, A. Fortin, T. Coupez, Y. Demay, B. Vergnes, J.F.
Agassant, J. Non Newt. Fluid Mech., to appear
10. A.S. Lodge, Elastic Liquids, Academic Press, New York (1964)
ll. M.H. Wagner, Rheol. Acta, 15 (1976) 136
12. A.C. Papanastasiou, L.E. Scriven, C. Macosko, J. Rheol., 27 (1983) 387
13. J.F. Agassant, P. Avenas, J.P. Sergent, B. Vergnes, M. Vincent, La Mise
en Forme des Mati~res Plastiques, Lavoisier, Paris (1996)
14. G.W.M. Peters, F.P.L. Baaijens, J. Non Newt. Fluid Mech., 68 (1997)
205
15. V. Girault, P.A. Raviart, Finite Element Approximation of the Navier-
Stokes Equations, Theory and Algorithm, Springer-Verlag, Berlin
(1986)
16. F. Brezzi, M. Fortin, Mixed and Hybrid Finite Element Methods,
Springer-Verlag ( 1991)
17. T.R.J. Hughes, W.K. Liu, A. Brooks, J. Comp. Phys., 30 (1979) 1
18. M. Fortin, R. Glowinski, Augmented Lagrangian Methods, Dunod
(1983)
19. M. Fortin, A. Fortin, Int. J. Num. Meth. Fluids, 5 (1985) 911
20. M.P. Robichaud, P.A. Tanguy, Comm. In Appl. Num. Meth., 3 (1987)
223
21. D.N Arnold, F. Brezzi, M. Fortin, Calcolo, 21 (1984) 337
22. T. Coupez, in" Numerical Grid Generation in Computational Fluid
Dynamics and Related Fields, N.P. Weatherill et al. eds., Pineridge Press
(1994)
23. J.F. Gobeau, T. Coupez, B. Vergnes, J.F. Agassant, in : Numerical
Methods in Industrials Forming Processes, Paul R. Dawson, Shan-Fu
Shen eds., A.A. Balkema (1995)
24. M. Fortin, Int. J. Num. Meth. Fluids, 1 (1981) 347
25. R. Pierre, SIAM J. Numer. Anal., 32 (1995) 1210
26. L.P. Franca, S.L. Frey, T.J.R. Hughes, Comput. Meth. Appl. Mech.
Eng., 99 (1992) 209
27. T. Coupez, S. Marie, Int. J. S. A., to appear
937

28. R. Pierre, Comput. Meth. Appl. Mech. Eng., 68 (1988) 205


29. A. Wathen, D. Silvester, SIAM J. Numer. Anal., 30 (1993) 630
30. Y. Saad, M. Schultz, SIAM J. Scient. Stat. Comput., 7 (1986) 856
31. S.F. Ashby, T.A. Manteufel, P.E. Saylor, SIAM J. Numer. Anal., 27
(1990) 1542
32. B. Magnin, Doctoral Dissertation, Ecole des Mines de Paris (1994)
33. A.N. Brooks, T.J.R. Hughes, Comp.Meth. in Appl. Mech. Eng., 32
(1982) 199
34. O. Pironneau, M6thodes des E16ments Finis pour les Fluides, Masson,
(1993)
35. J.-F. Gobeau, Doctoral Dissertation, Ecole des Mines de Paris (1996)
36. E. Pichelin, Doctoral Dissertation, Ecole des Mines de Paris (1998)
37. J.R.A. Pearson, Mechanics of Polymer Processing, Elsevier, Londres
(1985)
38. V.L. Streeter, Handbook of Fluid Dynamics, Mc Graw Hill, New York
(1961)
39. Z. Tadmor et C.G. Gogos, Principles of Polymer Processing, Wiley,
New York (1979)
40. B. Vergnes, Doctoral Dissertation, Ecole des Mines de Paris (1979)
41. R.T. Fenner, Principles of Polymer Processing, Mc Millan, London
(1979)
42. W. Michaeli, Extrusion Dies, Carl Hanser Verlag, Munich (1979)
43. B. Vergnes, P. Saillard, B. Plantamura, Kunststoffe, 11 (1980), 752
44. B. Vergnes, P. Saillard, J.F. Agassant, Polym. Eng. Sci., 24 (1984), 980
45. B. Arpin, P.G. Lafleur, B. Vergnes, Polym. Eng. Sci., 32 (1992), 206
46. E. Broyer, C. Gutfinger, Z. Tadmor, Trans. Soc. Rheol., 19 (1975) 423
47. C.A. Hieber, S.F. Shen, Israel J. Technology, 16 (1978) 248
48. S. Subbiah, D.L. Trafford, S.I. Gticeri, Int. J. Heat Mass Transfert, 32
(1989) 415
49. G. Boshouwers, J. Van der Werf, PhD Thesis, Technical University
Eindhoven, The Netherlands (1988)
50. H.H. Chiang, C.A. Hieber, K.K. Wang, Polym. Eng. Sci., 31 (1991) 116
51. A. Couniot, L. Dheur, F. Dupret, in" Numerical Methods in Industrial
Forming Processes, Balkema, Rotterdam (1989) 235
52. J.L. Willien, Doctoral Dissertation, Ecole des Mines de Paris (1992)
53. A.M. Benis, Chem. Eng. Sci., 22 (1967), 805
54. W.E. Langlois, Slow Viscous Flows, Mc Millan, London (1964)
55. I. Manas-Zloczower, J.W. Blake, C.W. Macosko, Polym. Eng. Sci., 27
(1987), 1229
56. F. Dupret, L. Vanderschuren, AIChE J., 34 (1988), 1959
57. J.M. Castro, J.W. Macosko, AIChE J., 28 (1982), 250
58. H. Mavridis, A.N. Hrymak, J. Vlachopoulos, Polym. Eng. Sci., 26
(1986), 449
938

59. S. Karam, Doctoral Dissertation, Ecole des Mines de Paris (1995)


60. L. Schmidt, Polym. Eng. Sci., 14 (1974), 797
61. S. D'Halewyn, J.-F. Agassant, Y. Demay, Polym. Eng. Sci., 30 (1990)
335
62. D. S ilagy, Doctoral Dissertation, Ecole des Mines de Paris (1996)
63. B. Debbaut, J.M. Marchal, M.J. Crochet, Z. Angew, Math. Phys., 46
(1995) 679
64. J.R.A. Pearson, M.A. Matovich, I & EC Fundamentals, 8 (1969) 605
65. Ph. Barq, J.-M. Haudin, J.-F. Agassant, Intern. Polym. Proc. 8 (1992)
334
66. D. Silgay, Y Demay, J.-F. Agassant, C.R. Acad. Sci. Paris, t.322, S6rie
lib (1996) 283
67. D. Silagy, Y Demay, J.-F. Agassant, Polym. Eng. Sci. 36 (1996) 2614
939

MODELLING AND SIMULATION OF INJECTION MOLDING

F. Dupret, A. Couniot 1, O. Mal, L. Vanderschuren g, O. Verhoyen

CESAME, Unitd de Mdcanique Appliqude, Universitd catholique de Louvain,


avenue G. Lemaftre 4-6, B-1348 Louvain-la-Neuve, Belgium
Tel :32 (0) 10 472350, E-mail : fd@mema.ucl.ac.be
1
current address : Siemens-Nixdorf lnformation Systems S.A., LoB "Major
Projects", chaussde de Charleroi 116, B-1060 Brussels, Belgium
2 current address : Shell Research S.A., avenue Jean Monnet 1, B-1348
Louvain-la-Neuve, Belgium

1. INTRODUCTION

The last decades have been marked by the spectacular development of synthetic
polymers, which are now commonly used in all of the major market sectors 9con-
sumer appliances, automotive, industrial, electric and medical. This sensational
success arises not only from the cost effectiveness of plastic parts, but also from
the wide range and amazing balance of chemical, mechanical and electrical prop-
erties of polymers. The growth of the polymer industry has been conducive to the
development of various processing techniques, among which injection molding
remains one of the most widely employed.
Injection molding is used for the mass production of identical parts obtained by
introducing a polymer melt into a cooled cavity. A major advantage of this pro-
cess is its suitability for the production, in a single operation, of ready-to-use arti-
cles of considerable geometrical complexity. As a result, its field of application is
very broad, ranging from the production of small parts such as tiny watch gears or
electronic components to large objects such as boats or automotive parts. Most
polymers may be injection molded, including amorphous and semi-crystalline
thermoplastics, thermosets, elastomers and fiber-reinforced plastics.
Injection molding is considered to be one of the most complex polymer proces-
sing techniques. A sketch of a typical molding machine is shown in Figure 1.
Primarily, two parts may be distinguished 9an injection unit comprising a hopper,
940

Figure 1. Schematic layout of a reciprocating screw injection molding machine,


and description of the main processing stages.

a rotating screw and a heated barrel, and a clamping unit containing the mold
typically made of two plates. The injection unit melts the polymer and injects it
into the cavity, whereas the clamping unit ensures the closure and opening of the
mold together with the final part ejection. In the case of thermoplastics, which
this chapter essentially addresses, the material is supplied under the form of
granules or powders introduced into the hopper. During plasticization, the screw
rotates and acts as a mixer and a pump : the material is taken from the hopper,
melted, homogenized and conveyed towards the barrel head while, at the same
941

time, the screw moves backwards. During injection, the screw acts as a piston,
moves forward and pushes the molten material into the mold. This operation
includes the filling stage, during which the polymer is injected into the mold, and
the packing stage, during which a high pressure is exerted by the screw to force
in additional melt to compensate for subsequent shrinkage due to cooling and
solidification. These steps are followed by the cooling stage during which the
material solidifies, and the ejection stage where the mold plates are separated,
thereby allowing the ejection of the part.
We shall not describe this process any further ~ the reader will find a thorough
description in [1,2]. Let us however mention three related developments which
will be addressed in this chapter. The first one is compression molding, where a
polymer charge, initially placed inside the cavity, is forced into the desired shape
by squeezing it via the motion of one half of the mold relative to the other [3,4].
Compression molding offers several advantages, in particular the use of simple
molds, the application of a low stress level during processing and a reduced mate-
rial waste. The second process is Reaction Injection Molding (RIM), where two
or more low viscosity liquids, which become reactive when brought together, are
mixed just before filling, the polymerization taking place partly during but essen-
tially after injection [5,6]. This process is typically applied for the production of
large objects due to the lower pressures required. The third class of related pro-
cesses is formed by Structural Reaction Injection Molding (SRIM) and Resin
Transfer Molding (RTM), which are both used for producing continuous fiber
composite parts by injecting a thermosetting resin into a mold filled with a fiber
preform [7,8].
In all these cases, the design of a new mold is a delicate and time consuming
task, often requiring dedicated trials and experiments. The aim of numerical
modelling is to prevent this costly and lengthy approach by providing an easier
way to apprehend the physics involved and to assess the influence of the key pro-
cess parameters. Quite clearly, the ultimate objective is, given a polymer, an ap-
plication and a desired type of geometry, to allow the numerical determination of
an optimum mold design together with the most appropriate operating conditions.
However, the numerical modelling of the filling and packing stages of the injec-
tion molding process (which represents the subject of this chapter) is plagued
with various difficulties. For the sake of simplicity, let us restrict ourselves here
to non-reinforced thermoplastics. During filling (Figure 2), the mold plates are
maintained at a temperature lower than that at which the injected material solidi-
fies. The problem is thus transient and 3D, in view of the front motion and the
important heat transfer between the melt and the cavity walls. Another difficulty
arises from the temperature dependence of the polymer viscosity, which couples
942

jjjjjjjj..~W~alJ~jjjjjjjj
Te
Jt e
Gate I ~ ~ ~
................................. 9 ........ i .......................
Front

Figure 2. Sketch of the injection molding problem.

the kinematics and the temperature distribution, especially close to the walls
where the lower temperatures increase the viscosity and slow down the fluid or
lead to the formation of a frozen layer. An additional difficulty relates to the
theology of polymer melts, which exhibit a strong non-Newtonian behavior as a
result of the very high shear rates prevailing during filling. Finally, the high
temperature gradients and pressures involved induce significant compressibility
effects, in particular during the packing stage.
A comprehensive numerical simulation of injection molding therefore appears to
be an outstandingly difficult task. Nonetheless, the geometry of molded parts is
typically characterized by a small thickness compared to the other characteristic
lengths or radii of curvature. As a consequence, Generalized Newtonian (GN)
stresses may be assumed and the goveming equations can be considerably simpli-
fied over most of the flow domain. The rheology of molded polymers is analyzed
in Section 2, while Section 3 details the flow and heat transfer model resulting
from the thin cavity assumption. The numerical algorithm accordingly developed
at the University of Louvain (UCL) [9-26] is presented in Section 4, and illustrat-
ed in Section 5 by means of simulation examples and experimental validations. In
these sections, only the injection and compression molding of thermoplastic mate-
rials are addressed. Extensions of the model to the cases of reactive resins, semi-
crystalline polymers and the filling of fiber mats are considered in Section 6. The
molding of fiber-reinforced polymers is investigated in the companion chapter
"Suspensions : Modelling the Flow of Fiber Suspensions in Narrow Gaps" (which
will be referred to as "Chapter FS" in the sequel).
943

It is worth noting that the numerical simulation of injection, compression and


related molding processes has received considerable attention for more than 25
years. Pioneer studies were devoted to modelling [2,27-34] and numerical [34-
46] issues, in order to scale the problem, to understand basic effects (such as the
formation of a solidified layer on the walls, the phenomena governing molecular
orientation in the final part, the role of viscoelasticity, etc.) and to develop first
simulation tools for 2D isothermal in-plane, or non-isothermal cross-sectional
calculations. The mathematical model was further elaborated [9,12,17,47-65],
with a view to correctly represent the fountain flow, to take into account the
presence of abrupt changes of thickness, bifurcations and other geometrical fea-
tures, to include polymer compressibility and to accurately predict the flow
stresses. More and more efficient numerical techniques were concurrently devel-
oped to solve the general 2lAD flow and heat transfer problem (which basically
consists of 2D pressure and 3D temperature calculations) [52,56a,64,66-81] or to
perform full 3D simulations [82]. At the same time, the model was extended to
reactive [5,6,25,32,49,58,72,83-86] and semi-crystalline [19,20,87-93] polymers,
and to compression molding [3,4,34,76,94,95], SRIM-RTM [7,8,21,25,76,81,96-
102] and other [103,104] molding processes.

2. RHEOLOGY OF INJECTION MOLDED POLYMERS

In continuum mechanics, constitutive equations relate the Cauchy stress aij,


heat flux qi, specific internal energy U and entropy S to the past history of
the material point thermodynamic variables [105,106]. For amorphous thermo-
plastics, the latter are the temperature T, the deformation gradient from initial to
actual state and the temperature gradient. Additional variables are required in the
case of reactive, semi-crystalline or fiber-filled polymers (see Section 6 and
Chapter FS). It is convenient with fluids to consider the pressure p as an addi-
tional thermodynamic variable, to start from the specific enthalpy H instead of
the internal energy, and to introduce a state equation for the specific mass p.
The stress is decomposed as ffij - -P~ij + "cij, where "cij denotes extra-stress
and Sij is Kronecker's symbol.
Amorphous thermoplastics obey viscoelastic rheology, which can be represent-
ed by a broad class of constitutive equations [ 106,107] among which the Leonov
model [ 108-111 ] (which is of the differential type and was developed on the basis
of irreversible thermodynamics) has been the object of particular attention in
injection molding [30,53,54,57,60a,64,65,67,112]. Since discussing the selection
of a viscoelastic model is outside the scope of this chapter, only the single mode
944

model will be presented. The elastic Finger strain tensor B (.e) and the symmet-
ric plastic rate of strain d o are defined to represent the partially relaxed elastic
deformation from initial to actual state and the irreversible relaxation rate of elas-
tic deformation, respectively (without volume change effect). Letting t , x i and
v i stand for time, Cartesian coordinates and velocity components, the constitu-
tive equations for B (.e) and dip can be put in the general form [57,111] 9
v
#~(e),4
B(;e),j + de pB(e),~ ,~J 4" "ik '*kjp -- 0 , (1)

and d p - b / 4 0 ( n ( e ) d - n(.e)-d ) , (2)

v
where the convected derivative _Bij(e) is def'med to maintain d e t ( B ( e ) ) - I 9

V' o3B(.e ) rgn ('e ) oavi R ( e ) - B [ f c ) oavj 20qVkB.(.e ) (3)


_ (e) _ ~ -t- v k "-'kj - q- --
l~ iJ -- oqt r;~Xk & ~ ~x k 3 & k u '

while b is a positive function of T and the invariants I 1 - B (e) , 12 - B ( e ) - l ,


and 0- O(T, p ) is a characteristic relaxation time for small amplitude motions.
R(e)d
The Einstein summation convention over repeated indices is used, and --ij

and Rv ij(e)-a denote the deviatoric parts of R--ij(e) and B (.e)-I Extra-stresses are
defined as follows 9

"gij -- 2 ( B(e)d
""
awlat,--itl(e)-d~ ij
V/012) 9
(4)

where the elastic potential W is a function of T, I 1 and 12 . Typically,

W - 3/2 G ( 2 n + 1)-1 [(1- f l ) ( I 1 / 3 ) 2n+l + f l ( I 2 / 3 ) 2 n + l - 1] , (5)

where G - G ( T ) is an elastic modulus and fl and n are constants.


Injection and compression molding are dominated by shear, since most often
thin parts are produced in practical applications. A quite accurate approximation
of the flow is therefore obtained by assuming a developed shear flow theology,
which provides the Criminale-Ericksen-Filbey ("CEF") model (see e.g. [106]) :
V
r ij - 2 rl d ij + 4 Iii 2 d ik d kj - ilt 1 d ij , (6)
945
v

represents the upper-convected derivative of dij,


v Odo ~0 o5,~ Ov
d ij = Ot
+ vk
Ox ~ & k d ~j
- dik J
Ox k ,
(7)

while the shear viscosity 0 and the 1st and 2 "d normal stress coefficients ~1
and I[/2 are functions of temperature, pressure and shear rate (~'), with

t2 -- 2 dqdq . (8)

In view of the numerical difficulty inherent to the viscoelastic problem, most of


the effort in injection and compression molding modelling has consisted in ne-
glecting the normal stress differences in (6), which gives rise to the g e n e r a l i z e d
N e w t o n i a n model. In that case, a large mass of data indicate that the Cross and
Carreau laws, which are particular cases of the Carreau-Yasuda law [113],

. - + , (9)

with a - 1 - n for the Cross mode! and a - 2 for the Carreau model, are suc-
cessful in describing the shear stress/shear rate relationship for a large variety of
polymer melts. This was demonstrated by Hieber et al. [52,59,114,115], Chiang
et al. [56,78], Chen and Liu [73], and Douven et al. [64], who obtained excellent
pressure validations in PP, PS, ABS, PC, nylon and PMMA filling and post-fill-
ing experiments. In (9), rl0 - ~10(T,p) and the constant v 0 are the zero shear
rate viscosity and shear stress, while n is the power-index (0< n < 1). The
Cross model adequately fits the viscosity curve for polymers of broad molecular
weight distribution [52,59,115] and should thus be generally preferred to the
Carreau model. In the log-log diagram, both laws exhibit a Newtonian plateau at
low shear rate and a power-law asymptote of slope ( n - 1) at high shear rate. It
is interesting to compare the shear viscosity of the Leonov model to the Cross and
Carreau laws. As an example, Simhambhatla and Leonov [111] propose to use
b(Ii,I2)-(I2/I1) m in (2) for LDPE melts. Equations (l-S) provide an equiv-
alent shear viscosity which exhibits the expected Newtonian plateau and power-
law asymptote. Other values can be obtained depending on the selection of
material fimctions.
A last simplification is often introduced. In view of the short process cycles in
injection and compression molding, the wall shear rate is generally quite high and
946

the power-law model,


FI -- Tl~),.gl-n,~, n-1 ,

provides a fairly good viscosity estimate to calculate the shear stress ( ~ - r/~') in
the cavity. Let us point out that Verhoyen and Dupret [23] have shown that the
kinematics of GN fluid flow in a thin cavity is relatively insensitive to the
viscosity dependence upon shear rate, even at low flow rates, whereas stresses
are highly viscosity dependent (see Section 3.3).
The GN model has been extensively used in the simulation of molding pro-
cesses, especially with power-law and Cross law viscosities [9,10,14,17,23,26,
31,37,40,41,52,56,64,66,69,72-74,76,78,79,104]. The present chapter focuses
on these models. The calculations of flow stresses via viscoelastic rheology and
GN kinematics is discussed in Section 3.4.
In polymer melts, the heat flux can be modelled by the Fourier law,
qi - - k o~/oqxi , (11)

where k - k (T,p) is the thermal conductivity. An empirical role and data are
proposed in [56,64]. According to Douven [ 116], it is not necessary to consider
anisotropic conductivity effects in injection and compression molding, even when
molecular chains are highly oriented by shear, since heat conduction within the
fluid is dominated by the transverse flux contribution (and thus only one compo-
nent plays a non-negligible role). It should be noted that most of the available
data do not take into account the pressure dependence of k, whereas compari-
son with experiments seems to indicate that this effect can be non-negligible (see
Section 5).
To complete the model, constitutive equations are needed for p , H , and S.
During filling, most of the fluid is in the liquid state and, for amorphous thermo-
plastics, p is a function of p and T (with very fast relaxation). The Spencer-
Gilmore state equation was often used in early studies [37,68]. Nonetheless, it
seems that the empirical equation of Tait [56,60,64,73b,116,117],
l i P - ~/Po - - C/Po ln(1 + p / B ) , (12)
where C is a dimensionless constant while l i P o and B are material func-
tions, is more appropriate to model l i p in both the liquid and the glassy states.
The glass transition temperature Tg is assumed to depend linearly upon p,
while 1/p o and B are linear and exponential functions of T,

Tg - rg~ + y p , Po -1 - fll + f12 ( T - Tg) , B - fl3exp(-flnT), (13)


947

with different coefficients ~i used above and below Tg. This approach is not
completely satisfactory in the vicinity and below Tg, where relaxation is much
slower. According to Kabanemi and Crochet [118], both stress and thermal ex-
pansion relaxations must be considered during the cooling stage. This is achieved
by introducing the concept of fictitious temperature (i.e. the temperature of local
thermodynamic equilibrium), together with additional relaxation functions (for
pressure and for fictitious temperature).
Finally, H is in principle a function of T, p and the shear elastic deforma-
tion. In the case of a GN fluid, shear elastic effects are neglected and
c)H/c?f - Cp , OH/Op -(1-~g/')/p , (14)

where T is the absolute temperature, Cp is the specific heat at constant


pressure and o~ - - p - a c)p/OT is the thermal dilation coefficient. The second
equation (14) is a direct consequence of the 2 nd law of Thermodynamics which, in
its rational form, imposes from the Clausius-Duhem inequality the definition of a
unique specific entropy from the differential dS - d H / T - d p / ( p T ) , a s well as
constraining r/ and k to be positive (see e.g. [105]). An empirical rule gov-
erning the thermal dependence of Cp is proposed in [56a] and material parame-
ters are given in [56b,64] for various polymers. The pressure dependence of Cp
is governed by (14), since c?2H/Op31" - c?2H/31"Op. However, in practice, this
effect is low.
To close this section, it is important to recall that amorphous thermoplastics
generally behave as thermorheologically simple materials [ 119]. The thermal de-
pendence of viscosity, relaxation times, etc., is therefore characterized by a shift
function a r, which expresses time-temperature equivalence and is defined by a
WLF-type relationship [56,64,73b,119])" below ( T o - C 2 ) , the material is fro-
zen and a r vanishes, while above (To-C2), a r is given by the expression

ar - exp(-C,(r-ro - Dp)(C 2 +r-ro)-'), (15)

where To is a reference temperature, and C 1, C2 and D are material con-


stants which depend on To (note that C 1 and C2 are more or less equal for all
polymers when To - T g , with C 1 - 3 2 , C2 - 5 0 ~ Hence, in (2) and (9), 0
and 7/ can be assumed to obey the following rules :

O(T, p)/O(To, O) - ar , rio(T, p)/rlo (T0,0) - ar 9 (16)


948

3. MATHEMATICAL MODELLING

3.1 The lubrication approximation and its limits


The aim of this section is to introduce the Hele Shaw model, which is basically
valid in the flow of most thermoplastics and some thermosets in thin cavities. In
particular, it applies during the filling and packing stages of injection and com-
pression molding. GN rheology and creeping flow will be assumed (with
negligible inertia and gravity effects [2]), while the influence of crystallization
and/or curing will not be considered.

3.1.1 A s y m p t o t i c analysis
Scaling is very complex in molding processes in view of the many dimensionless
numbers involved, and we refer to the book of Pearson [2] for more detail. Our
aim is, following Van Wijngaarden et al. [31] and Dupret and Vanderschuren [9],
to treat the lubrication approximation as an asymptotic model in order to provide
a sound basis for the construction of the boundary conditions. Under the assump-
tion of a low and slowly varying gap thickness 2h, asymptotic equations are
obtained by letting the dimensionless number e = h o / L tend towards zero,
where h0 and L are characteristic dimensions of the gap and the part, respec-
tively. This approach clearly distinguishes two flow regions [12,33,120], namely
the outer zone where the lubrication approximation is valid, and the inner zones
which are formed by the flow front, side wall ... boundary layers, that will be
investigated in Section 3.2.
It is sufficient to consider a model problem, where the fluid is assumed to be
incompressible, with a power-law viscosity and constant thermal properties,
while the part is symmetric with respect to a planar midsurface (Figure 2). In-
plane coordinates are denoted by x a , with Greek indices going from 1 to 2,
while the gapwise coordinate is denoted by z or x 3. Latin indices go from 1 to
3. The injected or initial fluid temperature Ti and the wall temperature Te are
given constants. The filling duration is denoted by it. The mass, momentum and
energy equations read as :

Ov, = 0 , (17)
C~Xi

Ox i Ox j
949

pCp - - ~ + OXi) /.if2 + k , (19)

with 71 - m o e bp-ar yn-1 . (20)


In equations (17-20), p, Cp, k, n, m o, b and a are material constants.
The gapwise velocity component v3 will also be denoted by w.
Physical quantities are scaled as follows :
I p P P
x a-Lx a, z-hoz , t=zt , h - hoh', (21)
1, , , )
va-L'c va, W-WoW , P-PoP , T=T i+ -T e T'.

The dynamic similitude assumption states that, when e--> 0, the dimensionless
fields v a' , w' p' and T' (as fimctions of x~, z' and t ' ) tend towards
given limits provided the function h'(x'#, t') is fixed and the less d e g e n e r a c y
principle is applied to determine the scaling factors w0 and P0 and to remove
any indeterminacy in the limiting process. Asymptotic equations are therefore
obtained by assuming that dimensionless pathlines tend towards given limits,
together with the fields p' and T ' , when the above conditions are satisfied.
Writing the mass equation (17) in the form

3va' + w0"t" 0w' = 0, (22)

less degeneracy [120] provides w0 :


w0 = eL/~" . (23)

Defining y,2 _ (Ov,a/Oz,)2, it follows from (8)that y2 _(y,2 +o(eZ))/(e~:)2.


The momentum equations become

Po - m~ eb'p'-a'T'~ "n-1 Ova + 0 e2


L oqx; E:2L,b'(E,g') n-1 ~ o~z' '

Po moe c? e b > ' - a ' r ' y "n-1 &'P, + (24)


F_L 03Z, C L~'(e~') n-1 0z'

+2Oz ~, 'T'y &" + ,


950

where a ' and b' are defmed by the relations


a" - a ( T i - W e) , b" - bpo . (25)

To remove any indeterminacy, a" and b" must be fixed asymptotically in such
a way that the products b'p" and a'T" do not degenerate when e---)0.
Further, considering equation (24.1), less degeneracy imposes to select P0 as

P0 - moe-ari/(El+n~n I . (26)
lk l

Finally, the energy equation (19) takes the form


o32T"
Gz + v a -Oxa
9 - 7 - + w" - B r eb'p'-a~'y "n+l + (27)
Oz ,

where the Graetz and Brinkman numbers (which represent the ratios between the
advected or dissipated heat and the heat which is diffused in the gapwise direc-
tion), are defined as

Gz = pCp(eL)2/(k'c), Br - mo e-aT~ L2El-n/('cn+lk(Ti - r e ) ) . (28)

Both numbers must be asymptotically fixed when e ~ 0. It is also convenient


to define the Nahme-Griffith number as Na = a ' B r .
The lubrication model is obtained by letting e tend towards 0 in (22), (24)
and (27), with w0 and P0 defined by (23) and (26) and the numbers n, a ' ,
b', Gz and Br fixed asymptotically. Retuming to dimensional quantities, the
simplified system is

av a g4, = 0

3p = 0 , (29)
oaxa 0 , 0z

v cp + v ~ -&--~ + w - - g - o
,nl

with 7/ given by (20) and ~' by the simplified expression

~2 _ (0 v~ / 0 z) 2 . (30)

To consider the presence of a solidified layer along the walls, the viscosity law
must be modified and additional dimensionless numbers are needed. In early
951

studies [2,28b,31,33], a no-flow temperature was often introduced but this con-
cept is difficult to define rigorously. It is better to use a WLF-type viscosity law
(equations (15) and (16.2)) throughout the entire cavity. Asymptotic equations
are obtained in the same way as in the present section.

3.1.2 Hele Shaw formulation


The Hele Shaw flow model results from integration over the gap of the simpli-
fied field equations. It is worth recalling the work of Berger and Gogos [37],
Williams and Lord [40], Broyer et al. [41 ], Hieber and Shen [66], Van Wijngaar-
den et al. [31] and many other authors [35a,38], who developed the first numer-
ical tools to simulate injection molding by this approach. Good reviews can be
found in references [47,64]. In the sequel, this model is presented in detail in
view of its connection with all subsequent developments. Non-synmaetric thermal
conditions are considered on the cavity walls, according to the approach of
Chiang et al. [56,78] (see also [16]). Compressibility effects are taken into ac-
count [ 14,56a,73b]. In order to separate the effects of pressure and temperature,
the specific mass constitutive equation is approximated as follows :
t9 -- Dr(1 -- ] ~ ( T ) - k - ) t ( T ) p ) , (31)

where Pr is the specific mass at the temperature Tr and zero pressure, while
fl(T) and 7(T) are material functions, the former being constrained to vanish
when T = Tr. Two kinds of gap average quantities will be used. Letting
~p(xa,z,t) stand for any physical field, the averages ~ and r are defined by
h
2h - f 0a , 0co, (32
-h
where to(x~,z,t) is a so-called profile function, which is proportional to the
velocity profile in the gap and will be defined later.
Consider first the in-plane momentum equations (29.2) (which are not affected
by compressibility). A first integration over the gap provides the relations
O~ ot _ Op z - z0
(33)
Oz 3x~ 71

where zo(x~ ,t) is the unknown level at which 3v~/Oz vanishes. From (33)
and (20), the viscosity is thus provided by the relation
o" mo - n-' . (34)
952

A second integration provides the in-plane velocity components from (33) and
(34), taking the wall no-slip boundary conditions into account"

_- - II /0x ll ' ' ~ e-bp/n mo -1In to(z,T(.)) , (35)

where the profile function to, which is defined by the integral

to(z,T(.)) - ~hz (( - z o)1~ - Zo] l/n-1 e ar(()/n d ( , (36)

must vanish on the walls. Hence, the additional relation


~o(-h,T(.))- o , (37)

can be used to determine the zero shear rate level z0.


A third integration over the gap provides the average mass flow rate p v a 9

2h pv a = - S o Opioaxa , (38)
where the fluidity S o is given from (31) and (32) by the expression

Sp - 2hPr II~p/~x=ll''n-' e -bp/n mo -1/n (1 - fiB + TBP) ~ , (39)

while ~ is obtained by carrying out an integration by parts 9


h
2h~- ~[Z-zoll/n+leaT(z)/ndz. (40)
-h
Consider now the mass equation

ap ~ ~ ( p v ~ ) + (~) - 0 (4~
oat Ox~
From (38), integration provides the 2D pressure equation"

c9 r (Sp O p ) (42)
~-(2hfi)- ~ ~, ,

where Sp is given by (39)and fi - Pr ( 1 - ~ + ~ p ) . It is also necessary to


integrate the mass equation (41) from z to h, in order to obtain w (which is
present in the energy equation). Taking into account the no-slip wall condition
w ( z - + h ) - +Oh / oat, one obtains the following relation"
953

pw - p-~ Vo~ dr + -~a~hz"v~dr , (43)

where the derivatives a*/o~ and O*/o~ are carried out while f'~ing z / h
instead of z (as in the a/& and O / ~ derivations). This expression is
convenient in numerical calculations because z / h (and not z) is used to posi-
tion the nodes in the gap.

~_<r //,,,./////////////,,1__/ "';~~~

Figure 3. Simplified model of the gapwise velocity component.

In practice [9], the expression of w is often simplified by dropping the a*/&


and O*/axa terms in (43), which consists in assuming a parallel, converging or
diverging velocity profile in the gap (Figure 3). This approximation is useful
because the exact calculation of w can be quite difficult since va varies very
rapidly in the vicinity of the side-wall re-emrant comers, due to the presence of a
singularity at these locations [20]. However, the presence of "solidified" layers
[2,28b,31,33] along the cold walls reduces the accuracy of this approximation,
which can be improved by taking the thicknesses c5+ and S- of the solidified
layers into account. In the molten zone ( - h + ~- <z < h - 6+), the approxima-
tion of w becomes

z + gb 0 6-) + (h ~- (60)+ v a (60) , (44)


5;
W ~ ~ ~ ~

with ~- - (~++S-)/2 , S0 - (~+-~-)/2 . (45)

The thicknesses ~+ and c5- must be evaluated from the "no-flow" temperature
Tm. Discretization nodes are optimally placed at the "solid"-liquid interface
(where T - Tm), and the mesh or grid must be adapted to ~+ and S- [31]~
Finally, when compressibility effects are taken into account and the lubrication
approximation is used, the energy equation takes the form (from (14))
954

0% + + w--s - ccc + +
(46)
-t- m 0 e b p - a T ~)n+l

The heat transfer in the steel mold can be simplified in the same way and the
governing energy equation within the walls is

PsCs(V3T
~ +-oat
O h ~o~) = (z>h
, or
_ z <_ - h ) (47)

where Ps, Cs and k s denote the steel material properties. Equation (47) is
coupled with the heat transfer in the gap. This model was improved by Chiang et
al. [78], who superpose a steady 3D contribution obtained by averaging the
successive molding cycles, on the transient contribution provided by (47).

3.1.3 Extensions
Extending the Hele Shaw model to curved midsurfaces is very easy [9], since
asymptotic analysis can show that curvature effects are negligible in Hele Shaw
flows because the gapwise dimension is "infinitely thin" as compared with in-
plane dimensions. The goveming equations must be adapted, since the easiest
approach is to use a unique Cartesian system of axes Ox i to represent all the
vector and tensor components, together with curvilinear coordinates (~1,~2,~3)
to represent the position. Any vector field (such as velocity) has 3 in-plane com-
ponents v i and an additional out-of-plane component w. Letting ~a denote
the in-plane coordinates and ~3 represent the distance to the midsurface, the in-
plane gradient of a quantity such as T is calculated by the matrix formula

Oxi
] = A(ATA)-I O~1 c?~2 ' (48)

with A ( 3 - E "]
~ 0xi , while DT/Dt is calculated as follows"

DT o~ ~T (ip)
= ~ + Vi-'---'X---t- W ~ . (49)
Dt & oaxi c?~3
Another very important extension consists in considering other viscosity laws,
such as the Cross law (see equation (9)). If a pressure-temperature decoupled
955

~e~o
103 . . . . . . . . . . . . . . . . I ' ' 103 .... 'r, "1 ..... "'''1 /' ' "

102 ............................ Z................................ !............... .-,

lOl .............................
i...............................
i/ ........ ::.. 101 ............................ + ............................... ~ '.t.~"......... .1

..............................................
i .........i............... 10 ~ ............................ i....................

1~176
I ........ .~..................... ::j:i!..----':::i~ .... ..'--::..::b'"" . . . . . . . . . . "1

10 -1 1 0 -1 ........................... -:::.... -:............... ..:..~.................

10 -2 ii..... 10-2 , ,
10 .2 10 4 10 ~ 10 -2 10 -1 10 ~

Z'/Z" 0 Z'/Z" 0

~/~o 1//1/o
10 x ........ t ........ J ........ I ........ t ...... ": 10~ ~: ........ ~ ........ i ........ ~ ' '"' ..... t ...... ".-

10 ~ 10 ~

! i ... f................!................i..............i
10 -1 . . . . . . . . . . . . . . . . . . . . . . .
1 0 -1

10 -2
10 3 10 -2 10 1 10 ~ 10 ~ 102
10-21
1( f3
!iii.i....iiiiiii10-2 10-1 100 101 102

Figure 4. Approximation of the Cross viscosity law by an expansion with 2 terms


(left) and 3 terms (fight: A3 = 1.78, B3 = 0.23) for a commercial grade PC
(Makrolon 2805, n = 0.093). Top : shear stress; bottom : viscosity.

numerical scheme is required, the approach of Verhoyen and Dupret [23] can be
used. The quantities 7/o and Vo are approximated as follows 9

1 _ Z 1 - m;(T) m~(p) (50)


7/o - k m,k(T) m2/C(p) ' ~'0

where the number of terms in the 1 st expansion must be limited to reduce the
computational cost, while the 2 no decomposition is sufficient because ~:o is most
often close to a constant. According to (29), (30) and (33), the shear stress ~:
can be written in the form

- oe - II~/ax~lllz- zol, (5~


so that Iz- zol/O - ~/ll~ 9 (S;>
956

On the other hand, from (9) and (50), the reduced shear stress z / z o is a given
function F n of the reduced shear rate (r/0 y/'c o )"

2" (~0 ~'/T0))I _Fn(T]O'~'] (53)


9o 1+ C/To -" "

This latter relation can easily be approximately inverted :

110 fl _ Fn_ 1 Z _ Ai z , (54)


%
where the first two terms produce the Newtonian viscosity plateau when z --->0,
and the power-law asymptote when z --->oo (by selecting A1 = 1, B 1 = n,
A2 = 1, B 2 = 1), while a third term, which is calculated from a best fit (with
n < B 3 < 1 ), is useful to improve the approximation (Figure 4).
Finally, from (51), (52) and (54), and using the lubrication approximation (equa-
tion (33)), the in-plane velocity gradient can be approximated by the sum
Ov,~
3z cgxa mlk(T)rn2k(p)
(55)

ItAI9 m;(T)~(p) I z _ z0lsi/n sign(z z0)1.

In the case of symmetric temperature and velocity profiles (z o = 0), each term of
(55) can be integrated as in the case of a power-law viscosity. The non-sym-
metric case requires to approximate [z - Zo[8i/n sign(z - z0) as a stun of powers
of z and z 0. At each time step, the gap integrations can therefore be performed
once and for all when the temperature profile is known.

3.2 B o u n d a r y conditions

3.2.1 Fountain flow


The boundary conditions prevailing at the moving front during filling give rise to
the fountain flow [121] (Figure 5). Due to the no-slip wall conditions, a strong
deflection of the pathlines is observed near the front. Hence, the front region is
continually fed with fluid coming from the core of the cavity, while the front feeds
the regions located closer to the walls.
957

X3
I
|

_ m R . ~
L
r Xt

Figure 5. Sketch of the fountain flow 9 relative motion of the fluid with respect to
the front.

Previous investigations of fountain flow first aimed at understanding the basic


mechanisms governing its kinematics and its effect on the frozen molecular orien-
tation [27,41]. Richardson [33] developed asymptotic expansions by matching
the front and Hele Shaw flow regions. Pearson [2] identified two zones in the
molded part, namely, the set of material points having experienced fountain flow
and the remaining region (Figure 6). Numerical experiments were performed by
several authors [45,48,51,54,55,122,123], in order to investigate the behavior of
tracers submitted to fountain flow and the influence of shear thinning, capillarity
and viscoelasticity on the front shape and the frozen orientation. The experimen-
tal validations performed by Castro and Macosko [32], Coyle et al. [50] and
Behrens et al. [51] showed good agreement between measurements and predic-
tions, without significant influence of non-Newtonian effects.
The 2 no objective of investigating the fountain flow was to develop approximate
models able to capture its basic kinematics without resolving all the flow details.
We here present a singular perturbation analysis [12] of the model developed
(independently) by Manas-Zloczower et al. [49] and Dupret and Vanderschuren
[9]. Starting from the model problem of Section 3.1.1, a particular system of axes
Px i is selected (Figure 5), such that P moves with the flow front, Px 1 is par-
allel to the front velocity and Px 3 is oriented in the gapwise direction. Asymp-
totic equations are established in each of the inner regions, which are all located
in the front zone (in contrast with the single outer Hele Shaw region). Limit
equations are obtained under the assumption of dynamic similitude and taking
into account the matching conditions between inner and outer zones. The scaling
used in the inner flow regions is based on considering a set of decoupled
958

(a) ~.o ~ ~2.a


08 \ ~...~..
-

(b)
0.0
.ok,
0.8 max ~ '

'~ 0.6

t~ 0.4

0.2 o.5"~ 0.8

0.0 L
o 50 100 150 200 250 300
x [mm]
Figure 6. Cross-sectional space-time distributions obtained by (a) the isothermal
and (b) the non-isothermal filling of a rectangular plate. The residence time of the
material points is in seconds. The separation line (in bold) is not the line of maxi-
mum residence time when w is taken into account [20].

"infinitely zoomed snapshots" at any location and time on the from. More
precisely, letting the origin of the axes move with the front and selecting any
initial time during filling, physical quantities are scaled as follows :
X i -- hox i , t = e "C t ,

{ ~-1' ' (56)


Vi = t v i9 ,
p =
PO( p " + Ep*) ,
T =
Ti + ( T i _
Ze)T" ,

where the matching conditions impose (and this is a key point) to keep the same
pressure and temperature scaling as in the outer zone.
The mass equation is not affected by the scaling process. Its dimensional
asymptotic form is thus exactly (17). In a similar way, all the terms defining ~'
in (8) have the same order of magnitude and the dimensional asymptotic shear
rate is thus exactly given by this equation. Also, the viscosity is not affected by
the limiting process. The dimensionless momentum equations reduce to
959
o3p'/cgx[- O(e), which means that their asymptotic dimensional form is
o ~ / ~ i - O. Therefore, the 0th-order term of the pressure expansion is constant
in every inner front region. In a further step, the 1St-order terms of the dimension-
less momentum equations are considered :

eb'p'-a'T" e (onv[ + +

which provides the asymptotic dimensional equations 9

O~+ t~ ((O~ i O~j)) (58)


0x~ =0xj 77 9xj+0x~ ,
where p + denotes an additional pressure contribution in the front zone.
The dimensionless energy equation is

Gz - ~ + v[ - e Br e b'p'-a'T" ~,n+, q. OqX;2 J O(~) , (59)

where, for matching purposes, Gz, B r , a" and b' have the same value when
e ~ 0 as in the outer flow. The dimensional asymptotic form of (59) is thus
O T / O t - OT/Ot + v i anT/oqxi - 0 . (60)
The temperature of each material point therefore remains constant when it crosses
the front region.
The matching conditions impose velocity, pressure and temperature "continuity"
between the inner and outer regions [120,124] :

lim ( v l , v 2 , v 3 , p , T ) -
HS , v 2 HS O, p HS T HS )
v1 (61)
xl-~-oo ' ' (xl =0) '

where the superscript (t-/s) indicates the Hele Shaw outer solution. The order
of magnitude of w in the Hele Shaw flow (equation (23)) imposes to match v 3
with 0. As different scales have been used for both x 2 and t in the inner and
outer expansions, the relations (61) indicate that inner flows are two-dimensional
and quasi-steady. This discussion indicates how to select appropriate front
boundary conditions in the Hele Shaw model :
9 The asymptotic front region is infinitely thin and consists of a set of straight
segments perpendicular to the midsurface.
9 The front pressure in the Hele Shaw flow is the atmospheric pressure when
960

surface tension is negligible [50].


The temperature of the material points is kept constant when they (infinitely
fast) cross the front region. Hence, the temperature of a material point re-
entering the Hele Shaw region at the level z2 must be the same as when it
entered the front region at the level z I (Figure 7) :

r(z 2) - r(z l) . (62)

In fact, the characteristics of the simplified energy equation (46) are leaving or
entering the flow domain depending on whether the material point is moving
faster or slower than the front. Hence, one thermal condition must be imposed
at the level z2 while no condition may be imposed at the level z~.

....../ / / / / / / w a " / f / / / S ////


~ z21

Zl~
w

Figure 7. Simplified fountain flow model.

The front condition (62) is not affected by compressibility effects, since the front
inner pressure is a constant. The pairs of levels associated through condition (62)
are governed by mass conservation. Letting r.of denote the mass averaged
profile function (see (36)),

p(T, Pa ) o)f - p(T, Pa)O) , (63)

the condition linking z 1 and z 2 reads as

p - ) - 0 (64

Note that (_Of provides the front in-plane velocity vaf from (35).
The above discussion shows that, according to the terminology of Garcia et al.
[58], the "mass balanced front model" introduced in [9,49] relies on a sound
mathematical basis and fits perfectly with the Hele Shaw approximation. Also,
this approach can be used to model the effect of fountain flow when fiber rein-
forced polymers are molded (see Chapter FS). All the "simplified and averaged
961

simplified front models", which set the front temperature uniform and equal to the
centerline [40b,69] or average temperature just upstream of the front, present
mathematical inconsistencies which can induce heat losses at the front. However,
these models generally behave quite well [58] since temperature variations are
weak in the front region. Hence, no significant problem can be detected with the
mesh or grid refinements typically used. Finally, the solution could be improved
by solving the simplified inner flow problem and improving matching between the
inner and outer solutions. This coupled technique is quite complex and, in this
respect, the simplified model of Castro and Macosko [32], as adapted to 3D
flows by Peters et al. [ 104], provides a good method to gain accuracy.

3.2.2 Wall boundary conditions


In Hele Shaw flows, the boundary conditions are treated in a different way along
the cavity side walls (whose area is O ( h o L ) ) and on the larger upper and lower
walls (of area O(L2)). T h e side walls conditions [124] can again be derived

from an asymptotic analysis. This theory will not be detailed. Using the scale L
along the side wall and h in the other directions, it is easy to develop the inner
flow governing equations and to conclude that the following conditions must be
imposed on the Hele Shaw solution at the side walls :
9 free slip of the averaged flow, which provides the condition 0p / ~gn = 0 ;
9 no thermal condition, because the characteristics of the energy equation are
tangent to the side wall.
It is also possible to define along the side walls a classical displacement boundary
layer thickness, which does not increase in the flow direction.
The situation is totally different on the upper and lower cavity walls, since the
objective is to approximate their thermal effect by means of boundary conditions.
For that purpose, it is sufficient to consider the upper wall ( z - h). The heat
transfer in the metal is governed by the 1D equation (47), with the following
boundary conditions :
9 continuity of temperature T w at the solid-fluid interface (z = h);
9 balance of heat flux at the interface (the heat flux entering the wall is qw);
9 matching with the initial wall temperature Te 9 T - T e , if (t < tf , z >__h ) ,

where t f denotes the time at which the front reaches location x~;
9 matching with the regulation temperature T e 9 T = T e , if (t > t f , z - h + ec),
where ec is the distance between the interface and the thermal regulation
channels; e c can depend on position x a .
962

Zt
h+e c coolant

h + e(t) T
wall
o.(t) T :::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
":::'::"":.'.".'5.'L':':.",':',':',':'.'L'.".':','L'.5"L
h . . . . . i ._
r

polymer
z)

Te Tw(O Temperature
Figure 8. Typical temperature profile and thermal penetration length.

An accurate wall thermal boundary condition can be established from the theory
developed by Dheur [16] (see also [9,26]), as long as a monotonic and smooth
behavior of qw prevails ( T w is generally higher than T e and qw is positive
and decreases with time with thermoplastics, while T w is lower than Te and
qw is negative and increases with time in the case of thermosets). In that case,
the temperature profile within the wall can be approximated by a linear function,

{ ~ -_ TeTW(i - (z - h)/e) + Te (z - h)/e , if h >< hz +


< he +, e (65)

where e is the (unknown) thermal penetration length (Figure 8). The required
boundary condition is established by means of the Galerkin method, using linear
test functions and taking into account the flux conditions 9

- k s -a-T~ ( z = h ) = q w , -
- - ~ ( z - h + e) = 0 . (66)

After some calculations, this yields the pair of equations


e 0"7"w 3 k s ( T .
PsCs--
40t
+2 e ~ w - Te)-qw -0
(67)
0
PsCs -~((Tw - Te )e 2) = 6ks (T w - Te ) ,

whichgovem T w, qw and e as long as e < e c. When e reaches e c, a


single condition coupling qw and T w must be applied :
963

-e~0Tw + ~ksl~Tw-Te)Z ~ _ qw = 0 . (68)


PsCs 3 0 t e

The thermal behavior of the upper and lower walls is complicated by the foun-
tain flow, which generates a thermal shock between the core polymer and the
walls, with a resulting singularity on the front-wall intersection lines [28a]. This
effect can be analyzed by means of the thermal shock theory between two semi-
infinite rigid bodies of initial constant temperatures Te and TZ o [44], since the
fluid motion and the non-uniform temperature field play no role close to the sin-
gularity during the first instants following the shock. This theory shows that Tw
remains constant after the shock. In a similar way, when a thermal shock is con-
sidered between a semi-infinite body (the fluid) and a model body (the steel wall)
where conditions (67) hold, Tw again remains constant after the shock [26] 9

T w - (or T e + [3 Tfo )/(o~ +/~) , (69)

with -(3S s sl8) , -(pc.kl ) ,

and where Tfo denotes the temperature of the fluid making contact with the
wall at time t f . After the shock, the temperature profile inside the "semi-infi-
nite" fluid is easily calculated in terms of the complementary error function, while
the thermal penetration length is

e -(6ks(t- tf)/PsCs) 1/2 . (71)

3.2.3 Bifurcations and abrupt changes of thickness


The theory developed to build up a simplified model of fountain flow can be
easily extended to the so-called singular regions, where thickness variations are
abrupt or very steep and take place in a region of the same order of magnitude as
the part thickness. Singular regions include edges, abrupt changes of thickness
and bifurcations as in the case of T shapes (Figure 9).
Inside the singular regions, the same scaling (56) is introduced and a fixed Eu-
lerian reference frame is used in each "infinitely small and rapidly fading" inner
zone, where asymptotic analysis shows that the scaled flow is essentially 2D and
quasi-steady, while pressure is almost constant, and heat conduction and viscous
heating are negligible. According to Couniot et al. [17] and Dheur [16], the
following jump conditions must therefore be imposed"
9 Each singular region is infinitely thin, consists of straight segments normal to
964

(a) 11111/I//// (b)


I//I/111/I

///// ~/i// I
I//11111/11111

(d) /
(c) f / / / / / /
IIII/
//////,
,,,..._
///I/IS
,,....-

/ / / / / / / / 7 / / / / / / / / 7 ~ 7 / / / / / / I / / / / / /

Figure 9. Singular regions in Hele Shaw flow" (a)edge; (b)abrupt change of


thickness; (c) bifurcation with separation; (d) bifurcation with junction.

the midsurface(s), and is represented by a line on the midsurface.


9 Pressure is govemed only by the parabolic conservation equation (42) and is
continuous across the singular line, without singular losses of head.
9 The temperature of a material point crossing the singular region remains con-
stant (Figure 10). This condition has the same form (62) as in fountain flow
modelling. The pairs of levels associated through condition (62) are deter-
mined by mass conservation. In the case of the separating flow in a bifurcation
of upstream and downstream half-thicknesses h-, h( and h~, the levels z 1
and z 2 are linked by the condition

,~22 p (v, p) <o d= - P (7"p)


, odz , (72)

or if2h~ p (T ' p)09 dz - ff ~h-Q (T, p)o9 dz (73)


depending on whether the upper or lower downstream branch is considered.
This analysis shows that the edges play no role in the simplified model. In fact,
the influence of the geometry reduces to the dependence of the thickness and the
midsurface metrics upon position. Moreover, from (36) and (37), it turns out that
the velocity and temperature profiles are not affected by an abrupt change of
thickness as long as a power-law viscosity is used. The case of a bifurcation is
more complex. Two situations are possible, viz. flow separation and junction
(Figure 9), and the flow pattern is governed by the pressure gradient orientation
965

T(z2)=T(Zl ) 2h2

2h-

~'///////// 'j~
2h +
1

Figure 10. Thermal jump condition across a bifurcation.

with respect to the singular line. Flow separation and junction can be simultane-
ously present along different segments of a bifurcation. In both cases, an impor-
tam consequence of (62) is that the temperature and velocity profiles are no
longer symmetric with respect to the midsurface in the downstream branch(es),
even when the upstream flow(s) is (are) symmetric [17,125]. This effect, which
can influence the flow far downstream from the singular line, will be investigated
in Section 5.1. It should also be noted that an internal thermal shock takes place
in the case of flow junction.
This theory has multiple extensions, in particular when dealing with semi-crys-
talline or reactive polymers, or with fiber suspensions, or with the filling of fiber
mats (see Section 6 and Chapter FS).

3.2.4 Influence of the feeding system


From the nozzle to the mold cavity, the delivery system in the injection molding
process consists of a tronconical sprue, a set of trapezoidal runners, and of one or
several gates which can take on various shapes. The role of the feeding system is
to transfer mass and heat inside the mold and thus to supply boundary conditions
downstream from the gate(s) for the calculation of pressure, temperature, species
concentration, crystallinity degree, fiber orientation ... in the cavity. Hence, the
importance of the delivery system is indirect. Nevertheless, its influence can
prove crucial in obtaining satisfactory results in production devices [20]. Since
the pioneer work of Williams and Lord [40a], Richardson [33] and Pearson [2], a
number of more or less simplified flow models have been developed with the aim
of balancing the delivery system when multiple cavities are simultaneously filled,
or when multiple gates are used [11,56a,70,78,79,126,127]. However, oversim-
plification can lead to misleading results, since the temperature, species, crystal-
966

linity, fiber orientation ... non-uniformity, which is most often observed just
downstream from the gate(s), is a consequence of the flow kinematics in both the
runners and the gates. A typical example will be analyzed in Section 5.2. In
general, the complexity of the role of the feeding system can be explained by the
very low thermal conductivity and quasi-vanishing diffusion coefficients of com-
mon molten polymers, which result in a lack of mixing when the fluid crosses the
runners and the gates.
The flow in the gates is highly complex, since it is 3D and viscoelastic, with
high Deborah number [2,22,79]. Model simplifications are hardly conceivable.
Quasi-steady flow may be assumed and heat conduction may be neglected, while
the effects of polymer compressibility and thermal dilation can be simplified. The
main issue is to determine whether a model coupling extensional and shear
viscosities is sufficient to represent the main transfer effects. Similar, but less
critical, difficulties appear in the bifurcations of the runner system.
To accurately calculate the flow in the runners, their trapezoidal shape must be
taken into account since it induces transverse recirculations that can affect the
heat and mass transfer. The simplified mass and 3D energy equations are estab-
lished in the same way as in the mold cavity by means of the lubrication approxi-
mation, and all details will be omitted. Letting x a a n d z denote the transverse
and axial coordinates, it is again convenient to introduce a profile function oJ,
which depends upon temperature profile only [22]. The axial velocity component
w is proportional to o~, which obeys the equation

c) e_aT 0(.0 _
~x~ [l ll ~ -1 , (74)

with o~ = 0 along the boundary of the nmner section. Evaluation of the trans-
verse velocity components va in the energy equation must be performed care-
fully, since mass conservation provides a single equation. The system is closed
by taking the 2D c u r l of the non-simplified transverse momentum equations 9

3
[ (();
0 0v/~ Ovr rl~-~-

3.3 Geometrical aspects of the front motion


From mass conservation, the front velocity vaf is given by the expression

vaf = pv a/p . (76)


967

-,, -.j x2

'", , Front
I I

m~ ~ p ~ - ~t/(2 m +1)
...... ::- ..... ~: m, ...............

1; Oblique 0 * Horizontal p*
side wall side wall

t ' ~ m -1/4 "m = -1/3

\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ ~ \ \ \ \ ~ ~ \ \ \ N \ \ \ \ \ \ \ \ \ \ ~ \ \ \ \

Figure 11. Hele Shaw flow past a comer" schematic diagram in the real and
complex potential spaces (top); successive fronts calculated from the analytical
solutions (bottom).

Hence, from the front pressure condition and equation (33), vcr is normal ev-
erywhere to the front line and the flow fronts are perpendicular to the side walls
(in Hele Shaw flows, side walls are equivalent to reflection symmetry lines).
In this section, the following problems will be addressed"
9 How does the front cross a side comer, or the vertex of a conical midsurface ?
9 How do weldlines form, and how does the front collide with the side walls ?
9 What is the effect of an abrupt change of thickness or a bifurcation ?
968

3.3.1 Effect of a side wall corner


The flow front shape downstream of a side wall comer can be determined by
means of conformal mapping in the Newtonian isothermal incompressible case.
Consider in Figure 11 a uniform flow of speed v0 that nms past a side wall
comer of angle ( 1 - m)zc (m is positive in the case of a re-entrant comer and
negative with m > - 1 / 2 in the case of a salient comer). The thickness is uni-
form. Define the velocity potential r and the stream function
(with ~-~ - ~9r - e~/~ o~/0X/~ ), in such a way that the complex potential
f = ~0+ i ~ is an analytical function of the complex variable ~" = x 1 + ix z , with
the front boundary condition r on F(t). The solution can be found easily
by expressing ~" as a function ( ( f ) since the functional domain of ~ is the
fixed region (r < 0 , V > 0).
The complex velocity ~ = V1 + i V2 is given by the relation

(77)

Self-similitude of the solution at different times imposes that ~ be orthogonal to


(~/t - ~) along the front, or equivalently that

Re 0~- -1 -0 0r - 2t , if t>_O, ~ - 0 . (78)

The other boundary conditions are

: Im(r : o, if - o, -Kt<_r
Im(~e-mn/) : 0, if lit-O, <-Kt, (79)

d~/df ~ eim~r/Vo, if Ifl with ~<0, ~>0,

where 2hp K t / S o is the comer pressure and the constant K needs to be deter-
mined. After a few trials, the solution is found to be

( 1 1
~ + ~ ~
(8o)
rot 2m + l v0 t 2m + l v0 t

Hence, K - Vo2/(2m + 1) and the front velocity along the side wall Vfe is, from
(77) and (80) with r - ~ - 0 , V fe -- Vo/(Zm + 1).
969

Solutions are shown in Figure 11. When the front runs past a salient comer, or a
cone vertex whose total midsurface aperture angle is lower than 2Ir, a stagnation
point forms immediately at the comer or the vertex, and a material line joining the
comer to the front extremity, or the cone vertex to the closest front point, starts to
stretch out. This line is initially infinitely small and forms a weld line in the latter
case. When the front runs past a re-entrant comer, or a cone vertex whose aper-
ture angle is higher than 2zt, the velocity immediately becomes infinite at the
comer or the vertex, which are continually crossed by material points. These
conclusions are reinforced when the viscosity obeys a power-law. With a Cross
or Carreau law, the flow in the vicinity of a stagnation point is governed by the
Newtonian viscosity plateau.
Analytical solutions of the flow in Hele Shaw cells were investigated in detail by
Richardson [128] and Entov et al. [129], while Aronsson [130] develoved very
simple solutions for the limiting case of a zero power-index fluid (n - 0).

3.3.2 Weld lines


The flow past a cone vertex of aperture angle lower than 27r, that was analyzed
in the previous section, exhibits a key aspect of the kinematics of weld lines in
Hele Shaw flow. When two fronts collide, a single pair of material points meet
and the weld line results from the fast stretch of a material line that initially
reduces to this pair of points. The evolution of weld lines is complicated by the
observation that these "lines" are in fact true deforming surfaces. Starting from a
straight segment perpendicular to the midsurface at the location of the front meet-
ing, the infinitesimal weld surface begins to stretch rapidly and becomes a mate-
rial surface, that moves at the fluid velocity. Hence, the weld surface is normally
strongly deformed by pressure gradients. Moreover, its extremities experience
fountain flow, which typically induces a 3-fold weld surface in the front vicinity,
while a 2-fold shape is generally prevalent elsewhere. More complex behaviors
due to changes of flow pattern are possible. The traces of the weld surface on the
cavity walls are governed by fountain flow, since they represent the successive
positions of the material points that experience a thermal shock with the walls.
These traces are always symmetric with respect to the midsurface, even when the
velocity and temperature profiles are not.
Beside the exceptional case of a flat weld surface, thermal diffusion will remove
the temperature discontinuities across the surface. This will not be the case for
any quantity (such as crystallinity degree, fiber orientation ...) that is not subject
to diffusion. In a similar way, the derivatives of these quantities are normally
discontinuous across the surface that separates the core fluid from the material
points having experienced fountain flow.
970

v2~2+ vlf+2 (b)

V+

+ + v~)

d/~ ~ffont
v;: ,
. ~__ x 1
Figure 12. Flow across an abrupt change of thickness" (a) sketch of the front
velocities; (b) uniqueness of the intersection of the front and the singular line.

When a front collides with a side wall, the kinematics are the same as when two
symmetric fronts meet. Boundary layers are however different.

3.3.3 Abrupt changes of thickness and bifurcations


The principles governing Hele Shaw flow in the presence of a singular line
were developed in Section 3.2.3. Following Couniot et al. [17], precise angles
are formed between the front segments and the singular line. Consider first an
abrupt change of thickness, and let h-, vlf, V2-f , Sp and h +, V ~ f , V ~ f ,
+
S o denote the upstream and downstream values of the half-thickness, the normal
and tangent velocity components, and the fluidity, respectively (Figure 12). Iso-
thermal flow and fluid incompressibility are assumed. Mass conservation yields
h+v[: - h Vl: , (81)
while pressure continuity implies continuity of the tangent pressure gradient,
which from (38) and (76) is written as

2h +p v2+f / S; - 2h-p v]f /Sp . (82)

Finally, uniqueness of the intersection between the upstream and downstream


front segments and the singular line provides the relation

EV +2]/v+ E 2
"k- V 2 f 2f Vlf q" V 2 f 21/v' 9 (83)
971
Equations (82) and (83) can be cast in a more convenient form. In the case of a
power-law viscosity, the fluidity is given by (39), or equivalently, from (38), (40)
and (76), by

= m0 -1 ( n - l + 2)-nh , (84)

while ][vaf[I- Vlf 2 -FY2f 2 )1/2 can be eliminated by means of (83). Hence,
after some calculations, the following relation is obtained :

v 2+f / h +2 -- V2T
- / h -2 . (85)
Also, multiplying (83) and (85) side by side provides the relation

-- + +
The 3 equations (81), (85) and (86) involve 4 tmknowns (vlf ,v2f, vlf, v2f ),
which can thus be determined to within a multiplying factor. In particular, the
angles o~- and t~ + formed by the upstream and downstream velocity direc-
tions with the singular line are functions of the ratio (h+/h -) only"

tan2 o~- - tan 2 o~ = . (87)

The same formulas apply whether h + is higher or lower than h-. Two patterns
are thus possible, depending on which is the feeding and which is the fed branch.
Also, it can be observed that a vanishing or infinite thickness ratio corresponds to
the case of a lateral side of the mold. When h + = h - , the angle a + = a- is
indeterminate and (87) is no longer valid. When h is close to h-, a front cur-
vature boundary layer is thus present near the singular line in the Hele Shaw flow.
The m-plane dimension of this boundary layer is governed by the flow pattern and
the thickness ratio, but not the actual thicknesses. Hence, although (87) is valid
in theory, numerical solutions can exhibit angles o~+ and a - that differ from
their theoretical values, except if a very strong mesh refinement is used. In
reality, the flow in the singular region is governed by the theory developed in
Section 3.2.3. The inner flow domain is partly or totally superposed on the front
curvature boundary layer, which can be absent from experimental results.
The analysis of a bifurcation can be handled in the same way. Two problems
972

must be set apart, viz. flow separation and junction. A different equation governs
mass conservation in these cases, while pressure continuity and uniqueness of the
intersection between the fronts and the singular line each provide a pair of
equations. A total of 5 equations with 6 unknowns is thus found and the same
conclusions can be drawn as in the case of an abrupt change of thickness.
Verhoyen and Dupret [23] have shown that this theory can be extended to more
general viscosities, including the Cross and Carreau laws. Indeed, the jump equa-
tions (81-83) remain valid, together with (33), (38) and (51) (from the lubrication
approximation). Hence, the following expressions are successively found for the
fluidity:

I1@/~11so -112~ 11, (88


IlOplo~<,llso -Ilop/O~ III_".paz Iz~(C-zo )l. dr
p dz [.z i' sign(~"- Zo)d~"
h
= Ip~ Iz- ~01dz (since 19 is a constant). (89)
-h

On the other hand, from (51) and (54), the shear rate is

-- TO 00 -1F,,-l(llop/Ox<,lll z - zol/~o), (90)


where the function F n is defined by (53). Hence, introducing (90) into (89)
yields, in the isothermal case (with zo - 0 ) 9

So - 2p't'o 2h 170-1 II@/ex~l1-2Hn(h'ro -1 I1@/~<~11), (91)

where H n (~) - /~~ I~o~ Fn-' (~) d~ (~" > O) . (92)

However, from (88), H0p/Oxal[ S o h - ' [l~o~l[-1 i~ continuous across the singular

line, as is II0p/0xall2 So 2 h -2 v # (from (83)), and thus"

2 -I1 / 11 s;
since h+v;f / S ; - h-v2f / S ; , (94)
in view of the pressure continuity gradient. Therefore, comparing (91) and (93)
973

shows that H n (h [lOp/ oaxaIi/*0) is continuous across the singular line, since 7/0
and "ro are functions only of p. Hence, h IlOP/Oqxall/'r0 is itself continuous
since H n is strictly increasing, and thus from (93) :

Sp+ // h +3 - S p / h -3 , (95)

which also involves from (94) the continuity of v2f/h 2 (equation (85)).
Further developments proceed as with a power-law viscosity.
Similarly, it can be shown [23] that, when an unbalanced bifurcation is fed with
incompressible isothermal fluid through the upstream branch, and the incident
velocity is uniform and normal to the singular line, the motion of the fronts down-
stream of the bifurcation depends only upon the thickness ratio, but not on the
viscosity law. These considerations have far reaching consequences and explain
why the filling kinematics are, in general, only weakly dependent upon shear
thinning effects and dominated by thermal effects.

3.4 Viscoelastic effects


Since the aim of this chapter is not to analyze in detail the role of viscoelasticity
in injection and compression molding, this section is devoted to a review of the
existing literature. Viscoelastic effects are significant in critical regions, mainly
the front zone, the edges, abrupt changes of thickness and bifurcations of the part,
the gates, and the branchings of the delivery system. They also influence the flow
stresses (via normal stress differences) and the cooling stage [118], and are
directly connected to the generation of frozen-in orientation. Nevertheless, the
GN model generally provides accurate results since the flow is dominated by
shear, and also thanks to the rather low sensitivity of the gap-averaged kinematics
upon shear thinning effects as shown in the previous section.
Early studies (see the review of Grmela [47]) were limited by the tremendous
difficulty of the viscoelastic problem in normal processing conditions [107].
Tadmor [27] and Dietz et al. [29] developed simplified models to evaluate foun-
tain flow extensional effects and the resulting frozen orientation. Kamal and co-
workers (see e.g. [70,122]) investigated 2D cross-sectional flows in the cavity
with a White-Metzner rheology, but this approach is limited by the poor proper-
ties of this model [ 131 ]. Most of the effort bore upon using the Leonov constitu-
tive equation (see Section 2), since its material constants can be determined from
standard experiments. The 1D or 2D incompressible multi-mode model [108-
111] was used by Isayev and Hieber [30], followed by several authors [53,54,65,
67,112], to analyze the effect of the processing conditions on the flow and
974

N1 (MPa)
0.1
Direct
- Indirect

6
0.06

0.04

0.02
tO
0
-' -0.5 0 0.5 1

Distance midplane (mm)


Figure 13. Predicted first normal stress difference after 4.7 s in the filling of a
80 50 2 mm PC rectangular plate (vf = 120 mm/s, T/ = 320~ Left" direct
simulation; right" comparison between direct (solid line) and indirect (dotted
lines) calculations 8 mm from the linear gate. (From [57]).

residual stress distributions in junctures or in the front region. Experimental


comparison by means of birefringence measurements showed reasonably good
agreement, such as in the subsequent investigations performed by Haman [60],
using the 2D compressible Leonov model [57], and by Chang and Chiou [63],
using the K-BKZ model. In these latter studies, such as in the packing stage
analysis of Nguyen and Kamal [61] and in [65], an attempt is made to extend the
lubrication approximation to viscoelastic rheology by assuming a priori equal
scales for the extra-stress components (without considering the less degeneracy
principle), while simplifications are introduced more carefully in [60a]. The
scaling problem still needs to be solved for viscoelastic flows in thin parts.
An important step was made when Baaijens [57] demonstrated that simulating
GN kinematics, and re-calculating the stresses from a viscoelastic rheology and
this pre-calculated velocity field usually provides a fairly accurate estimate of the
solution in shear dominated flows. This indirect method, which was further used
by Douven et al. [64], Caspers [132] and Kabanemi et al. [62] to simulate the
molding of Leonov and Wagner fluids, represents a key step forward since it is
much easier to solve the indirect than the direct (coupled) problem. Figure 13
depicts a typical result from [57], where the agreement between indirect and
direct calculations is obviously excellent.
975

4. N U M E R I C A L SOLUTIONS

4.1 Review of available methods


Beside pioneer studies performed by means of the finite difference (FD) method
in order to solve the 189 problem [31,35a,40b], most of the effort in injection
molding modelling has borne upon developing mixed algorithms, where the 2D
pressure field is evaluated on the midsurface, while the temperature and other 3D
fields are integrated by means of a hybrid scheme over the entire filled region.
These techniques, which are detailed in the literature below and will not be
discussed, will be classified according to the solution of the free boundary
problem since the main difficulty to be overcome in the simulation of molding
processes arises from the fact that the filled domain evolves considerably with
time. The particular method developed at UCL is described in the next section.
Two general approaches were followed, the 1 st consisting in using control
volumes on a fixed mesh covering the entire midsurface, while the 2 nd class was
based on tracking the flow front(s) and adapting the mesh (or grid) at each time
step in order to cover only the filled domain. The l~t technique, which is very
robust and by far the easiest to implement, does not involve a discrete front repre-
sentation. Front phenomena are thus more difficult to model and strong mesh re-
finement is required in the front zone. This approach, which was 1~t used by Lord
and Williams [40b] and Broyer et al. [41] to model 2D non-isothermal cross-
sectional, or 2D isothermal in-plane flows, has been the object of intense investi-
gations from the Cornell group [56a,71,72,78]. It was further used to simulate
compression molding [95] and extended to the K-BKZ model [63], see also [75,
76,79,80]. Following Chiang et al. [56a,78], this technique attaches to each node
of the 2D mesh (surrounded by a given control volume) a variable which repre-
sents the volume fraction of the injected polymer within the control volume.
Solving alternatively for this variable and the pressure field, and for the tempera-
ture field (using subvolumes in the gap and a streamline-upwinding technique)
provides a reliable and efficient way to simulate the cavity filling, with an easy
representation of all the geometrical details and the delivery system. Normally, at
most one element should be filled per time step, but the VOF method developed
by Voller and Peng [81 ] to simulate RTM shows great promise as to the removal
of this limitation. Finally, the Taylor-Galerkin method developed recently by
Pichelin and Coupez [82] to simulate the 3D free surface flow of a GN fluid
represents a very important step forward obtained by means of an approach
which can be associated with the control volume technique.
The front tracking-mesh adaptation methods are based on mesh deformation or
remeshing. In the former case, the 2D filled zone is mapped onto a fixed domain
976

over which the problem is discretized. There is no longer any free boundary,
with the drawback that it becomes very difficult to address the complexity of
typical industrial parts. Whereas conformal mapping was used in early studies
[44,128a], the Thompson algorithm seems to behave better, see the results
obtained by Gtigeri and co-workers [74,77,98,133]. Particular methods are de-
veloped in [77,98] to couple the 2lAD Hele Shaw model with full 3D calculations
in the singular flow regions (front zone and junctures). On the other hand, the 1st
remeshing algorithms were based on the generation of successive element layers
according to the 2D front progression [52,66,73a], but this technique can hardly
be extended to general shapes. The best approach, as used in the MOLDSYS
software (see the next section) and in [80,104,132,134], consists in tracking the
front(s), recovering as many elements as possible from a fixed mesh covering the
entire midsurface, and generating new elements in the remaining portion of the
filled area.
To close this review, it is worth recalling the various techniques elaborated to
simulate the packing stage [14,56a,60a,61,64,68,70,78,104,132] and to under-
stand the fountain flow goveming mechanisms [48,50,51,54,55,70,77,122,123].

4.2 Description of the MOLDSYS approach


The MOLDSYS software developed by Dupret and co-workers at UCL [9-23,
25,26] (see also Chapter FS) solves the free boundary problem on the basis of the
following concepts : (i)front tracking and automatic remeshing, in order to cap-
ture the shape of the moving free boundary while providing a systematic method
of adapting to geometrical events such as weldline formation or front split after
collision with a side wall; (ii) Eulerian integration of the temperature or any
other 3D field (such as crystallization degree, reaction degree, fiber orientation);
(iii) extrapolation of these fields in the region located between the successive
flow fronts, using an extrapolation mesh and taking the fountain flow conditions
into account; (iv) 2D implicit FEM pressure calculation.
The front region plays a critical role in molding processes in view of the key
effect of fountain flow, which causes fast out-of-plane advection and material
deformation in the front zone. As far as a general purpose algorithm is desired, in
order to simulate injection and compression molding, RIM, RTM .... and predict
crystallization, curing, fiber orientation and their effects on the flow, it is essential
to accurately track the fronts. A remeshing strategy was selected instead of
deforming meshes in order to adapt easily to the topological changes of the filled
domain. Moreover, as in-plane diffusion is neglected in the lubrication approxi-
mation, the simplified energy equation (46) is such that, during any given time
interval, the fountain flow thermal condition (62) has no influence on the tempera-
977

rare evolution upstream of the initial front of the time step, which can be treated
as an outlet section (without thermal boundary condition). The characteristics of
(46) are indeed leaving the domain along this "initial" flow front. On the contra-
ry, the effect of fountain flow must be taken into account in the region located
between the initial and final fronts, since characteristics are leaving or entering
this region across the moving front depending on whether the material points
move faster or slower than the front. The extrapolation mesh is therefore built on
in the inter-front region and equation (62) is considered for the fluid layers that
re-enter the domain after having experienced fountain flow. This method is
general and can be applied to any field for which in-plane diffusion is neglected.
To describe the time-dependent algorithm, it is sufficient to consider a single
time step. Let M n, F n, ~ n denote the mesh, front(s) and discrete domain at
time t,, and [Phi, [T~] the associated column vectors of nodal pressures and
temperatures, while [Vn] and [Wn] stand for the discrete in-plane and out-of-
plane velocity components along the front(s) and inside the domain at the integra-
tion nodes. Additional fields [ A n ] represent the nodal crystallization or curing
degrees, or orientations ... if need be. While the flow fronts move from time t n
to tn+ 1 , they scan the "inter-front region" An+1 . This area is discretized as a
quadrilateral "extrapolation" mesh denoted by En+ 1. T h e sequence of operations
needed to calculate these quantities at time tn+ 1 is organized as follows :
9 Front displacement so as to obtain the extrapolation mesh En+ 1 together with
^

provisional representations of the flow front(s) Fn+ 1 .


9 Tracking of front-front and front-side wall meetings, and front conditioning to
generate the definite front(s) Fn+ 1 . Remeshing to generate Mn+ 1 on ~n+l"
9 Eulerian time integration of the temperature and other 3D fields on the previous
domain soastoobtainprovisionalvalues [7~n+1], ['4n+1] on M n. Extrapo-

lation of [7~n+l], [/~n+l] on the inter-front zone An+1 , taking fountain flow
conditions into account. Correction of the temperature and other fields so as to
obtain I T n +l ] , [an+l] on M n +l .
9 Implicit pressure time integration so as to obtain [Pn+l]" Calculation of in-
plane and out-of-plane velocities [Vn+1] and [Wn+1].

4.2.1 D e t e r m i n a t i o n o f the n e x t c o m p u t a t i o n a l d o m a i n
Front velocities are oriented along the boundary for vertices lying on the side
walls, and along the bisectors at the front for other vertices. The fronts are
978

I |

Figure 14. Extrapolation mesh in the presence of a bifurcation.

moved using the Euler scheme [13]. A more sophisticated technique would be
very difficult to implement since, during the time step, not only can the shape of
the domain evolve, but so can its topology as flow fronts merge, divide, or meet
the boundaries. Also, it was shown by Couniot [13] that the present remeshing
algorithm is of a very low order and that there is no point in devising sophisticat-
ed techniques. The Euler scheme has the benefit that the area scanned by a front
segment is represented by a quadrilateral, possibly degenerating into a triangle.
The union of these elements defines the extrapolation mesh. In the case of thick-
ness discontinuities or bifurcations, as velocities undergo a transformation when
the material points cross the singularity (see Section 3.3.3), each quadrilateral
originating from an upstream facet (relative to the singularity) yields one trans-
formed quadrilateral per downstream facet (Figure 14). Managing the extra-
polation mesh then becomes relatively easy [ 17]. The situation is more complex
regarding front junctions.
Computing the new position of the fronts features the difficulties of computa-
tional geometry [ 13], these stemming from the discrepancy between pure geomet-
ry and its numerical transposition. As a consequence, geometrical algorithms
involve tolerances, which means that the transitivity of geometrical properties is
lost (e.g., two straight lines which are parallel to a third line to a given tolerance
are not necessarily parallel to each other to the same tolerance). The complexity
of these algorithms is therefore often dramatic, particularly in the simulation of
molding processes when a front tracking strategy is selected, since the size of the
element sides must be controlled in order to obtain well-behaved meshes. To
overcome these problems, the solution must be implemented as a mix of numeri-
cal and logical operations. In particular, the present front motion algorithm is
979

Completely filled

[ I Partially filled

/t//tltl Completely empty

Figure 15. Partition of the elements of the fixed mesh.

based on the concepts of "virtual front nodes" and "events" [13], and uses as
reference the fixed mesh. Basically, a virtual node corresponds to the intersection
of one front segment with one side of a mesh element. As the front moves, the
virtual node moves along the element side until it reaches a mesh vertex, which
constitutes a first type of event. Other event types include the crossing of a mesh
boundary or a re-entrant comer along the boundary. All the events are sure to be
detected and simultaneous events can be processed asynchronously (which
consists in removing or inserting virtual nodes in the front description). Hence,
operations are mostly logical and not numerical.

4.2.2 Generation of the temporary mesh used for the next iteration
The discrete front description provides a partition of the elements of the fixed
mesh into three classes, viz. the elements which are (i) completely filled, (ii)
completely empty, and (iii) partially filled (Figure 15). A simple approach would
be to re-use as such the elements of class (i) and to generate additional elements
in the remaining portion of the filled area ~ the so-called generation zone. This
approach is workable with some enhancements. First, the discrete front descrip-
tion must be modified in order to handle front-side wall and front-front meetings,
and filtered to ensure that the local front discretization scale corresponds to the
local fixed mesh scale. Such conditioning can be achieved by redistributing the
front vertices so as to preserve mass balance. As front collisions are treated by
accepting a slight overlap between the colliding fronts, or a slight crossing of the
side walls, it is easy to detect and remove the dead front segments. Front filtering
must also remove the unfilled areas which are regarded as negligible and increase
the filled areas which are still negligible (removing matter can lead to a halt or a
non-natural motion of the fronts). The simulation time is finally adjusted on the
980

basis of a comparison between the total mass integrated at the nozzle and the
mass currently present in the mold.
After front filtering, the temporary mesh is generated. First, the fronts are
located with respect to the fixed mesh and, if strictions are detected, some totally
filled elements are considered as partially filled. Generation is then performed.
As a 1 st step, triangles are built using a Delaunay triangulation that is con-
strained to use the boundary vertices of the generation zone. This method, which
leads to optimal triangulations with respect to some criteria, can be implemented
quite efficiently and extended to curved surfaces. Preserving imposed boundaries
can be achieved in several ways [13]. As a 2 "d step, as many triangles as possible
are merged two by two into quadrilaterals. The temporary mesh Mn+ 1 is built
by merging the filled elements and the generated ones.

4.2.3 Integration of temperature and additional fields


The temperature field is discretized by means of fimte elements along the mid-
surface and cubic splines in the gap [25,26] (full polynomials as in [9,11] are less
stable):
T --- ~, T~i) (z,t) r ) , (96)
i

where the shape functions r are linear or bilinear on the parent element,
while T~i) (z,t) is the temperature profile approximation at node x~ ) .
Space discretization of the energy equation is performed by means of a hybrid
scheme combining a S UPG technique along the midsurface [ 11,25] and colloca-
tion in the gap. Weighted residuals of equation (46) are l~t integrated over f~n,
and are further considered for each node z (k) in the gap. The original SUPG
method [135] consisted in defining the test functions ~i)(xa ) as"

~(i) -- ~)(i) + A'i" v a o3~)(i) /oaxot , (97)

where A~" = (l~'~lA~ +lVo[Ao)/(-~/~Va 2) is an element characteristic time (~


and 7"/ denote local coordinates, with - 1 < ~,7/< 1 for quadrilaterals and
0 < ~ , 0 , ~ + 7/<1 for triangles, while v~ and V,, A~ and A n are character-
istic velocities and lengths at the element centroid). Vanderschuren [11 ] observ-
ed that using the same test functions at each level z <k) greatly reduces the com-
putational cost. Hence, instead of considering AT at the centroid, a position
dependent At is defined while, from (35), velocities are developed in terms of
981

oJ and Op/o3ca . After some calculations, the following expression is found"

Atva = ~i5 + ' (98)

which replaces A~" va in (97). For triangular elements, the following ad hoc
expression was found to take triangular symmetry into account :

Atv a - - 2~]-511 IO~i


(1- q) + (1- r ~i + (r + 7/)i0~ - ~ l ] ~_-q2----_
ll_2oxo
Op aOp "
OX (99)
After in-plane space discretization, the energy equation (46) reads as :

j
-1pCp OZ~ (k OZ
)l ~~mfiOjdS
m o ebp_aT~n+l o~
- I pc; - Oz pCp -~ + va ~i)dS , (100)
~-~m
where T(j) denotes the time derivative of T(j)(z,t). Following Mal et al. [25,
26], cubic spline collocation is performed in the gap by considering (100) for
each node z (t) and expressing the derivatives o32T(j)/oaz2 in terms of the
nodal values of T~j) and the end derivatives 31"~j)/cgz( z - 0 or h, or z - + h ) .
The latter are then eliminated using the wall conditions (67) or (68) and the rela-
tion -kOT(j)/Oz ( z - +h)= + qw (a vanishing derivative is imposed along the
midplane in the synmaetric case). When k is temperature dependent, oqT(j)/0z
is evaluated at internal nodes by a 2nd-order FD approximation. Eventually, the
semi-discrete energy equation takes the genetic form"
[r]- K([rl,[pl...)[rl - Y(VI,[P]-.-) , (101)
where the matrix K (related to heat diffusion) is block-diagonal, each block
being associated with a given midsurface node x~aj) and having dimension m
(the number of nodes in the gap), while the right-hand side includes all the terms
arising from heat advection, friction and compression. Moreover, K is of the
form A-1B, where A and B are block-tridiagonal matrices, while in (101) f
is of the form M-~q, where the mass matrix M consists of m equal blocks
982

whose elements are the integrals ~nm ~((i)~(j)dS.


Eulerian time integration of (101) is carried out in MOLDSYS using the 3rd-
order scheme developed by Dupret and Vanderschuren [9,11], as extended by
Mal [25] to temperature and time dependent matrices K. The method is implicit
for heat diffusion and explicit for other effects, with improved stability for heat
advection. In the absence of diffusion, it reduces to the 4th-order Runge-Kutta
scheme. During integration, the nodal values of p, 'i", va, w and p are
frozen. The thermal penetration inside the walls can be imposed or calculated
(equations (71) or (67)). In the latter case, according to Dheur [16], the best
approach consists in introducing instead of e the unknown X - (Tw - T e)e 2,
which is governed by the equation )~ - 6ks(T w - T e)/pscs.
When abrupt changes of thickness or bifurcations are present, particular tech-
niques must be applied. First, midsurface nodes are doubled, or tripled ... along
the singular line depending on the number of adjacent branches, while the asso-
ciated shape and test functions are split accordingly. Secondly, constraint (62)
must be taken into account. As va and p are frozen during temperature
integration, the relations (72-73) are themselves frozen and the correspondence
between feeding and fed branches and layers can be established before integra-
tion. When a given node is fed with fluid from the adjacent branch(es), the tem-
perature profile must be imposed, from the temperature of the feeding branch(es),
by means of this correspondence (Figure 10). The different branches of the part
can often be arranged in such an order that feeding branches are calculated before
fed branches. In some circumstances, iterations are necessary.
Time integration provides provisional values of [7~n+l] and [An+,] on the old
mesh M , . Extrapolation in the inter-front region An+~ is performed in 2 steps
using the extrapolation mesh En+ ~. First, quantities are evaluated at the nodes of
En+ 1 that belong to the provisional front /~n+~ without considering fountain
flow. This is achieved by least square approximation of the field gradients along
F n . Secondly, the effect of fountain flow is taken into account for those fluid
layers moving slower than the front. Since o9 and p are temperature-profile
dependent (from (31), (36) and (37)), the condition (64) cannot be used directly
to determine the pairs z 1 and z2 governed by condition (62). An iterative
technique is thus implemented. At each node x~) of /~n+l, a l~t guess as to the
profile function ~o(i) is obtained by extrapolation. When, at a given level z~2k) ,
983

the fluid is moving slower than the front \/r


z z" "\o9)')), the temperature is
corrected by picking the temperature of the level z 1 determined by (64). The
profile function is re-constructed and the sequence of operations is repeated up to
convergence.
Unacceptable wiggles were generated in the solution when applying this proce-
dure, due to the thermal shock caused by fountain flow (see Section 3.2.2). After
lengthy investigations [16], the algorithm was modified [25,26] by splitting the
temperature field into a numerical and an analytical part (T* and Ta). The
former is integrated and extrapolated using the above algorithm. The latter is
provided by the thermal shock between two semi-infinite bodies and reads as

+ [ 1 (lO2)
+ (T~,- T f o ) e f f c [ ( h + z ) ( 4 k o ( t - t f ) / P o C - ~ o ) - l / 2 ] ,

where Tw+ and Tw are the upper and lower shock temperatures (equations (69)
and (70)), while t f - t f (x~) is the time at which the front reaches x a , Tfo is
+
the temperature of the core fluid meeting the walls at time tf, and k~, p~, Cpo,
ko, Po and Cpo are the fluid properties at room pressure and temperatures Tw+
and Tw . A single contribution is considered in the symmetric case. The calcula-
tion of Ta requires to track Tfo and tf as functions of x a [25].
As the definite front(s) and mesh Fn+ 1 and Mn+ 1 differ from their provisional
counterparts /~n+l and MnUEn+l, t h e n o d a l v a l u e s o f [7~n+1] and JAn+l]
must be re-evaluated on Mn+l, together with tf and Tfo. This is performed
ypi mg te o,ate va,ues [an+l] on pro
cedure is more stable than least square approximation [16]. When a given node
of Mn+ 1 is located outside f~n u An+l, an appropriate extension of the ele-
ments of En+ 1 might be required [ 16]. In the presence of front-front collisions,
the extrapolated values are averaged in the overlap area and the weights are
obtained from the distances to the previous fronts F n provided by En+ 1 [ 10,16].
The elements of En+ 1 are generally quite flat, especially at the end of a short
shot. Therefore, letting ~l denote the inter-front distance and SXn+1 represent
a characteristic element size of Mn+ 1 in the direction normal to Fn+ 1, the ratio
984

(~l/~Xn+ 1 is often much lower than 1. A relaxation technique was developed by


Dheur [16] to address the resulting discrepancy between theoretical and calcu-
lated temperatures, since in the numerical procedure the fountain flow condition
(62) entirely affects the elements adjacent to Fn+1 (of size 6Xn+l), instead of
the inter-front zone An+1 (of size 51). The "numerical" temperature profile
T(j) (z,t,+ 1) is therefore corrected at any front node x(aj) by the formula

T(j) - (1 - 0) T(j)o + 0 T(j)n , (103)

where T(j)o and T(j)n represent the extrapolated profiles obtained just before
and after applying the condition (62), while the relaxation coefficient 0 is
defined by O-min(51/Sx,,+l,1). The normal characteristic element length
5x,,+1 and inter-front distance 51 are estimated for any node of Fn+1 on the
basis of the new mesh Mn+ 1 and the extrapolation mesh En+1 [16,25]. It
should be noted [16] that the exact value of 0 is not important, since extrapo-
lated values are already influenced by fountain flow before integration. Relaxa-
tion is a numerical stabilization procedure (likewise to upwinding), which is
indispensable at low flow rate.

4.2.4 Calculation of the pressure and velocity fields


The 2D mass equation (42), which governs the pressure field, is space discre-
tized using Galerkin finite elements [9,14,66,136]. Details are unnecessary. The
pressure equation couples the calculations in the part and the delivery system.
Time integration is carried out by means of the implicit Euler scheme, since equa-
tions are stiff in view of the low compressibility of molten polymers. Tempera-
tures are frozen duringintegration, together with ~ , /3, ~, /3B and ?'B (see
(31), (32), (39) and (40)). The system is solved using the Newton scheme.
The last substep is devoted to calculating the nodal velocities where needed.
The front velocity (equations (63) and (35-37)) is discretized by means of 1D
linear elements. A single unknown (the velocity norm) is calculated at each front
node, using a least square technique applied along Fn+ 1. In-plane velocity
components are given by (35-37). Iterations are necessary to determine the zero-
shear rate level z0 in the case of a non-synmaetric flow profile. A specific
drawback of the Hele Shaw approximation is that velocity tends to infinity in the
vicinity of side wall re-entrant comers [20], while viscous heating causes tempe-
rature to tend to infinity at the same time. This is a model, not a numerical,
problem which all the more perturbs the discrete solution as the mesh is refined.
985

Hence, calculating in-plane velocities is difficult and it is advisable to smooth the


result [20]. Finally, w is calculated from (43) via the Galerkin technique. In
practice, (44) is much more stable and generally produces satisfactory results.

5. S I M U L A T I O N R E S U L T S AND E X P E R I M E N T A L VALIDATION

5.1 Results of numerical simulations


Much can be learned by means of numerical simulation from in-depth analysis
of results obtained on simple geometries. In this section, we consider the non-
isothermal filling of a b i f u r c a t i o n composed of three 10 cm x 10 cm square facets
(Figure 16); the cavity features a single point injection gate, located in the comer
of one of the facets ~ which is identified as the upstream facet. To facilitate
simulation, the fluid (a polycarbonate) was assumed to be Newtonian, with con-
stant material properties (r/= 400 Pa.s, p - 1200 kg/m s , k = 0.215 W/m.K and
Cp = 2200 J/kg.K). Hence, thermal calculations were decoupled from the kine-
matics. Details on the material data can be found in [17]. Several alternatives are
examined, featuring an increasing discrepancy in the thicknesses of the down-
stream facets, designed so that the total volume remains constant. The influence
of the geometrical asymmetry of the part is clear, since the thicker downstream
branch gets filled first. Just after it gets filled, the front velocity in the thinner
branch exhibits a sharp increase, as a consequence of flow redistribution.

(a) (b) (c)

Figure 16. Sketch of the test bifurcation and successive flow fronts, for a flow
rate of 60 cc/s. Upstream, and horizontal and vertical downstream branch thick-
nesses : (a) 4/4/4; (b) 4/5/3; (c) 4/6/2 (in mm).
986

35000
p (kPa)

30(0 - /6/2

25000

t = 1.3 s t = 3.52 s
20000

15000
I

10000 , , I , , , , , t(s)

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0


Figure 17. Evolution of the pressure at the spree : (a) 4/4/4 mm thickness
distribution; (b) 4/5/3 mm; (c) 4/6/2 mm.

The pressure evolution at the sprue is shown in Figure 17. As it is observed


quite frequently on complex parts, the crossing of the bifurcation and the
thickness discrepancy have a barely noticeable pressure effect before the end of
the filling of the 1=t downstream facet. The latter event occurs at variable times as
the asymmetry increases. A detailed examination of the velocity field would
reveal that a slight recirculation is induced in the branch that gets filled first, since
the pressure along the bifurcation is higher in front of the injection gate than at the
other bifurcation extremity. Such a recirculation can have an important effect in
the case of fiber-reinforced composite parts.
The bifurcation induces a usually asymmetric downstream repartition of the flow
and, along with the flow, the temperature profile across the gap is split. As a
consequence, the downstream temperature and velocity profiles are asymmetric
(see Section 3.2.3). The four phenomena governing temperature evolution in

Figure 18 (next pages). Layer by layer temperature field at the end of the filling,
from T = 400 K (black) to T = 600 K (white) (approximate range). The bifur-
cation (in bold) is located at the top of the upstream branch and the bottom of the
2 downstream branches. Flow rates : (a) 60 cc/s and (b) 120 cc/s.
987
988
989

injection molding (fountain flow, heat advection, viscous heating and heat dif-
fusion) do not occur at the same time. Fountain flow occurs first, when the poly-
mer reaches a given point in the cavity. Its main effect is to make the temperature
profile quite uniform. However, this situation is short-lived at a given Eulerian
location, in view of the effect of heat transport (which generates asytmnetric pro-
files far downstream from the bifurcation) and the viscous dissipation taking place
in the sheared fluid layers. Soon the temperature profile starts exhibiting a well-
known "seagull" shape. After a while, heat diffusion starts exerting an increasing
cooling influence, especially when the local flow rate is low.
Figure 18(a) shows the temperature field at the end of the filling, in different
layers across the thickness, each identified by its relative distance to the midsur-
face. In the core (layers 3 to 5), temperature is relatively uniform and has a value
close to injection temperature. Only at very few locations is this observation not
valid, viz. in stagnation areas (visible at the comer in layers 2 and 6 of the up-
stream branch) and in the region filled last, where the temperature is higher due to
viscous heating (see layer 3 of the vertical branch). On the other hand, near to
the skin, where the cold mold effect is higher, temperature distribution appears
more or less symmetric in the downstream branches, except in the bifurcation
vicinity. This is easily understood in view of the flow and temperature profiles
splitting at the bifurcation, where the inner skin layers are fed with a much hotter
material than on the outer side. In fact, the asymmetry tends to increase when the
front moves away, but it falls in competition with viscous heating or cooling.
Finally, between the core and the skin, a noticeable effect is the temperature
increase induced by viscous heating in layers 2, 3, 5, 6 of the thin vertical branch,
as a consequence of the sharp increase in velocity occurring after the thicker
branch has been filled. Also, the temperature profile appears to be highly non-
symmetric in these layers, due to the kinematic bifurcation. This effect is also
visible in the horizontal downstream branch (layers 2, 3, 5, 6). All these observa-
tions are confirmed by results obtained for other flow rates (Figure 18(b)).

5.2 Experimental validations


Experimental validation has been a constant concern since the beginning of
injection molding numerical modelling. As the melt flow is by nature unsteady,
non-isothermal and 3D, while the fluid is compressible and non-Newtonian, few
analytical solutions exist and their use is highly restricted. Therefore, laboratory
experiments are necessary to validate the models. In a 1~t step, experimental flow
fronts and gate pressure measurements in 1D filling were compared to numerical
predictions [35b,40a,42]. In the early 80's, the developmem of more sophisticat-
ed tools allowed to predict and validate the 2D front shape and pressure evolution
990

Sprue Pressure
(a)
Heat flux
Optical fibers
r Gate / Pressure
--- ~ ~ _ _ - ~ ~ . . ~ e a t flux

pos. 5
pos. 4
pos. 3
pos. 2
pos. 1
after
gate gate
(b) Ill /
Ill I
-111 I
57 -la 51
qi
qH \
9 qll I

17 365
Figure 19. (a) Sketch of the injected plate with a close-up of the gate and the
locations of the sensors. (b) Finite element mesh representation (where use is
made of the symmetry of the part), together with the thickness distribution.

in the cavity [43,46,66]. Later, important steps consisted in introducing accurate


characterization of the viscosity and other material properties, while extending the
model to the process packing and cooling stages [52,56b]. Nowadays, experi-
mental validation focuses on the prediction of the part's final properties (such as
density, residual stresses, shrinkage and warpage), in some cases taking visco-
elastic effects into account [60b,64,132,137]. It is clear that precise on-line mea-
surements and material characterization, together with rigorous numerical tools,
are required to obtain accurate validations. As an example, injection molding
experiments were performed in the laboratories, and with the help and expertise
of the staff of the "P61e Europ6en de Plasturgie" of Oyonnax (France), using a
420-ton hydraulic Billion press. The mold cavity was a 367x102x4 mm rectan-
gular plate (Figure 19). In order to ensure linear filling, the gate was designed
991

311
309
~ 307
305
~ 303
301
299
297
295 .....................................
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
time (s)
Figure 20. Filling of a rectangular PS plate. Comparison between predicted ( . )
and measured (---) mold surface temperatures for the first sensor in the cavity.

with a large rib. The delivery system consisted of a 11 cm sprue and a 32.5 cm
trapezoidal nmner. The steel mold temperature was regulated by water circula-
tion in channels situated 10 mm inside the walls. The mold was equipped with
transducers and optical fiber sensors. Pressure transducers were positioned at the
inlet of the sprue, just upstream and downstream from the linear gate, and in the
plane of symmetry of the cavity (40, 115, 189 and 264 mm downstream from the
inlet). Heat flux transducers consisted of pairs of thermocouples separated by a
steel cylinder (one positioned at the mold surface and the other 7 mm inside the
mold), which could produce a very fast response time of 0.2 ms with a precision
of 1 K. Heat flux transducers were placed at the inlet of the runner, just upstream
and downstream from the gate, and in the cavity at distances of 40, 115, 189, 264
and 319 mm downstream from the inlet, and 20 mm from the plane of synmaetry.
Although different polymers were injected, only the results obtained with PS
(Dow Styron 678E, for which the material parameters are given in [20,25,116]),
will be detailed. The injection and mold temperatures were 483 K and 294 K,
respectively, while the injection flow rate was 250 cm3/s. The evolution of the
mold surface temperature is presented in Figure 20. As filling progresses, the
temperature at the sprue rises first, followed by all the sensors. The increase is
sudden and accurately records the front motion. After the front has passed, the
polymer-wall interface temperature remains approximately constant during the
remainder of the filling. This effect, which is a direct consequence of the thermal
shock theory presented in Section 3.2.2, was obtained in both the numerical
predictions and the experiments, and for materials as different as PS, PC, PP and
992

Figure 21. Successive flow fronts obtained from simulation with a very simple
runner and gate model (top), and from short shot experiments (bottom).

PET, thereby validating the thermal model of Mal et al. [25,26]. Also, this shows
that in-plane heat conduction within the polymer and the walls is negligible,
whereas the effect of the thermal singularity caused by fountain flow is important.
Short shot experiments were performed in order to visualize the flow front pat-
tern (Figure 21). It is obvious that the gate, which was designed to ensure linear
injection into the cavity, is not efficient. The polymer first enters the cavity cen-
ter, before reaching the lateral extremities of the thick rib. Later on, a linear front
forms, while the polymer flows faster close to the lateral side walls. This finally
leads to a 'V'front shape, which grows when the flow rate is increased. Verho-
yen [20] has proved that this effect, which sometimes generates a front weldline
at the end of the filling and was observed with all other injected polymers, is a
direct consequence of the non-uniform viscous dissipation taking place in the
runner during filling, leading to a non-uniform and transient gate temperature
profile. The resulting non-uniform fluidity in the cavity is responsible for the flow
front 'V' shape (Figure 21).
-.
993

6. EXTENSIONS OF THE M O D E L

6.1 Chemical reactions


In Thermoplastics Injection Molding (TIM), solidification results from the
cooling, below the glass transition temperature Tg, of the injected polymer in
contact with the cold cavity walls. In reaction and thermoset injection molding
(cf. Macosko [5], Kamal and Ryan [6], and the references therein), the irreversi-
ble curing reaction of a resin is used to build the solid stiff part. While reaction
injection molding usually refers to mixing activated systems (such as polyure-
thanes or epoxies), thermoset injection molding refers to heat activated systems
(such as polyesters). In this section however, the terminology of RIM is used in
the general sense. The advantages of RIM compared to TIM are numerous. The
low viscosity of thermosetting resins before curing allows one to have high fiber
reinforcement contents, as needed for good mechanical properties, and also the
injection of large parts (of several m 2) on classical machines. In addition, the
solidification resulting from curing is easily controlled by changing the resin com-
position and using catalysts or inhibitors.
The same modelling approach used in TIM may be used to analyze RIM
provided some particularities are taken into account. First, additional thermody-
namic variables must be introduced to represent the evolution of the resin compo-
sition. Each material property of the melt becomes a function of p, T, possibly
~', and of the different species mass fractions w i . Neglecting species diffusion,
each toi is governed by an equation of the form Dooi/Dt-fi(p,T,y,to~),
where fi is the kinetics of species production. Secondly, as curing reactions are
exothermic, reaction heat must be incorporated in the energy equation. Hence,
the numerical procedure presented in Section 4.2 may easily be extended to
simulate RIM processes. To this end, T and toi are solved simultaneously,
thereby ensuring stability in this strongly coupled problem [25].
The modelling of RIM has been the object of constant attention in the literature.
While simplified models were used in earlier studies [83], a major concern has
been to understand the space-time distributions induced by fountain flow and
their effect on the curing reaction. This problem was intensively investigated by
Macosko and co-workers [4,32,49, 58], see also [84]. Other modelling issues
relate to the effect of inertia when the viscosity of the injected polymer is very
low [85], or to the molding of highly viscous reactive polymers [72,86].
A precise determination of the material properties of the reactive mixture is of
primary importance for accurate simulation [3,5,6]. While the influence of the
reaction on p and k can often be neglected, its effect on 7/ must generally be
994

(a) z = h 0.5 z - 0 (b)

Figure 22. Reaction injection molding of an automobile front hood (of lm x lm


x 3mm). Distribution of the degree of cure and temperature at end of filling (after
17 s) (a) on the wall; (b) along the midplane of the cavity. (From [25]).

taken into account, especially in the vicinity of the so-called gel point where r/
rapidly tends to infinity. Even more critical is the accuracy of the model used for
reaction kinetics. When the resin formula may be considered fixed, simple n th-
order models can often be used, and all the information on the reaction is lumped
into a single variable called the degree of cure. When the influence of a modifica-
tion in the resin composition is to be investigated, mechanistic models are neces-
sary. A most interesting and widely used example of this approach is the model
developed by Stevenson [138] for unsaturated polyester, where the influences of
initiator and inhibitor concentrations and types on the reaction are considered.
This model is used in the example presented in Figure 22 (Mal [25]). In order to
reduce computational time, use is made of the part symmetry in the simulations.
The injection temperature is 356 K, the heating channels being kept at 423 K in
order to activate the curing reaction.

6.2 P o l y m e r crystallization
For semi-crystalline polymers, the thermo-mechanical conditions prevailing
during the different stages of the process interact with the development of micro-
995

structures that in turn affect the product end-use properties [139-147]. Polymer
crystallization has been extensively studied and several kinetic models developed
[24,148-151], most of them based on the Kolmogorov-Avrami-Evans approach
[152,153]. However, little work has been performed to introduce these models
into injection molding simulation. Dufoss6 [154] investigated the effect of non-
isothermality on PE molding. The effect of shearing on PPS crystallization was
analyzed by Hsiung et al. [87,91 ], who extended the Hele Shaw model to include
crystallization by means of the Nakamura model [155]. More recently, Isayev et
al. [88] simulated PP molding, taking shearing influence into account by combin-
ing the models of Liedauer et al [156] and Nakamura et al. [155]. The combined
effects of temperature, shear and pressure were analyzed by Ito et al. [92] and
Verhoyen et al. [20,93], in order to predict PP and PET molding.
Crystallization models can be implemented in Hele Shaw simulation by using
the scheme of Section 4.2. This requires (i) to introduce a kinetics model in order
to integrate the crystallization degree a v as an additional 3D field, and (ii) to
consider the effect of crystallization on r/, k, p and H. The goal of the
present section is to summarize the investigations carried out by Verhoyen et al.
[20,93], showing that combining crystallization modelling and simulation pro-
vides a very convenient methodology to better understand the crystallization
mechanisms under high pressure and shear conditions.

6.2.1 The crystallization model


The model developed in [20,93] combines the consecutive Avrami model [19,
24] with the approach of Schneider et al. [157] and Andreucci et al. [158]. The
former is an extension of the Avrami theory, which takes into accotmt induction
time, primary and secondary crystallization, ultimate degree of crystallinity, and
the effects of temperature and shear stress. The latter allows to separate the
nucleation and crystal growth processes. In the sequel, subscripts 1 and 2 refer
r
to the primary and secondary mechanisms, while avi denotes the relative degree
of crystallinity of mechanism (i). The kinetics model reads as

DO1 -- 1 / t ~ , 002 -lit 2 , (104)


Dt Dt

I DN1 = 1V DR1 = G N 1 if 0 1 > 1 , else DR1 = O ,


Dt ' Dt Dt
(105)
DS1 = 8 ~ G R 1 , DV1 = G S 1 ,
LN Dt
996

)
9
|
Q
9 9 "
(a) ( ~':
(b)

Figure 23. Sketch of the crystallization model. Interactions affect the nucleation
and growth of real spherulites, but have no effect on the fictive crystals.

Da~Cl c DV~
Dt = ( 1 - a V l ) Dt
DavC2
: (, )(
Dt
(106)
if 02>1 , else D C-aw = O,
Dt
Da v - Otv ( Wl DavlC + ~ c)
Dt Dt
(1- W1 j Dav2
Dt

The parameters t oi represent the isothermal induction times, while non-


isothermal induction times are calculated from the accumulated factors 0i [159],
which reach 1 when induction is complete and the associated mechanism starts.
Equations (105) refer to the model of [157,158] : the time derivative of the
specific volume VI of the fictive nuclei is the product of their specific surface
S 1 by the growth rate G, while DS1/Dt is 4~r times the product of the
fictive specific diameter 2R 1 by G, and DR 1/Dt is the product of the specific
number of fictive nuclei N 1 by G; the evolution of N 1 is governed by the
nucleation rate /~ (Figure 23). As the fictive specific volume (V1) is also the
fictive crystallinity degree, the Avrami postulate provides the relative crystallinity
C
degree Ctvl of the first mechanism. The relative degree for secondary
C
crystallization av2 is calculated by means of the differential form of the Avrami
equation, with k 2 and n 2 as kinetic parameter and Avrami index. The real
degree of crystallinity is calculated by combining the 1st and 2 nd mechanisms
997

(with the weight W 1 ), taking into account the asymptotic degree of crystallinity
oo

tzv . The model of [157,158] was not used for the secondary mechanism, whose
physical behavior is totally different. It is assumed that nuclei pre-exist in the
melt, and that, in static conditions, the number of activated nuclei /V1 is
independent of time (with instantaneous nucleation). Hence, /V1 is a given
function of T, ~' and p, which can be determined experimentally. When the
material undergoes a complex thermo-mechanical history, /V is obtained from
the relation /V - max (D1Vm(t)/Dt, 0), in order to avoid any decrease of N 1
when the polymer is heated. This expression should be improved.
This model can predict PET crystallization in typical laboratory apparatus (such
as DSC or cone and plate rheometer) in both isothermal and non-isothermal
conditions [20,93]. In principle, the dependence of PET material properties on
txv should be specified. Detailed material functions are given in [20]. However,
these effects may be neglected in PET injection molding since crystallization
develops mainly during the packing and cooling stages.

6.2.2 Experimental and numerical results


Injection molding experiments were performed using the equipment described in
Section 5.2. The half-thickness of the plate was 3 ram. Processing conditions are
detailed in [20]. The simulation was peformed by imposing a gate flow rate of 60
cm3/s during filling, while the experimental pressure evolution was prescribed
during packing and cooling until the gate was frozen. A non-uniform wall tempe-
rature (from 293 K to 303 K) was imposed, corresponding to wall sensor mea-
surements performed at the beginning of the filling stage.
The results obtained by prescribing a uniform inlet temperature of 568 K are
depicted in Figures 24(a) and 24(b), where the predicted and experimental crys-
tallinity profiles are represented at two locations in the plate, together with the
profiles of N 1 . The results obtained by taking only thermal effects into account,
and only thermal and shear effects, are shown for comparison purposes. In all
cases, a zero a v is predicted along the walls, indicating a quench of the polymer
when it meets the walls, while a uniform crystalline core is obtained only if pres-
sure-induced crystallization is taken into account (when only the thermal effect is
considered, the core crystallinity reaches 3 %, which is far from the experimental
28%; when the shear stress effect is added, no significant difference is observed).
However, the predicted and experimental crystallinity profiles do not fit well in
Figure 24(a). As explained in Section 5.2, this results from the important tempe-
rature increase observed in an annular region located at the inlet of the gate and
998

0.6 i i "' I I '" I i 1020


(a)

II ~ ~ . . . . . . . . . . ~ cr
0.4 .............................. ..-.................. x .............. :.....-.........::...........................
"~ rf/ k1~6l ~O: , 0

-~ 0.3 ..... -"-~176 -""--~ \~1014 ~


0.2 ~.
r,.) 0.1 I ~176
~ ....- ,L. ! % .. %e ~ "] 1012 :tt:
__.~ _=_- : ...............k ........... : - - ~..o %LA
0.0 ~----'~""i'~"-'" I I I .......~ - . . ~ . % - ~ l 10 lo
-3 -2 -1 0 1 2 3

0.6 . . . . . . 1020
(b)
z
0.5 1018
cur
0.4
"~ 101s O

"9 0.3
o ,,,,,~

b
o.2 -/" 1014

1012
~,,~.

ro 0.1 - I" ?~1


o~ .-- -.:.-:..:...-...~...(....-..-'=.....':...=...--" - \ v4 c_.._l

0.0 t.7 .... _:..r ............ i i i ...............t..':..___.. L/ 101~


_1 -2 -1 0 1 2 3
0.6 102o
(c)
0.5 Z
10 TM
exl)
0.4
1016 0

t i
, ,...r
1014 c~
r~ 0.2 i,,,~ 9

b 1012
0.1 L _ ~ ~ r % -
L--..-I

0.0 I ~ - I ...... I..... I ~ 101~


-3 -2 -1 0 1 2 3
z [mm]
Figure 24. Gapwise crystallinity profiles at distances of 55 mm (a and c) and 310
mm (b) from the gate in the symmetry plane. Comparison between experiments
(e) and numerical predictions with (a and b) uniform and (c) improved inlet tem-
perature profiles. Dotted lines" only thermal effects considered; dashed lines"
only thermal and shear effects; solid lines : all effects. (From [20]).
999

due to viscous heating in the nmner. Since nuclei are generated mainly upstream
from the delivery system (i.e. where p is the highest), they are partially destroy-
ed in the nmner under the action of the high peripheral temperature. Hence, their
inlet distribution is not uniform, and this strongly affects crystallization. Also, the
non-uniform inlet temperature profile again had a major impact on the front
shape. A simple model was developed to evaluate the distribution of nuclei
downstream from the gate [20]. The resulting profiles are shown in Figure 24(c),
which exhibits a much better agreement between calculation and experiments.

6.3 Filling of fiber mats


SRIM and RTM differ fundamentally from classical injection molding. Usually,
in these processes, a thermosetting resin is injected into a cavity where a fiber re-
inforcement has been placed beforehand [7]. Using SRIM or RTM, it is possible
to achieve, in one shot, complex composites with a high ratio of continuous fibers
and excellent mechanical properties, while the low resin viscosity (< 0.1 Pa.s
typically) allows the manufacturing of large parts (of several m 2) with low cost
tools. In addition, good orientation control is obtained.
Although the reinforcement considerably modifies the physics, the same simula-
tion tool developed for TIM and RIM may be used to simulate SRIM and RTM
provided some particularities are taken into account [21,25,80,96,97,100,133,
160-163]. The equations must be established via a multiphase analysis, in which
the liquid flows through a solid porous medium (cf. Tucker and Dessenberger [8]
and the references therein). For an incompressible Newtonian resin, stationary
fiber reinforcement, negligible inertia and long range viscous forces, the resin
flow is governed by the Darcy law, which relates the intrinsic fluid velocity v i to
the pressure gradient O~/OqXi ,

v i = - / ~ ~ Kij Op/Oxj , (107)

where q~ is the porosity (i.e. the liquid volume fraction), 7"/ is the resin vis-
cosity and Kij the permeability tensor. The Darcy law is useless in describing
how the resin penetrates the bundles and wets each individual fiber.
As in TIM and RIM, most SRIM or RTM parts are thin (see Section 3.1). Di-
mensional analysis thus leads to combining the mass and Darcy equations as

c~xa
1000

h
where S ao - I / ~ KaO dz . (109)
-h

Greek indices mean that the quantity is limited to its in-plane components.
Equation (109) has the same form as (42), the fluidity factor S being replaced
by a fluidity tensor Sag. The procedure outlined in Section 4.2 is thus applic-
able, provided minor modifications of the front tracking algorithm in singular
regions are introduced [21,25].
The in-plane permeability Ka# is generally determined by eigenvector-eigen-
value measurements in simple flows [7,97]. In addition, progress has been made
in relating the reinforcement geometry and draping with Ka~ (taking into
account in some cases the elasticity of the preform) [164-170]. To illustrate the
effect of permeability, the injection of an automobile front hood has been simulat-
ed using two different kinds of reinforcement (Figure 25). The filling time is
141 s. Crosses indicate where vents should be placed. This result shows that
permeability is a key factor in determining the geometrical and operating parame-
ters for a given part. It should be noted that, as perfect draping along the borders
or the junctions is difficult, high permeability zones are often found, resulting in
easy flow channels or race tracking [162,171-173].
In SRIM and RTM, solidification generally results from a chemical reaction
initiated by heat or mixing during filling. In order to reduce the cycle time, the
mold walls are set to a high temperature so as to enhance this mechanism; hence,
thermal effects must be taken into account in the model [ 100,102]. However, the
heat transfer in the porous medium is much more complex than in Section 3.1,
since hydrodynamic dispersion must be considered due to the tortuous nature of
the flow paths. Assuming local thermal equilibrium between the fiber bundles
and the resin, the energy equation [99,101] reads as :

(op, c, p,c, v,

O [Kij+Kii~o~I" 1 ap @ (11o)

where P ph and Cph denote the specific mass and specific heat (with subscripts
l and s referring to the liquid and the solid), Ki~ and Kia are the thermal
conductivity and mechanical dispersion tensors, and ~ is the specific heat
source due to chemical reactions. In (1 lO), K~ is a function of the conductivity
1001

(a) (b)

(c) (d)
Figure 25. Isothermal simulation of the filling of an automobile front hood" (a)
finite element mesh; (b) example of temporary mesh; (c and d) evolutions of the
flow front for a non-isotropic (Kxx/Kyy - 3 ) and an isotropic reinforcement.

of the two phases and the reinforcement geometry [21,174], while KiJ, which
formally accounts for the mixing convection heat transfer, is often considered to
be proportional to Uv,II
Mal et al. [21] propose the equation"

Kid =" ~Pl Cl IIVml1-1 Zijkl Vk Vl , (111)

where Lijkt is a fourth order tensor. This relation can induce different mixing
effects parallel or perpendicular to the flow, in agreement with experimental
observations [175]. Also, it is observed that hydrodynamic dispersion is invariant
with respect to time scale changes, corresponding to a purely geometrical mixing
model. As for Kij, the symmetries required for Lijkl are not clear. A complex
dimensional analysis shows that, in SRIM and RTM, gapwise mixing is
1002

t = 70s 31530~ t = 70s 305 ,,,-~


(end) 32.5~ (end~ 315
335~/'~ (a) " " 325/'~~ (b)

3 4 ~

305 305
t = 141s ,,,,315~-,,7"~ t = 141s ...315~ 5>
(end) 33~.~@r" (c) (end) 33~~, / (d)
3 4 ~ 345~~~~,,~

Figure 26. Simulation of the non-isothermal filling of an automobile front hood.


Average temperatures (in K) at end of filling. Top : filling time = 70 s; bottom:
filling time = 141 s; (a, c) no dispersion; (b, d) Lzz = 0.0017 mm. (From [21]).

predominant when compared with in-plane contributions, and tijkl V k. v t thus


has a single non-negligible component Lzzal3 v a v/3. ff it is further assumed that
Lzzoc~ is transverse isotropic [21] ( L z z a o - L z z tSaO), the equations involve a
single mixing coefficient Lzz and are equivalent to the model of Tucker and
Dessenberger [8].
Thanks to the material data provided in [99], realistic numerical experiments
may be carded out. In particular, the non-isothermal filling of the automobile
front hood of Figure 25 has been simulated by Mal et al. [21 ] in order to compare
1003

the results obtained with and without neglecting mechanical dispersion (Figu-
re 26). Even though Lzz is small, its effect on the average temperature distribu-
tion leads to a difference of up to 10 K in the entry zone when filling is faster. It
is essentially in the immediate vicinity of the gate that its effect is important,
because mechanical mixing is very effective in the regions where thermal
gradients and velocities are high. A complex balance between dispersive and
conductive heat transfer takes place during filling and, the higher the flow rate,
the less effective conductive heat transfer is (while mixing acts on higher thermal
gradients). In conclusion, the global heat transfer is increased in the vicinity of
the gate, while it is reduced in regions of lower velocities.
Two closing remarks are in order. First, chemical reactions can be modelled us-
ing the methodology of Sections 4.2 and 6.1, as long as the kinetics and material
data are known. The evolution equation of species concentration is similar to
(110), with an additional term relating to curing kinetics, while molecular dif-
fusion can generally be neglected and mechanical dispersion is governed by the
tensor Lijkt. Secondly, it is easy to extend to SRIM and RTM the theory of
abrupt changes of thickness and bifurcations developed in Sections 3.2.3 and
3.3.3. In the case of isothermal flow, tmiform material properties and a trans-
verse isotropic permeability, mass conservation exactly provides (81), while
pressure continuity across the singular line forces the continuity of the tangent
velocity (v2f +- v2f-), instead of (82) (see Figure 12). At the front, equation
(83) remains valid and thus, after a few calculations, (81) and (83) show that,
when h + ~ h - , the normal velocity component vanishes at the flow front
+
(Vlf -Vlf -0). This simple effect, which holds true in the case of bifurca-
tions, can be observed in practice. It can also be exploited for computational
purposes.

7. CONCLUDING REMARKS AND A C K N O W L E D G M E N T S

This chapter summarizes the models and numerical techniques that were elabo-
rated to develop the MOLDSYS simulation software. Combining sound physical
knowledge with rigorous mathematical approximations and accurate algorithms
proves efficient in addressing the main issues of process modelling. Various
polymer molding techniques, all based on the thin cavity assumption, have been
implemented in the simulation program. The numerical method, which is based
on front tracking, automatic remeshing and extrapolation (taking fountain flow
into account), has been successfully extended to the filling of complex parts.
1004

Results prove the validity of this approach. Also, it is shown that the simulation
tool can be used to better understand the tmderlying physics when complex
materials are molded.
Besides the authors of this chapter, several people participated in this research,
including Marcel Crochet, Luc Dheur, Olivier Hansen, Kali Kabanemi, Natasha
Van Rutten and Vincent Verleye, whom the authors wish to thank here for their
contributions. The work was carded out within the framework of the European
BRITE project RI1B-0087-F(CD), the "Multimat6riaux" project of the Walloon
Region of Belgium, and the program of Interuniversity Attraction Poles of the
Belgian state, and in collaboration with the Shell Research and Technology
Center in Amsterdam (the Netherlands) and the "P61e Europ6en de Plasturgie" of
Oyonnax (France). The experimental work was possible thanks to Rapha61
Favier's permission. Grants from the IRSIA (Belgium) and the FRIA (French
Commtmity of Belgium) are acknowledged. The authors wish to thank Jean-
Pierre Gazonnet, G6rard Dechavanne and Virginie Durand for their friendly help
and advice in performing the molding experiments. The efficient page setting
work of Victor Vermeulen was also appreciated.

REFERENCES

1. Z. Tadmor and C.G. Gogos, Principles of Polymer Processing, John Wiley &
Sons, 1979.
2. J.R.A. Pearson, Mechanics of Polymer Processing, Elsevier, London, 1985.
3. C.L. Tucker, in : A.I. Isayev (ed.), Injection and Compression Molding
Fundamentals, Marcel Dekker, New York, 1987.
4. T.A. Osswald and S.C. Tseng, in : S.G. Advani (ed.), Flow and Rheology in
Polymer Composites Manufacturing, Elsevier, Amsterdam, 1994.
5. C.W. Macosko, RIM, Fundamentals of Reaction Injection Molding, Hanser,
Munich, 1989.
6. M.R. Kamal and M.R. Ryan, in : A.I. Isayev (ed.), Injection and
Compression Molding Fundamentals, Marcel Dekker, New York, 1987.
7. C.D. Rudd, A.C. Long, K.N. Kendall and C.G.E. Mangin, Liquid Moulding
Technologies, Woodhead publ., Cambridge, 1997.
8. C.L. Tucker and R.B. Dessenberger, in 9 S.G. Advani (ed.), Flow and
Rheology in Polymer Composites Manufacturing, Elsevier, Amsterdam, 1994.
9. F. Dupret and L. Vanderschuren, AIChE J., 34 (1988) 1959.
10. A. Couniot, L. Dheur, O. Hansen and F. Dupret, Proc. Numiform'89,
Balkema, Rotterdam (1989) 235.
11. L. Vanderschuren, Ph.D. Thesis, Universit6 catholique de Louvain, 1989.
1005

12. F. Dupret and L. Dheur, Proc. Oji Int. Seminar, Tomakomai, Japan (1990),
Hemisphere (1992) 583.
13. A. Couniot, Ph.D. Thesis, Universit6 catholique de Louvain, 1991.
14. K.K. Kabanemi and F. Dupret, Proc. Numiform'92, Balkema, Rotterdam
(1992) 357.
15. K.K. Kabanemi, Ph.D. Thesis, Universit6 catholique de Louvain, 1992.
16. L. Dheur, Ph.D. Thesis, Universit6 catholique de Louvain, 1992.
17. A. Couniot, L. Dheur and F. Dupret, IMA Conf. Series, New Series Nr. 42,
Clarendon Press, Oxford (1993), 381.
18. V. Verleye, Ph.D. Thesis, Universit6 catholique de Louvain, 1995.
19. O. Verhoyen, F. Dupret and R. Legras, Proc. Numiform'95, Balkema,
Rotterdam (1995) 1209.
20. O. Verhoyen, Ph.D. Thesis, Universit6 catholique de Louvain, 1997.
21. O. Mal, A. Couniot and F. Dupret, Composites Part A, 29A (1998) 189.
22. N. Van Rutten and F. Dupret, Proc. ASME'97, Dallas, Texas, MD-Vol. 79
(1997) 399.
23. O. Verhoyen and F. Dupret, J. Non-Newtonian Fluid Mech., in press (1998).
24. O. Verhoyen, F. Dupret and R. Legras, Polym. Eng. Sci., in press (1998).
25. O. Mal, Ph.D. Thesis, Universit6 catholique de Louvain, 1998.
26. O. Mal, L. Dheur, P. Pirotte, N. Van Rutten, A. Couniot and F. Dupret, Proc.
Numiform'95, Balkema, Rotterdam (1995) 1165.
27. Z. Tadmor, J. Appl. Polym. Sci., 18 (1974) 1753.
28. H. Janeschitz-Kriegl, Rheol. Acta, 16 (1977) 327; 18 (1979) 693.
29. W. Dietz, J.L. White and E.S. Clark, Polym. Eng. Sci., 18 (1978) 273.
30. A.I. Isayev and C.A. Hieber, Rheol. Acta, 19 (1980) 168.
31. H. Van Wijngaarden, J.F. Dijksman and P. Wesseling, J. Non-Newtonian
Fluid Mech., 11 (1982) 175.
32. J.M. Castro and C.W. Macosko, AIChE J., 28 (1982) 250.
33. S.M. Richardson, Rheol. Acta, 22 (1983) 223.
34. C.L. Tucker and F. Folgar, Polym. Eng. Sci., 23 (1983) 69.
35. M.R. Kamal and S. Kenig, Polym. Eng. Sci., 12 (1972) 294; 302.
36. J. Vlachopoulos and C.K.J. Keung, AIChE J., 18 (1972) 1272.
37. J.L. Berger and C.G. Gogos, Polym. Eng. Sci., 13 (1973) 102.
38. P.-C. Wu, C.F. Huang and C.G. Gogos, Polym. Eng. Sci., 14 (1974) 223.
39. Z. Tadmor, E. Broyer and C. Gutfinger, Polym. Eng. Sci., 15 (1975) 515.
40. G. Williams and H.A. Lord, Polym. Eng. Sci., 15 (1975) 553; 569.
41. E. Broyer, C. Gutfinger and Z. Tadmor, Trans. Soc. Rheol., 19 (1975) 423.
42. Y. Kuo and M.R. Kamal, AIChE J., 22 (1976) 661.
43. W.L. Krueger and Z. Tadmor, Polym. Eng. Sci., 20 (1980) 426.
1006

44. M.E. Ryan and T.-S. Chung, Polym. Eng. Sci., 20 (1980) 642.
45. C.G. Gogos, C.-F. Huang and L.R. Schmidt, Polym. Eng. Sci., 26 (1986)
1457.
46. A. Couniot and M.J. Crochet, Proc. Numiform'86, Balkema, Rotterdam
(1986) 165.
47. M. Grmela, Polym. Eng. Sci., 24 (1984) 673.
48. H. Mavridis, A.N. Hrymak and J. Vlachopoulos, Polym. Eng. Sci., 26 (1986)
449.
49. I. Manas-Zloczower, J.W. Blake and C.W. Macosko, Polym. Eng. Sci., 27
(1987) 1229.
50. D.J. Coyle, J.W. Blake and C.W. Macosko, AIChE J., 33 (1987) 1168.
51. R.A. Behrens, M.J. Crochet, C.D. Denson and A.B. Metzner, AIChE J., 33
(1987) 1178.
52. C.A. Hieber, in : A.I. Isayev (ed.), Injection and Compression Molding
Fundamentals, Marcel Dekker, New York, 1987.
53. N. Famili and A.I. Isayev, in : A.I. Isayev (ed.), Modeling of Polymer
Processing, Recent Developments, Hanser, Munich, 1991.
54. H. Mavridis, A.N. Hrymak and J. Vlachopoulos, J. Rheology, 32 (1988) 639.
55. H. Mavridis, A.N. Hrymak and J. Vlachopoulos, AIChE J., 34 (1988) 403.
56. H.H. Chiang, C.A. Hieber and K.K. Wang, Polym. Eng. Sci., 31 (1991) 116;
125.
57. F.P.T. Baaijens, Rheol. Acta, 30 (1991) 284.
58. M.A. Garcia, C.W. Macosko, S. Subbiah and S.I. Gii~eri, Intem. Polym.
Proc., 6 (1991) 73.
59. C.A. Hieber and H.H. Chiang, Polym. Eng. Sci., 32 (1992) 931.
60. A.A.M. Flaman, Polym. Eng. Sci., 33 (1993) 193; 202.
61. K.T. Nguyen and M.R. Kamal, Polym. Eng. Sci., 33 (1993) 665.
62. K.K. Kabanemi, A. Mt-Kadi and P.A. Tanguy, Rheol. Acta, 34 (1995) 97.
63. R.-Y. Chang and S.-Y. Chiou, Polym. Eng. Sci., 35 (1995) 1733.
64. L.F.A. Douven, F.P.T. Baaijens and H.E.H. Meijer, Prog. Polym. Sci., 20
(1995) 403.
65. S.C. Chen and Y.C. Chen, J. Appl. Polym. Sci., 55 (1995) 1757.
66. C.A. Hieber and S.F. Shen, J. Non-Newtonian Fluid Mech., 7 (1980) 1.
67. S.-F. Shen, Int. J. Num. Meth. in Fluids, 4 (1984) 171.
68. T.-S. Chung, Polym. Eng. Sci., 25 (1985) 772.
69. H. Alles, S. Philipon, J.-F. Agassant, M. Vincent, G. Dehay and P. Lerebours,
Polym. Proc. Eng., 4 (1986) 71.
70. M.R. Kamal, in :L.T. Manzione (ed.), Applications of CAE in Injection
Molding, Hanser, Munich, 1987.
10o7

71. K.K. Wang and V.W. Wang, in : A.I. Isayev (ed.), Injection and Compression
Molding Fundamentals, Marcel Dekker, New York, 1987.
72. M. Sobhanie and A.I. Isayev, in : A.I. Isayev (ed.), Modeling of Polymer
Processing, Recent Developments, Hanser, Munich, 1991.
73. B.S. Chen and W.H. Liu, Polym. Eng. Sci., 29 (1989) 1039; 34 (1994) 835.
74. S. Subbiah, D.L. Trafford and S.I. GiJ~eri, Int. J. Heat Mass Transfer, 32
(1989) 415.
75. P.A. Tanguy and R. Lacroix, Int. Polym. Process., 6 (1991) 19.
76. E.W. Liang, H.P. Wang and E.M. Perry, Adv. Polym. Tech., 12 (1993) 243.
77. B. Friedrichs and S.I. Gii~eri, J. Non-Newtonian Fluid Mech., 49 (1993) 141.
78. H.H. Chiang, K. Himasekhar, N. Santhanam and K.K. Wang, J. of Eng.
Mater. and Tech., 115 (1993) 37.
79. P. Kennedy, Flow Analysis of Injection Molds, Hanser, Munich, 1995.
80. T.J. Wang, L.J. Lee and W.B. Young, Int. Polym. Process., 10 (1995) 82.
81. V.R. Voller and S. Peng, Polym. Eng. Sci., 35 (1995) 1758.
82. E. Pichelin and T. Coupez, to be published in Comput. Meth. Appl. Mech.
Eng. (1998).
83. E. Broyer and C.W. Macosko, AIChE J., 22 (1976) 268.
84. J.M. Castro, Polym. Eng. Sci., 32 (1992) 715.
85. R.K. Mohammed, T.A. Osswald, T.J. Spiegelhoff and E.M. Sun, Intern.
Polym. Proc., 9 (1994) 279.
86. G.A.A.V. Haagh, G.W.M. Peters and H.E.H. Meijer, Polym. Eng. Sci., 36
(1996) 2579.
87. C.M. Hsiung and M. Cakmak, Polym. Eng. Sci., 31 (1991) 1372.
88. A.I. Isayev, T.W. Chang, K. Shimoyo and M. Gmerek, J. Appl. Polym. Sci.,
55 (1995) 807; 821.
89. G. Titomanlio, V. Speranza and V. Brucato, Intern. Polym. Process., l0
(1995) 1.
90. T.D. Papathanasiou, Chem. Eng. Sci., 50 (1995) 3433.
91. C.M. Hsiung, M. Cakmak and Y. Ulcer, Polymer, 37 (1996) 4555.
92. H. Ito, K. Minagawa, J. Takimoto, K. Tada and K. Koyama, Intern. Polym.
Proc., 11 (1996) 363.
93. O. Verhoyen, N. Van Rutten, R. Legras and F. Dupret, in preparation.
94. C.C. Lee, F. Folgar and C.L. Tucker, J. of Eng. for Industry, Trans. ASME,
106 (1984) 114.
95. T.A. Osswald and C.L. Tucker, Int. Polym. Process., 5 (1990) 79.
96. F. Trochu, R. Gauvin and D.-M. Gao, Adv. Polym. Tech., 12 (1993) 329.
97. S.G. Advani, M.V. Bmschke & R.S. Pamas, in" S.G. Advani (ed.), Flow and
Rheology in Polymer Composites Manufacturing, Elsevier, Amsterdam, 1994.
1008

98. B. Friedrichs and S.I. Gti~eri, Polym. Eng. Sci., 35 (1995) 1834.
99. R.B. Dessenberger and C.L. Tucker, Polymer Composites, 16 (1995) 495.
100. T.J. Wang, R.J. Lin and L.J. Lee, Intern. Polym. Proc., 10 (1995) 364.
101. C. L. Tucker, Polymer Composites, 17 (1996) 60.
102. G. Lebrun, R. Gauvin and K.N. Kendall, Composites Part A, 27 (1996) 347.
103. K. Himasekhar, L.S. Tumg, V.W. Wang, H.H. Chiang and K.K. Wang,
Adv. Polym. Tech., 12 (1993) 233.
104. G.W.M. Peters, P.J.L. van der Velden, H.E.H. Meijer and P. Schoone,
Intern. Polym. Proc., 9 (1994) 258.
105. C. Truesdell and W. Noll, The Non-Linear Field Theories of Mechanics,
Springer, Berlin, 1965.
106. R.B. Bird, R.C. Armstrong and O. Hassager, Dynamics of Polymeric
Liquids, Vol. 1 : Fluid mechanics (2nd ed.), Wiley, New York, 1987.
107. R. Keunings, in : C.L. Tucker m (ed.), Fundamentals of Computer Modeling
for Polymer Processing, Hanser, Munich, 1989.
108. A.I. Leonov, Rheol. Acta, 15 (1976) 85.
109. R.K. Upadhyay, A.I. Isayev and S.F. Shen, Rheol. Acta, 20 (1981) 443.
110. A.I. Leonov, J. Non-Newtonian Fluid Mech., 25 (1987) 1.
111. M. Simhambhatla and A.I. Leonov, Rheol. Acta, 34 (1995) 259.
112. A.I. Isayev and R.K. Upadhyay, in : A.I. Isayev (ed.), Injection and
Compression Molding Fundamentals, Marcel Dekker, New York, 1987.
113. K. Yasuda, R.C. Armstrong and R.E. Cohen, Rheol. Acta, 20 (1981) 163.
114. C.A. Hieber and H.H. Chiang, Rheol. Acta, 28 (1989) 321.
115. C.A. Hieber and K.K. Wang, Proc. Antec'90 (1990) 943.
116. L. Douven, Ph.D. Thesis, T.U. Eindhoven, 1991.
117. J. Greener, Polym. Eng. Sci., 26 (1986) 886.
118. K.K. Kabanemi and M.J. Crochet, Intern. Polym. Proc., 7 (1992) 60.
119. J.D. Ferry, Viscoelastic Properties of Polymers (3ra ed.), John Wiley &
Sons, New York, 1980.
120. M. Van Dyke, Applied Mathematics and Mechanics, Vol. 8, Perturbation
Methods in Fluid Mechanics, Academic Press, New York, 1964.
121. W. Rose, Nature, 191 (1961) 242.
122. M.R. Kamal, S.K. Goyal and E. Chu, AIChE J., 34 (1988) 94.
123. T. Sato and S. Richardson, Polym. Eng. Sci., 35 (1995) 805.
124. H. Schlichting, Boundary-Layer Theory, McGraw-Hill, 1968.
125. W.F. Zoetelief, Ph.D. Thesis, T.U. Eindhoven, 1995.
126. B.H. Kim and M.C. Ramesh, J. Eng. Industry, 117 (1995) 508.
127. B.H. Lee and B.H. Kim, Polym.-Plast. Technol. Eng., 35 (1996) 147.
128. S. Richardson, J. Fluid Mech., 56 (1972) 609; 102 (1981) 263.
1009

129. V.M. Entov, P.I. Etingof and D.YA. Kleinbock, Euro. J. Appl. Math., 6
(1995) 399.
130. G. Aronsson, Euro. Jnl. Appl. Math., 7 (1996) 417.
131. F. Dupret and J.M. Marchal, J. Non-Newtonian Fluid Mech., 20 (1986) 143.
132. L.W. Caspers, Ph.D. Thesis, T.U. Eindhoven, 1995.
133. H. Aoyagi, M. Uenoyama and S.I. Gtigeri, Intern. Polym. Proc., 7 (1992)
71.
134. C.W.M. Sitters, Ph.D. Thesis, T.U. Eindhoven, 1988.
135. A.N. Brooks and T.J.R. Hughes, Comput. Meth. Appl. Mech. Eng., 32
(1982) 199.
136. O.C. Zienkiewicz and R.L. Taylor, The Finite Element Method, Vol. 1 (4t~
ed.), McGraw-Hill, London, 1989.
137. W.F. Zoetelief, L.F.A. Douven and A.J. Ingen Housz, Polym. Eng. Sci., 36
(1996) 1886.
138. J.F. Stevenson, Polym. Eng. Sci., 26 (1986) 746.
139. M.R. Kantz, H.D. Neuman and F.H. Stigale, J. Appl. Polym. Sci., 16 (1972)
1249.
140. M.R. Kamal and F.H. Moy, J. Appl. Polym. Sci., 28 (1983) 1787.
141. J.P. Trotignon and J. Verdu, J. Appl. Polym. Sci., 34 (1987) 1.
142. C.M. Hsiung, M. Cakmak and J.L. White, Polym. Eng. Sci., 30 (1990) 967.
143. C.M. Hsiung, M. Cakmak and J.L. White, Int. Polym. Proc., 5 (1990) 109.
144. M. Fujiyama, Intern. Polym. Proc., 7 (1992) 84.
145. M. Saiu, V. Brucato, S. Piccarolo and G. Titomanlio, Intern. Polym. Proc., 7
(1992) 267.
146. C.M. Hsiung and M. Cakmak, J. Appl. Polym. Sci., 47 (1993) 125; 149.
147. P. Jerschow and H. Janeschitz-Kriegl, Rheol. Acta, 35 (1996) 127.
148. H. Janeschitz-Kriegl and G. Eder, J. Macromol. Sci.-Chem., A 27 (1990)
1733.
149. I.H. Hillier, J. Polym. Sci., Part A - Polym. Chem., 3 (1965) 3067.
150. N. Billon and J.M. Haudin, Ann. Chim. Fr., 15 (1990) 249.
151. T. Ozawa, Polymer, 12 (1971) 150.
152. M. Avrami, J. Chem. Phys., 7 (1939) 1103; 8 (1940) 212; 9 (1941) 177.
153. U.R. Evans, Trans. Faraday Soc., 41 (1945) 365.
154. C. Dufoss6, Ph.D. Thesis, ENSMP, Sofia-Antipolis, France (1990).
155. K. Nakamura, K. Katayama and T. Amano, J. Appl. Pol. Sci., 17 (1973)
1031.
156. S. Liedauer, G. Eder, H. Janeschitz-Kriegl, P. Jerschow, W. Geymayer and
E. Ingolic, Int. Polym. Proc., 8 (1993) 236.
157. W. Schneider, A. Ktippl and J. Berger, Intern. Polym. Proc., 2 (1988) 151.
1010

158. D. Andreucci, A. Fasano, M. Primiciero, M. Paolini and C. Verdi, Math.


Mod. & Meth. in Appl. Sci., 4 (1994) 135.
159. W.L. Sifleet, N. Dinos, J.R. Collier, Polym. Eng. Sci., 13 (1973) 10.
160. R. Lin, L.J. Lee and M. Liou, Intern. Polym. Process., 6 (1991) 356.
161. A.W. Chan and S.-T. Hwang, Polym. Eng. Sci., 31 (1991) 1149; 32 (1992)
310.
162. B. Liu, S. Bickerton and S.G. Advani, Composites Part A, 27A (1996) 135.
163. Y.-E. Yoo and W.I. Lee, Polym. Composites, 17 (1996) 368.
164. N. Patel and L.J. Lee, Polymer Composites, 16 (1995) 386.
165. J. Van der Westhuizen and J. Prieur du Plessis, Composites Part A, 27
(1996) 263.
166. S. Ranganathan, F.R. Phelan JR. and S.G. Advani, Polymer Composites, 17
(1996) 222.
167. F.R. Phelan JR. and G. Wise, Composites Part A, 27 (1996) 25.
168. A.C. Long, C.D. Rudd, M. Blagdon and P. Smith, Composites Part A, 27
(1996) 247.
169. W. Chang and N. Kikuchi, Comp. Fluid Dyn., 7 (1996) 49.
170. D. Ambrosi and L. Preziosi, Composites Part A, 29A (1998) 5.
171. C.-J. Wu and L.-W. Houmg, Polym. Eng. Sci., 35 (1995) 1272.
172. D.A. Steenkamer, D.J. Wilkins and V.M. Karbhari, Composites Manufact.,
6 (1995) 23.
173. D.R. Calhoun, S. Yalva~, D.G. Wetters, C.-H. Wu, T.J. Wang, J.S. Tsai and
L.J. Lee, Polymer Composites, 17 (1996) 251.
174. M. Kaviany, Principles of Heat Transfer in Porous Media, Springer, New
York, 1991.
175. R.A. Greenkom, Flow Phenomena in Porous Media, Marcel Dekker, New
York, 1983.
1011

F l o w o f P o l y m e r i c Melts in Channels with M o v i n g B o u n d a r i e s

A.I. Isayev a, C. Z o o k and Y . Z h a n g

alnstitute of Polymer Engineering, The University of Akron


Akron, OH 44325-0301, U.S.A.

1. I N T R O D U C T I O N

1.1 Significance of the Problem


The development of numerical techniques to accurately approximate the flow of
polymer melts in channels with moving boundaries is of paramount importance in
polymer processing. In polymer processing, such a flow situation occurs in
injection molding, extrusion and simultaneous injection/compression molding. In
the case of injection molding, such a flow situation occurs in the non-return valve
located on the front of the machine screw. During the injection stage of the
molding process, the valve is required to close to stop the flow of polymer melt
back into the screw region. Typically, a ring, ball or piston is utilized to close the
flow passage into the screw to facilitate this shut-off. In extrusion, varying the
geometry of the die by using choker bars or deformable lips allows the control of
melt flow to obtain products according to desired specifications. In simultaneous
injection/compression molding, polymer enters a mold and is compressed by a
moving boundary that is perpendicular to the flow direction. In all of the
examples, the boundary moves perpendicular to the dominant flow direction.
Current research in our laboratory has focused on the simulation of the non-
return valve during the molding process. Figure 1 shows a typical non-return
valve used in an injection molding machine. This valves has a cylindrical ring
which closes during the injection stage to close the passage into the screw. This
closure stops polymer melt from flowing back into the screw. During the
recovery stage, this ring opens to allow melt to accumulate in front of the
valve/screw assembly. This melt will be injected into the mold during the next
injection step.
1012

Ring
Valve o.~tflo~.~,( \ _Inflow

Outflow / Inflow

Figure 1. Sketch of typical ring type non-return valve for the injection molding of
thermoplastics.

A plot of the pressure traces in front of the valve on the barrel wall
(downstream) and behind the valve (upstream) in the screw section and screw
displacement during the injection and recovery stages is shown in Figure 2. At
the start of injection both pressures increase rapidly. The upstream pressure
reaches a steady value, while the downstream pressure in the screw metering
section decreases drastically. This decrease in pressure indicates the ring closing
the flow passage into the screw. Thus, the valve is closed. The closing time of
the valve can then be determined. For the recovery cycle, an oscillating pressure
is observed in the upstream pressure transducer. This is caused by the movement
of the flights of the rotating screw over the stationary pressure transducer. The
downstream pressure measurement has a steady pressure due to the accumulation
of polymer melt in the large reservoir (shot size) in front of the valve for the next
injection stage. The pressure drop across the valve/screw determines the amount
of resistance during the recovery stage, which affects the length of the recovery
time. The ability to simulate this process of the valve opening during the
recovery stage and closing during the injection stage will determine the time for
the valve to open, close and the forces acting on the ring, which will help in the
development of more efficient valves [1 ].
In order to understand the phenomena that takes place during the polymer
processing with moving boundaries, an experimental slit die with a moving
boundary has been designed (Figure 3). The polymer flows through the slit die as
a wall closes the flow passage. The transient pressure drop across this wall is
measured. To numerically simulate the transient channel flow with a moving
boundary, computational code based on the f'mite element method has been
developed using the Giesekus viscoelastic model. In addition, a commercial
computational package is utilized to understand the transient polymer flow using
a quasi-steady approach. The pressure drop from the die experiments and
numerical simulations will be compared in an attempt to validate the numerical
technique. The contribution of the elasticity of the polymer will be investigated
1013

30 14
Injection Recovery
25- Stage / -12~
/ -10~
~, 2 0 -
-a
15-
-- 6 ~
r~
~ 1==,4
10-
-4 ~

5 1--3" A ,. -2

~, i i i i I I I I I I

0 2 4 6 8 10 12 14 16 18 20
t, S
Figure 2. Pressure measurements across valve and screw displacement versus
time during the injection molding of HDPE with an injection speed of 2.54 cm/s
and a melt temperature of 204.4~

by comparing viscoelastic simulations with generalized Newtonian-inelastic


simulations using the Cross model. Changes in the velocity and stress
components will be discussed to indicate the effect the moving boundary has on
these variables. The usefulness of these methods in polymer processing will then
be discussed.
V,
I
U I,._~ Inlet
I
L......... - = q
zx]L.... ~ - - S J m ovt n g
L____v__~ !

XI~> ~Upstream DownstreamR Melt


i
mPressure Pressure j~ Probe
~ Transducer Transducer

Figure 3. Sketch of slit die with moving boundary.

1.2 B r i e f o v e r v i e w of viscoelastic simulations


In the past two decades, polymer melt flow has been extensively researched.
This research has led to an understanding of the basic flow of viscoelastic fluids
in many different polymer processing applications. The advancement of this
knowledge base has been expanded mainly through the implementation of
computational simulations, which can only approximate realistic polymer
1014

processes. In particular, the simulations is carried out for two or three


dimensional flow domains, complex geometry such as abrupt and tapered
contraction flows, and flow with transient or unsteady boundary conditions. Non-
isothermal boundary conditions have also had some limited development in
extrusion, but not in the area of abrupt or complex geometry.
Advanced numerical techniques available to simulate the complex flow of
viscoelastic fluids include the fufite difference and the fimte element techniques.
The books of Crochet [2] and Tucker [3] give a good review of these techniques.
To properly model viscoelastic flow, one needs to understand numerical
simulation techniques, the non-linear viscoelastic constitutive equations and the
mathematical characterization of these systems of equations. Techniques used in
past research include analytical solutions, decoupled fimte difference, decoupled
f'mite element, and coupled/mixed f'mite element methods. In a decoupled
approach, the goveming equations and comtitutive model are solved at different
steps of the solution process. In a mixed/coupled approach, the goveming
equations and constitutive equation are solved simultaneously. The choice of
governing equations (energy, momentum and continuity) and the assumptiom
which constrain the applicability of these equations, as well as the choice of
constitutive equation, determine the type of viscoelastic flow a numerical
technique can realistically simulate.
Together, the goveming equations and constitutive equation form a set of
nonlinear differential equations. This nonlinear set of differential equations is
descritized into an algebraic system of equations, which can be solved with
numerical techniques. The flow domain is divided up into elements or nodes and
solved locally for each set of algebraic equations. The solutions from these local
elements are then related globally to the entire flow domain to reach an accurate
approximation. One technique to solve this system of equations is the finite
difference method, although, much of the recent interest in numerically modeling
viscoelastic flow has been concerned with the finite element method. In this type
of approximation, the flow domain is discretized into a set of nodal points. The
governing equations are then approximated at each node point using a Taylor
series approximation. The constitutive equation is then solved separately using
an iterative approach. Past work in this area has been conducted by many
authors including Perera et al [4,5,6], Davies et al [7,8], Gatski and Lumley [9],
Townsend [10,11], Tiefenbruck and Leal [12], Yoo et al [13,14], Phelan et al
[ 15] and Yuan et al [ 16].
The method of choice for the simulation of viscoelastic flow is the finite element
method. In this method, the flow domain is descretized into triangular or
quadrilateral elements. The system of equations is approximated using the
1015

Galerkin method or method of virtual work. The equations are then solved by
standard elimination techniques such as Gauss or Gauss-Siedel elimination. The
solution of these equations gives approximations of the pressure, velocity
components, extra stress components, etc. The two techniques to solve these
equations are the decoupled and mixed/coupled methods.
In the decoupled method, the continuity and momentum equations are solved for
the Newtonian flow field. Then, the constitutive equation is solved separately
using some type of streamwise integration. Usually, a Picard iteration, which is
also called successive substitution, is applied to the constitutive equation to attain
convergence. The extra stress contribution from the constitutive equation is then
incorporated into the momentmn equations and then a new solution is found for
the momentum and continuity equation. This iterative procedure is repeated until
the convergence criteria is attained. Researchers to have used this method
include Upadhyay and Isayev [17,18], Isayev and Huang [19], Bush et al [20],
Lou and Tanner [21,22,23], Lou and Mitsoulis [24], and Hulsen and Van der
Zander [25,26]. Since the streamlines start as flow boundary, the recirculation
zones in contraction flows are solved by a numerical technique found in
Upadhyay and Isayev [17,18]. The use of decoupled techniques has also been
utilized for integral constitutive equations by Viriyayuthakom and Caswell [27],
and Dupont et al [28] using particle tracking and Lou [29] using the control
volume approach. Other methods include the Choleski decomposition method
with a Picard iteration by Mitsoulis et al [30,31 ], streamline upwinding (SU) and
Lesaint-Raviart methods by Fortin and Fortin [32], streamline-upwinding/Petrov-
Galerkin (SUPG) by Lou and Tanner [22], Barakos and Mitsoulis [33], and Lou
[34], the adaptive viscoelastic stress splitting/streamline integration (AVSS/SI)
scheme and adaptive viscoelastic stress splitting/streamline-upwind Petrov-
Galerkin (AVSS/SUPG) by Sun et al [35] and botmdary integral methods by
Bush and Phan-Thien [36,37].
The use of mixed methods to simulate viscoelastic flow has received a large
amount of attention due to the inability of these schemes to converge at high
Deborah numbers. Early work on these techniques employed the method
proposed by Kawahara and Takeuchi [38], where the Galerkin method and a
Newton-Raphson solver converged only at low Deborah numbers. Researhers to
have used this method include Crochet and Bezy [39], Crochet and Keunings
[40,41,42], Davies et al [43], Dupret et al [44], Debbaut and Crochet [45],
Keunings [46], Musarra and Keunings [47], Mendelson et al [48] and Kajiwara et
al [49]. Other earlier research of Yeh et al [50] applied numerical methods
designated as quadratic stress interpolation (QLL) and stress interpolation using
bilinear polynomials (QQL) with Newtons method to solve the system of
1016

equations. The QQL method was found to converge at higher Deborah numbers
and have a more stable convergence. This QQL method was also used by Brown
et al [51].
Later, Marchal and Crochet [52] developed a more advanced mixed method
where each element is divided into several bilinear sub-elements for stress to
improve the algorithms convergence ability. In addition, some authors used the
SU and SUPG methods introduced by Hughes and Brooks [53,54] to improve
convergence. Other authors to have used this method include Debbaut et al [55],
Pumode and Crochet [56], Hartt and Baird [57] and Yunm and Crochet [58].
Methods to improve the convergence of mixed numerical schemes have been
researched by Brown, Armstrong and coworkers [59,60,61,62]. King et al [59]
implemented a numerical method into the momenaun equations which makes the
elliptic property of this equation explicit called the Explicitly Elliptic Momentum
Equation (EEME). SUPG is also applied to this numerical scheme. This type of
method has also been utilized by Burdette et al [60] and Coates et al [61]. One
other formulation developed by Rajagopalan et al [62] is the elastic viscous split
stress (EVSS) where the viscoelastic stress is split into the viscous and elastic
stresses defined by x= _~ + ft. In addition, Sasmal [63] used this method.
Other methods applied to the mixed solution technique include the
spectral/fmite-element method by Beds et al [64], the time-
discontinuous/Galerkin least-squares (TD/GLS) method by Baaijens [65,66], the
Taylor-streamline upwind-Petrov-Galerkin/pressure correction formulation by
Townsend and Webster [67,68], and the modified EVSS by Geunette and Fortin
[69] and Azaiez [70]. Integral models have also been solved by mixed methods.
An example of this is the paper by Papanastasiou et al [71 ].
All of these numerical techniques incur some difficulties in the approximation of
viscoelastic flow. One problem studied by Holstein and Paddon [72] and
Lipscomb et al [73] is the singularities which occur at the re-entrant comer of
contraction flows. The extra stress terms can reach physically unrealistic values
since the flow behavior of the fluid at these comers is not well understood. The
gradients of the stresses at the boundaries can become excessively high. To
overcome this problem, either the mesh size is increased in the region of the re-
entrant comer or a slip condition is applied at the boundary. Another source of
instabilities is inherent in the constitutive equations themselves. Kwon and
Leonov [74,75] have investigated the Hadamard stability and the positive
def'miteness of the configuration tensor of different constitutive equations. Models
proposed by Larson, Leonov and Giesekus, which are known to accurately
predict viscoelastic flow, experience blow-up instabilities. These instabilities can
be overcome by properly modeling the dissipative terms, adding a small
1017

Newtonian viscosity or using a multi-mode approach with respect to the


relaxation spectrtma [17]. Van der Zanden and Hulsen [26] have also reported on
these types of instabilities in the simulation viscoelastic flow. Joseph [76,77]
investigated instabilities during the flow of viscoelastic fluids. He determined that
hyperbolicity changes the system of equations from an elliptic to a hyperbolic
type, thus making numerical solutions difficult.
The simulation of the moving boundary requires the use of a transient numerical
scheme. The basic understanding of the unsteady flow of constitutive equations
with a time dependent history of the fluid deformation is well documented. The
nonlinear response of viscoelastic fluid is different from the response typically
seen for a Newtonian fluid. To more accurtely simulate this fluid, a viscoelastic
model is applied in conjunction with a multi-mode approach to the deformation
history. One type of flow situation, the sudden imposition of flow, occurs in
many polymer process applications. The types of unsteady flow include a
suddenly imposed velocity gradient, constant pressure gradient, a periodic
pressure gradient, start-up flow with a wall moving parallel to flow, etc. In
unsteady flows, the elasticity does not effect the f'mal velocity, but at the
imposition of the velocity or pressure gradient a nonlinear response occurs in the
velocities and stresses of the fluid. For a suddenly imposed velocity gradient, an
overshoot in the pressure occurs. For the suddenly imposed pressure gradiem, a
velocity overshoot is observed and is followed by a velocity minimum before a
steady value is attained. Researchers to have investigated this phenomena
include Fielder and Thomas [78], Waters and King [79], Chong and Franks [80],
Chong and Vezzi [81], Townsend [82], Akay [83], Duffy [84,85], Balmer and
Fiorina [86], Ryan and Dutta [87], Upadhyay et al [88], Isayev [89] and Kolkka
and Ierley [90].
The solution of time dependent problems for the full set of governing equations
has held much interest in the modeling of viscoelastic fluids. Along with the
inclusion of a time variable, numerical stability and convergence difficulties are
already confronted when trying to model the steady flow of a viscoelastic fluid.
In addition, time dependent flows add an increased amount of computational time
to solve problems. The numerical scheme typically used to approximate the time
variable is a fmite difference approximation. Past publications applying transient
simulations include the research of Townsend [10] using a De-Fort Frankel
scheme, Keunings [91], Northey et al [92] using fully implicit integrator and
semi-implicit integration, Keiller [93], Olsson [94] using a predictor-corrector
method for time stepping, Rasmussen and Hassager [95] using a Lagrangian
integral method and Baloch et al [68]. Results indicated that stable time
dependent solutions were attainable, but flow occurring at high Deborah numbers
1018

were not presented.


In order to understand the behavior of a viscoelasstic fluid, the experimental
techniques that have been employed include streak photography, stress
birefringence studies, multiple flash technique and laser doppled velocimetry. An
excellent review of this research has been published by White et al [96] and
lsayev and Upadhyay [97]. These methods work well when clear polymers are
used in the experimental research. Unfortunately, fluids such as rubber
compounds which cannot be seen through or polymers which can only be
processed with industrial processing machinery cannot take advantage of these
experimental methods. The only other experimental variable to measure is the
pressure at various locations in the flow channel. Experimental work on pressure
prediction to numerical simulations has been conducted by Isayev and Upadhyay
[97] and Huang et al [98,19]. Results though indicate that the prediction of
pressure is not as accurate as the prediction of the other field variables. In
addition, this error can become significantly high for higher flow rates and in an
abrupt contraction or expansion.
In the present research, a decoupled finite element method previously used in
two dimensional unsteady viscoelastic modeling is applied to the solution of
polymer flow in channels with a moving boundary. The momentum and
continuity equations are solved using the method of virtual work. The
constitutive equation is solved by a Picard iteration technique. This formulation
has been shown stability up to a Deborah number of 50 and a Weisenberg
number of 4.2 by Upadhyay and Isayev [17], a Deborah number of 270 by Isayev
and Upadhyay [98], and recently to a Deborah number of 845 by Isayev and
Huang [19]. The only limitation in achieving a higher value of the Deborah
number is the density of the mesh. A higher density mesh, hence, allows
convergence at higher Deborah numbers. Added to this past f'mite element code
is the simulation of transient flow with a moving boundary, which is closing the
flow domain. This boundary moves in the direction perpendicular to the main
direction of flow. In addition, the computational package FLUENT has been
utilized to simulate the flow of a generalized Newtonian fluid based on the Cross
model. The simulation using this package applies a quasi-steady approach with a
velocity imposed at the moving boundary where the wall is closing the channel.
Through these two paths of numerical simulations, the understanding of the effect
the transient nature of flow with a moving boundary has on viscoelastic flow will
be developed and understood.
1019

2. THEORETICAL

The modeling of polymer processing can be grouped into three particular areas
of research. One is based on the generalized Newtonian theory, which takes into
account only the nonlinear shear rate dependent viscosity of the material. The
other is based on the viscoelastic theory of polymer melts, which has many
different forms, but is mainly concerned with incorporating the non-linear stress
terms into some type of constitutive equation along with the nonlinear viscosity.
To complete the theory of viscoelastic modeling, the viscoelastic-plastic fluid
theory is also covered.

2.1 Generalized Newtonian Fluid


For most types of fluids, the viscosity of the material does not change as the
shear rate on the material increases. For an incompressible Newtonian fluid the
shear stress is related to the strain rate by the equation

r~ = -ja~ (1)

where ~t is the Newtonian viscosity which is a constant for a given temperature,


pressure and composition. Early rheologists used the empiricism that the
viscosity could be a function of shear rate to model the shear rate dependence of
the viscosity of suspensions, pastes, polymer melts and solutions. Thus, they
incorporated the dependence into the previous equation as

(2)

where r/(~;) is the apparent viscosity, which is a function of shear rate 2, pressure
and temperature.
Many generalized Newtonian equations have been postulated since the
realization of this type of relation. The types of available models are the Cross-
Arrhenius, Power Law, Klein, Carreau-Yasuda, Spriggs tnmcated power law,
Eyring, Powell-Eyring, Sutterby, Ellis, and Bingham. Bird gives a good review of
many of these models in chapter 4 of reference [99].
Two very useful generalized Newtonian relations, are the power law and the
Cross-Arrhenius equations. The power law equation of Oswald and de Wade
[ 100,101 ] is written as

-n--1
r/=mg (3)
1020

where m is a constant and n is the power-law index. This equation can only fit the
high shear rate range of the viscosity curve called the power law region. This
equation is useful since many analytical solutions are available, which can be
easily understood. Many of these solutions can be found in the book of Bird
[99]. In addition, this equation is capable of being implemented into numerical
solutions easily. An example of this is a paper by Hieber and Shen [ 102] and also
in Hieber [103] in which a finite element/finite difference is used to simulate the
injection-molding filling process. Hieber [104] and Isayev and Upadyhay [97]
also used the power law equation along with the finite element method to
estimate the pressure drop and extra entrance length across planar and
axisymmetric contractions. The main drawback of this equation is the inability to
accurately model the Newtonian or low shear rate range of viscosity.
The model by Cross [ 105, 106] is a three parameter equation of the form

7/
n(Y) = o (4)
i-n

where rio is the zero shear viscosity and ~* is the shear-stress level at which rl is
in transition between the Newtonian limit rio and the power law region. The
effect of temperature and pressure on the melt viscosity can be taken into account
through an Arrhenius type dependence. Hieber gives a thorough description of
these types of equations in chapter 1 of reference [ 103].
The modeling of the injection molding process by the Cross-Arrhenius equation
has been well documented since the early 1980's. Hieber et al [107] used the
Cross-Arrhenius type equation to model material viscosity to understand the
effect of juncture losses resulting from the elasticity of the fluid. By comparing
the numerical viscous simulation and experimental data, they were able to deduce
the amount of juncture pressure from the elastic effects. Sobhanie, Deng and
Isayev [ 108,109,110,111 ] have extensively used the Cross-Arrhenius equation for
the injection molding of robber compotmds. An additional parameter was added
to this simulation to account for the cure kinetics of the rubber compounds during
injection molding. The numerical pressure predictions showed qualitative
agreement with the experimental pressure data.
Other researchers who have used generalized Newtonian models include Duda
and Vrentras [112]. These authors used the Powell-Eyring equation to
understand the effect nonlinear viscosity has on the amount of extra pressure loss
1021

in juncture regions. Kim-E, Brown and Armstrong [113] used a Carreau type
viscosity equation to simulate the flow through an axisymmetric contraction.
Results indicate steep velocity gradients occurring near the wall caused high
shear rates and a shear thinning viscosity. These steep gradients effect the
accuracy of the finite element approximations.

2.2 Viscoelastic Fluids


The limiting factor in the ability of generalized Newtonian theory to predict
polymer melt flow is the inadequacy of the equations to predict the elastic effects
in a flowing polymer. These effects are extremely important in flows of polymers
through sudden contractions and expansions. In addition, nonlinear stresses
arising from polymer melts contribute to large pressure drops and steep stress
gradients. Also, the elastic effects can dominate in start-up or transient flows of
polymers. Thus, these elastic effects have a significant impact on the flow of
polymer melts in many areas of processing. The modeling of viscoelastic
phenomena employs two main types of constitutive equations to model polymer
flow. The two types are the integral and differential type equations [99]. In the
present paper, we will consider the differential models only.
To characterize the flow of viscoelastic fluids, the dimensionless Deborah
number defines the rate of straining on a fluid as

De = .3U ~, (5)

where 3U/b is the characteristic shear rate with U being an average velocity
downstream of the contraction and ~, is the mean relaxation time defined as

~rlkkk
k = x_-, (6)
N
with TIk being defined as 770= 77~+~ = k, TI~ is a parameter similar to the lower
Newtonian viscosity, and TIk, and Ek are the viscosity and the relaxation time in
the k th mode.
Most differential models are only an extension of the Maxwell [114,99,115]
type equations from the 1860's. The Maxwell model assumes a linear relation for
the viscous and elastic responses of the fluid. This relation can be written as
1022

Xxy+ G /)t - -tt~xy (7)

where 17xy is the shear stress and G is the shear modulus. For steady-state
motions, the equation simplifies to the Newtonian fluid. This equation shows
time-dependent response upon an imposition of flow. In addition, other types of
constitutve equations can be explained depending on the type of time derivative
utilized for equation (7). For sudden changes with time, the equation simplifies to
a Hookean solid. One can introduce the time derivative 5/8t in the convected
coordinate system such as

V T
~-,__=a=$-Vv.a-~=.Vv (8)

where ~ is the substantial time derivative, which translates with the material
particle, Vv is the velocity gradient with Vv~ denoting the transpose. This
system is called the upper convected coordinate system where the base vectors
are parallel to material lines. Therefore, the vectors are stretched and rotated
with the material lines, t, the time derivative, is written as

D a
~ = b--Tx== ~-x=+ v__.Vx__ (9)

Limitations for this model exist which exclude it from properly predicting realistic
polymer flow at high Deborah numbers. For instance, the model does not include
any type of shear rate dependent viscosity. The first normal stress coefficient is
not shear rate dependent either. The elongational viscosity becomes infinite at
f'mite strains and moreover, the recoverable strain is over-predicted at high strain
rates. In addition, the model cannot predict polymer behavior after the imposition
displacements strains.
In later developments, Oldroyd [116] proposed a quasi-linear differential model
which was frame invariant called the Oldroyd-B model or the convected Jeffrey's
model [117]. This model included a time dependent deformation tensor and is
written as

~.,~-~+~ = 2no D+ X.,~D (lo)


1023

For this equation, one needs to provide three parameters where rio is the zero
shear rate viscosity, D is the deformation tensor, s is the relaxation time, and
is the retardation time. This equation is capable of describing time-dependent
flows, but is still unable to correctly predict rheological behavior. Equations of
this type are limited to polymer flows with small deformation rates or low
Deborah numbers. To account for these deficiencies more elaborate differential
models have been developed such as the Maxwell corotational, White-Metzner,
Gordon-Schowalter, Johnson-Segalman and Oldroyd-8 constant models which
appear in Bird [99] and Larson [115]. These models were capable of describing
more of the rheological properties of melts, but were still unable to describe all of
the material properties correctly.
In the mid-1960's, models were proposed that could describe polymer flows
more accurately. All of these equations contain nonlinear stress terms. With the
incorporation of this term, the constitutive equations are more suitable in
predicting stresses in shear flow. One model which includes the quadratic stress
term is the nonlinear viscoelastic equations developed by Giesekus [118-122].
This constitutive equation is based on the concept of a deformation-dependent
tensorial mobility or drag. The equation is derived from the theory for
concentrated solutions and melts using the dumbbell theory for dilute solutions.
The assumption is made that the mobility is not dependent upon the individual
configuration of each polymer segment. Instead, an average configuration of all
the segments is used to relate the mobility tensor to the configuration tensors.
This averaging of the polymer segment configuration bridges the gap between the
molecular ideas from which the constitutive equation comes from and the treating
of the polymer chains as a continuum of polymer melt or solution. This equation
replaces the scalar mobility constants Bk with a non-isotropic mobility symmetric
second order tensor 13to give an equation of the form

~.a+~-~-~= = 2rid (11)

where $ is the elastic stress term, I"1 is the shear rate dependent viscosity, G is the
shear modulus and D is the rate-of-deformation tensor. To achieve realistic
predictions, two assumptions are made on the dependencies of $ and l~ on the
elastic strain tensor C. First, the relation [G-5] of _ to ~ is of the neo-Hookean
type dependence and is written as
1024

where ~ is the unit tensor. The shear modulus, G, is related to nkT, where n is
the number of beads per unit volume, T is the temperature, and k, is Boltzmann's
constant. The second assumption [120] is that a linear relation exists between 13
to C and is written as

= ~ + oc(C- ~)= (1- ct)8 + txC (13)

In this equation tx is an empirical constant of proportionality (mobility factor)


and is related to the compressibility of the material, oc must satisfy the condition
of 0 < cz < 1. The most general form of the constitutive equation can be written
as

a 2 3
+ ~-~ + ~,~-~ = 2r/D (14)

When oc equals O, equation (14) reduces to an isotropic mobility tensor and the
UC Maxwell model is retained. When oc equals 1, the anisotropy is at its
maximum and equation (14) produces results in shear and extension similar to
those for the corotational Maxwell model. For the research presented in this
paper, the assumption of incompressibility of the polymer is assumed and oc is
equal to 1A. In addition, a multi-model approach is applied to equation (14)
resulting in

1 2 3
l: + ~ ~ k + 2,k-~_~ = 2 r/kD (15)

where k denotes the k th mode.


The basic understanding of this model has been well documemed for many types
of simple flows. Giesekus [120,121] published results on the predictuon for
simple shear and simple extension which predict shear thinning and a non-
vanishing first and second normal stress difference. In addition, this research has
defined where the Giesekus model predicts real solutions by Yoo and Choi [123]
and viscoelastic instabilities in shear by Oztekin et al [124]. Other f'mdings
include the Giesekus and other models ability to predict Poiseuille flows by
Schleiniger and Weinacht [ 125], sinusoidally undulating channel flow and Taylor-
Couette flow instabilities by Beds [126], shear flow and experimental
comparisons by Vlassopoulos and Hatzikiriakos [127], exponential shear stress
coefficient and elongational flow by Schieber and Weist [128], shear flow, and
1025

uniaxial and biaxial extension by Khan and Larson [129], steady and transient
shear flow by Quinzani et al [130], and tmiaxial, biaxial and elongational flow by
Isake et al [131]. Larson [132] has compared several different models including
the Giesekus model for steady-state flows, start-up steady straining, stress
relaxation following cessation of steady straining, single and double step strains,
and elastic recovery, in sheafing, or uiaxial, biaxial and planar extension. Other
work has discussed the compatibility of equations with equilibrium
thermodynamics by Grmela and Carreau [133] and the restriction of the extra
stress tensor and requirements that this tensor must be positive def'mite by Hulsen
[134,135].
Other differential equations which are often used in the modeling of viscoelastic
fluids are the Leonov [136,137] equations, the Phan-Thien Tanner (PTT) model
[138,139],the f'mite extensible nonlinear elastic (FENE) model [140,141] and the
Doi-Edwards [142] model. For practical purposes, constitutive equations
quadratic in stress are the most popular for today researchers. More complex
models can be derived, but the implementation of these models would probably
be to cumbersome due to the complexity of the mathematical equations. In
addition, to many parameters would need to be specified in order to implement
the models. Excellent reviews of constitutive equations and the relations between
the different models is available in the books by Bird et al [99] and Larson [115].

2.3 Viscoelastic-plastic fluids


In the modeling of polymeric fluids, one other area of interest is the modeling of
filled polymer systems. The inclusion of the interaction between polymer and
filler adds additional assumptions to the constitutive model. These filled systems
are characterized by the rheological behavior of the matrix, the particle
characteristics, the dispersed state, the interaction between particles, and the
interaction between the filler and the polymer matrix. In addition, the particle
size effects the rheological properties of the filled polymer melt. For large
particles, an increase in shear and elongational viscosity if observed. In systems
with small particles, yield values in shear and elongational flow occur, as well as,
other changes in the rheological properties. The approach to modeling these
filled systems is either a continuum approach or a micro-mechanics approach.
The first to propose a stress-deformation rate equation for a fluid with yield
stress was Bingham [143]. This was subsequently followed by the equations of
Prager and Hohenemser [144]. These authors proposed a constitutive model for
the flow of an incompressible viscoplastic material as

D = 0 for J2 -<Y (16)


1026

J2-Y
- - - - - - x = 2riD for J2 > Y (17)
J2 = -"

for rate-of-deformation tensor, where Y is the yield function, and gj is the


deviatoric stress tensor def'med as

x = ~o..-/3o=
m J
8..
~J
(18)

The second invariant of the deviatoric stress tensor is def'med by

J2 = 4~-1;. q; (19)

Subsequently, Slibar and Parsley [145,146] proposed equations based on


thixotropic phenomena with time-dependent yield values. The rate of
deformation tensor and deformation were both incorporated into the yield
function Y. Later, a viscosity function was proposed by Harris [147] which
assumed an initial viscosity could be recovered after a long enough time.
The first to develop a viscoplastic rheological model with yield was Schwedoff
[ 148]. Using a modified form of the Maxwell model, this equation is written as

=0 when x>__Xy (20)


~,d x + (x - Xy)= rl~/ when x>_Xy (21)
dt

In steady state shear flow, this model reduces to

x = Xy+ rl~ when x __.Xy (22)

Later, HuRon [ 149] proposed a model which uses a yield criterion with a memory
function. White [150,151] also developed a more specific viscoplastic model
based on the combination of avon Mises yield criterion and a hereditary function.
To more accurately model the thixotropy of fluids, Suetsugu and White [152],
Montes and White [153] implemented a time dependent yield criterion. Isayev
and Fan [ 154] used a von Mises yield criterion and the Leonov viscoelastic model
to propose an equation for f'flled polymers. This model though does not predict
thixotropy. Often, filled systems will have time dependent properties caused by
the breadkown and build-up of particle structure. To account for this time
dependence, Cheng and Evans [155] have developed an isotropic,
1027

incompressible, inelastic, thixotropic model where the particle breakdown and


build-up process is included. Others to have developed this type of theory
include Cheng [ 156] and Kemblowski and Perera [ 157].
More advanced models in filled systems include the work of Leonov [158] and
Simhambhatla [159]. These authors developed a rheological model based on the
Leonov viscoelastic model [136]. This model assumes that the total stress is the
sum of viscoelastic stresses developed from micro-flow of the polymer matrix
around flocs, and the stresses due to particle-particle interactions of the dispersed
particle phase. In addition, Coussot et al [160] developed a similar model for low
molecular weight matrices. Sobhanie et al [ 161] developed a viscoelastic-plastic
model where the total stress is assumed to be the sum of the mean stresses due to
micro-flow of polymer melt arotmd a particle and the mean stress due to particle-
particle interactions. The Leonov model and a yield criterion are used to
approximate the mean stresses in the polymer matrix and filler network. Yield is
def'med by a function in terms of a differential equation with an internal
parameter describing the evolution of scructure changes during floc rupture and
restoration. In this model, the total stress ~ is represented by the equation

m p
x=x +x (23)

m
where _x_ is the mean stress in the polymer matrix and ~P is the mean stress in the
filler network. The mean stress in the polymer matrix is represented by the
equation

N
"rm= P~ + 2 r/~D + 2~.~(Wk.lCk - Wk,2Ck
-1) (24)
k=l

where ~ is the elastic strain tensor and Wk is the elastic potential. The mean
stress in the filler network is represented by a scalar yield function

p Y
x =--x (25)
= J2 =

The particle network is modeled by a Voight-Kelvin viscoelastic equation which


assumes that there is an initial modulus go which can decay and recover based on
the amount of damage to the filler network.
1028

3. N U M E R I C A L SIMULATIONS

3.1 Finite difference formulation for a generalized Newtonian fluid


The first numerical approach used to simulating the flow of a polymer fluid in
this paper is the finite difference technique. This numerical approach along with
the finite element technique in the section 3.2 are applied since the equations to
model two dimensional polymer flow are too complex for the derivation of
analytical solutions. For viscous-inelastic simulations, the commercially available
computational fluid dynamics code FLUENT is utilized. To more accurately
model the nonlinear viscosity of the polymer melt, FLUENT was modified to
incorporate a Cross model viscosity function. A quasi-steady analysis is applied
to the viscous solutions to account for the time dependent nature of the moving
boundary. The solution of this problem using finite difference code will provide
an understanding of the dynamic response of non-return valves.
The finite difference code FLUENT uses a pressure based segregated finite-
volume method to solve the governing equations. Further aspects of this
numerical formulation are found in the FLUENT manual [162] and the book of
Patankar [163]. The incompressible form of the continuity equation is written as

V.v=O (26)

where v is the velocity vector and the isothermal, incompressible and steady form
of the conservation of momentum equations with no body forces is written as

V.~=O (27)

where ~ is the total stress acting on the fluid. The model implemented is the
Cross model defined by equation (4). To determine the shear rate dependent
viscosity of this model the local strain rate is determined by

=
3ui ~3u~ (28)
r'~ ~ + Ox,

where i and j represent the direction of the velocity or displacement. The scalar
invariant of this local strain rate is determined as the local shear rate.
The two dimensional continuity (26) and momentum equations (27) are solved
by integration. The differential form of the momentum equations are written as
1029

(OVx v OV ]
P v'-~"x + Y Oy )=--~-x + ~9-T- (29)

v Ov,]
P vx-~x + '/9y ) = /gx + i)---~ (30)

for the x and y directions respectively. The differential form of the continuity
equation is written as

~9v_._Lx ~)Vy
Ox +--~y = 0 (31)

where Vx is the velocity component in the x-direction or axial direction, and Vy is


the velocity component in the y-direction or gap wise direction. (Yijis the stress
components in the indicated directions. The stress tensor ~ for the generalized
Newtonian flow is written as

g=-p +n(V.v+V.v") (32)


and the individual stress components are written as

~Vx
Oxx = - p + rl Ox (33)
~gvy
o . = -p + rl ~)y (34)

( Vx
~x, = -P + rlt--~-y + Ox ) (35)

where p is defined as the pressure.


To approximate the differential equations the flow domain is divided using a
finite-volume method as shown in Figure 4. The node points are located along
the grid lines. The scalar and vector variables are stored at the center of each
control volume. The governing differential equations are solved by integration
across each control volume using the divergence theorem. In the solution of these
field variables, the pressure at each control volume face is determined. In
addition, variables such as the velocity components, which obey a momentum and
mass balance across each face, must be determined. Across each face, there is an
1030

assumed average mass balance. The mass balance is applied to ensure that
continuity exists between the variables in a given control volume and the
variables of neighboring control volumes.

node,. : node
(i,j+l) (i+l,j+l)

Cell Center
face Q
location of
variable storage

node node
(i,j) 9 (i+l,j)

Figure 4. Control volume scheme for use in FLUENT.

In the solution of equations (29), (30) and (31) for each control volume,
FLUENT uses several different interpolation schemes to give a more accurate
approximation of the field variables across each control volume. The choice of
interpolation schemes includes the power law scheme [163], blended Second
Order Upwind/Central Difference [164] and QUICK [165]. The power law
scheme uses a one-dimensional convection-diffusion equations for interpolation.
The blended second order upwind/central difference and QUICK higher order
interpolation schemes, which determine the faces values between control volumes
based on the values stored at the centers of two control volumes and an adjacent
control volume upstream of the first two cell centers. The higher order
interpolations schemes result is more accurate approximations in the solution of
the flow domain.

3.1.1 Numerical Scheme for Finite Difference Method


For the solution of the flow domain, FLUENT uses an iterative procedure First
the velocity components are solved for using an initial guess for the pressure.
Mass balance is then applied to correct the pressure. The viscosity from the
Cross model is then updated. The governing equations for each control volume is
solved based on the values of neighboring control volumes. The procedure
continues using a line-by-line solution of the equations until a satisfactory error in
the residuals has been obtained.
A quasi-steady approach is applied to solve for the transient moving boundary.
This type of approach is applied to very viscous fluids from knowing that the
Reynolds number is very low and applying the assumption that creeping flow is
1031

occurl'ing. In addition, the assumption is made that the flow is isothermal. In the
simulations the position of the wall from fully open to eighty-five percent closed
is divided into a given number of displacements. A control volume grid is built
for each displacement using the FLUENT preprocessor. The boundary
conditions are applied to each grid including the flow rate, material properties,
and constant closing velocity of the wall. To approximate the velocity of the
moving wall, the sides of the wall use a wall boundary velocity, which is parallel
to the wall. The velocity of the bottom face uses a flow inlet condition with a
uniform velocity across the entire face. The velocity boundary conditions are
shown in Figure 5.

v, =vy =0.0 vy= wall v~ =vy--0.0


velocity i

v~= walll<:::]v,=O.O ~7 v, =0.o[:~L[Ivy--velocity


WGII
i
Inle~y velocitt
[[] Inlet condition'Exit
J" ,with uniform
vx =O. Op f x ' , , ~ _ . j ~/~vel~

v~ =vy =0.0
Figure 5. Boundary conditions for numerical simulations of moving wall and die.

3.2 Finite element formulation for a viscoelastic fluid


To fully understand the effect of the moving boundary, the viscoelastic
rheological properties of a polymer fluid must be incorporated into the simulation.
Thus, numerical methods such as the f'mite element method are used to properly
describe the polymer melt flow. In order to simulate the closing of the slit die
with the FEM, the continuity and momentum equations must be put into f'mite
element formulations. The Giesekus constitutive equation will be solved by the
streamline integration technique of Viriyayuthakom and Caswell [27] and Shen
[166]. Finite element code using the FORTRAN programming language already
exists for the Leonov model. This code was originally developed by Upadhyay
and Isayev [97,18,17] and Huang and Isayev [19,98]. This code simulates the
viscoelastic flow through tapered and abrupt planar contractions. Based on this
code, quasi-three dimensional transient moving boundary f'mite element code for
planar complex geometry has been developed by Zhang. As mentioned
previously, the solution of this problem using the f'mite element technique will
help to understand the dynamic response of a non-retum valve.
For the f'mite element formulations, incompressible form of the continuity
equation is written as
1032

V.v =0 (36)

and the isothermal and incompressible form of the conservation of momentum


equation with no body forces is written as

Dv
V.~= p Dt (37)

where v is the velocity vector, p is the fluid density, and r is the total stress
acting on the fluid. The operator D( )/Dt is the material time
derivative b( )/bt + v. V(). The constitutive equation to be solved is the multi-
mode Giesekus model with the incompressiblity condition a equal to 89
corresponding to the incompressibility condition. This equation is written as

1 2 0
"t" + - ~ k ~ + ~,k~-_Tk = 2r/kD (381

where ~.k is the relaxation time, Gk is the shear modulus and rlk is the viscosity of
the k th mode. The extra stress tensor is incorporated into the Cauchy stress tensor
and is written as

r +~ (39)

where p is the pressure, 5__is the unit tensor and r__L_is the extra-stress tensor. For
the Giesekus model, the Cauchy stress tensor will be written as

Npara

~--- "" -- = n~k

where w is the extra-stress tensor of the k th mode with Npara denoting the total
-'k

number of nodes and p is the pressure contribution to the stress.


The continuity and conservation of momentum equations are solved using the
virtual work f'mite element method with six-node triangular elements. The two
dimensional momentum equations in the x and y directions are written as

(~v,~ aw ~yX1 aC~x ~r (41)


1033

(Ovy Ov, Ovr~ o,=~ o,=. (42)

respectively, and the continuity equation is written as

0vx 0vr (43)


0x +-~-- = 0

where p is the fluid density, Vx is the velocity component in the x-direction or


axial direction, and Vy is the velocity component in the y-direction or gapwise
direction, oij is the stress components in the indicated directions. For the Cauchy
stress tensor, the individual stress components are written as

Ov, (44)
k

Ovy
o-,,,, = - p + r/,--~-- + ~ r,,,,,,, (45)

O-y-- 7r/,C(Bv~
T +~) (46)
k

The extra stress terms x~j,k are the contribution of stress from the constitutive
equation for different modes of relaxation.

3,v3,P3)

(Ul,Vl,P~.."........~ (u2'v2'P2)

Figure 6. Six-node triangular element for u-v-p formulation with triangular area
coordinate system.

To find an approximate solution, the flow domain is discretized into a series of


sub domains or elements as shown in Figure 6. These subelements are applied to
the flow domain for the solution of the continuity and conservation of momentum
1034

equations with triangular elements. The velocity components are approximated


by second order polynomials with the node locations at the comers and
midpoints. The pressure is approximated by f'~st order polynomials with the
location of the nodes at the comers of the triangular elements. The velocity
components and pressures are approximated by shape functions N and M,
respectively. The shape functions can be written simply as, v~ = N__~u,v~ = N_yv,and
p=M_.p. Further development shows that these shape functions are a set of
coordinates, which determine the location of each triangular element and the
subsequent node points in the flow domain. The relation def'lnes these
coordinates by

L = Ai / A = (ai + bi x + Ci y) / 2A (47)

where A is the total area of the element, Ai is the are the area coordinates as
indicated in Figure 6 and ai, bi and ci are the coordinate values of the three comer
nodes. The three area coordinates satisfy the following relation

L, + L2 + L3 = 1 (48)

The area coordinates have the unique relation that L,= 1 when x = xj, and y = yj if
i= j. Furthermore, the area coordinates have the relation that L,= 0 when x = xj,
and y = yj if i ~ j. One other constraint is that 0 < Li < 1 w h e n x ~ x j and y ~ yj. The
velocity components use the same quadratic shape functions, which are def'med
by six different variables for each element such as

N=[N,,N2,N3,N4,N5,N6] (49)

with each Ni being def'med by the standard triangular quadratic basis relations of

N~= L~(2L~-I), N2= L2(2L2-1), N3 = / _ o ( 2 L 3 - 1 ) ,


N4=4LIL2, Ns=4L2L3 ~ Nn=4L3LI (50)

The pressures use the linear shape functions defined by three different variables
for each element written as

M_M_=[M~,M2,M3] (51)

with each M being defined by the standard triangular linear basis relations of
1035

Ml = L1, M2=/_,2 & M3 = L3 (52)

The derivaties of the velocity components and pressure shape functions are
defined as

(9Vx ON 0Vx 0N
- - (53.a)
~gx - igx u, iOy ~gy u
0Vy 0N 0Vy 0N
~ ~ ~ ~ V
0x 0x v, by by - (53.b)
Op OM Op a M
&- &- - p__, Oy- ~- - p__ (53.c)

Equations (41), (42) and (43) are written into the finite element formulation using
the method of virtual work in two dimensions as follows

~=6v= ~ + = i= v9 otw+v w+v ' dR (54)

.l',,<~v, ~ +-~--)d=: .1",,<~Vy. p t . ~ + v.-~- + v, (55)

and J=~. 0x + =0 (56)

respectively, where dR is def'med as the integration in a given region. Applying


the Gaussian theorem written in the form of

~v~~) -gd (~v~.())---g-()~(avD (57)

The left-hand side of the momentum equation (54) can now be rewritten as

which can be written in the form


1036

f~(~v"r')a~-I,[ ~(~v')~x~ ~(~v')


aR~y~" ] (59)

and equation (55) is

(~v,o.)+# (60)

which can be written in the form

I~(~v,.~)~-I~
[~v
o( ,)~. +.v,/}.
~y ~. (61)

where F is the integration along the boundary of the domain and T1 and T2 are a
simplified way to write the relations in the brackets. Substituting these relations
into the conservation of momentum equations (54) and (55) results in the
following set of equations

I, [~v
o( ,)oo+ ~(~v~,
Oy ~ ]dR+I,~v~.o~-~+v~-~+v,
~v~ ~vx ~)~ =
(62)

and

~(~ v,. T2)dF (63)

Introducing the shape functions for the velocity components from equations
(53.a) and (53.b), the conservation of momentum equations (54) and (55) can
now be written in the following form.

+ ay Ox, dR+j'u_ u _ _ . t , - ~ + u+ v =

(64)
1037

and

L- ~ 0y r /
dR +~R__
Nr N ' p /
"~ + u + -~z v
_dR =

8Vr~r(Nr" T2)dI" (65)

In addition, the continuity equation (56) can be written as

r_ R r(ON o (66)

by introducing the pressure shape function from equation (53.c). The stress
components in the momentum equations are represented by the following
equations with an artificial damping parameter co to control the stability of the
equations. These equations are written as

axx = - p " + 2 +co ~ - 2 ( o ' & + ~ ' r m-x' x,k (67)


k

ayy=- +2 +oJ)--~---2o90y +Z'r~yJ:k (68)

a, = + co + - o~ 9 v7 -~ a Vy -~ m-~ (69)
ay + oe + T.:

where m is the variable to be solved and m-1 is the variable from the previous
solution. This set of equations can be simplified and written in a matrix
representation to form a set of equations, which will be solved for each element
to give the approximate solution of the flow domain. This matrix is written as

i ONr ON-~R r
m

r/'+m ax +

f N r ONdR
1038

fOv,, r O3Vx T m
- - I NrNdR + J-7--N NdR ---~-N_ N_aR 0
Rm - R o'x----

_ -

o 0

T T
og~ N o
2~~+~ -
R-L~ m-1

r ~N~
R_L ~, ~+2--~---~~ o
o o o]

1 r
~N NdR 0 o
R u
+p o
0
~ _NTNdR o
0 0
fi t"

_r ,~
~N ] + N r T2 (70)

L 0
0

In these matrix equations, the N r and N vectors indicate the shape functions for
the velocity components with six constants and the M indicates the shape
functions for the pressures with three constants. The terms which include the At
terms indicate the f'mite difference approximation of the time stepping. The
velocity components are the initial conditions for the time step. The fmite
difference representations of the velocity components are given as
1039

m+l m m+l m
Ou u -u Ov v -v
- - - - and - - - - (71)
Ot At & At

in the x and y directions, respectively. Equation (70) can be written in a simpler


form as

u~ TM][(R,)q
(72)
{K31}Um {K32}V TM 0- J

where the K matrix is banded and symmetric. Therefore, the matrix set is solved
using LU decomposition.
The system of equations can be numerically integrated for each element in the
flow domain using standard integration techniques. The integration across each
element can be summed to give the contribution of the velocity components,
stresses and pressure across the flow domain from the continuity and momentum
equations. This summation can be written as

NE
I F(x,y)dR = ~_, I F(x,y)dR (73)
R i=1 R,

where NE is the total number of elements, i is the element number from 1 to NE,
and Ri is the domain of the ith element.

3.2.1 Numerical Scheme for Finite Element Method


The method applied to solve this system of equations is a decoupled scheme
where the pressures and velocities are solved separately to the Giesekus
constitutive equation in relation to the extra stress terms. This constitutive
equation and the extra stresses are solved using a streamwise integration
technique introduced by Viriyayuthakom and Caswell [27] and Shen [161 ]. In
this technique, a Newtonian flow field is d e t e m ~ e d using the momentum and
continuity equations. Then the the stress terms _~j,k are determined using the
streamline integration of the constitutive equation and the known values of the
previously found velocity field. Next, the pressures and velocities are solved
again with the extra stress terms treated as a known variable. This iterative
procedure is repeated until a converged solution is found and the error has been
1040

minimized to an assumed value.


In the streamwise integration procedure, a streamwise coordinate s is introduced
which is positive downstream along each streamline. The constitutive equations
can be written in the form of

dv )r o~ 1 ak 2 2rhD (
=k = Vv-~ + ~ 9( Vv - ~ - ~ - v. VL - ~-~'k~ - 2Gk AkL + Ak----= F Vv,
[v_J--~-- (74)

by taking advantage of the convective term in the constitutive equation. The


resulting equations are a system of first-order nonlinear differential equations,
which are integrated using Euler integration. The integration is def'med as

x =x----'k + ~ - r l v v '~k ) (75)

(n+l) ,l(n)
where ----k
x and ------k denote the stress tensor at two consecutive points n+l and n,
respectively, and As is the arc length between the two points. The integration
starts at the upstream boundary and continues along the streamlines of the extra
stresses. This integration can be performed for any relaxation time ~,k as long as
the streamwise coordinate is chosen such that ~ <<~,klV_J. However, if the
relaxation time is small, the assumption is made that xlVv_J<<l, and a
perturbation solution is determined by

x = Gk[~+ ~,k(VV+ V : ) - 11
------'k
(76)

which is related to the Giesekus constitutive equation. The accuracy of this extra
stress tensor is directly related to the accuracy of the velocity and rate of strain
tensors.
This method does have difficulties when calculating the flow in the recirculation
zones since the streamline for a vortex is closed and does not start or end at an
inflow boundary. Thus, no initial conditions can be applied to solve this
situation. The streamlines in these flow regions are solved separately. The
assumption is made that the velocity components and strain rates are very small
compared to the flow, which drives the recirculation. Thus, from the assumption
that Z,k[VV_Jis small, equation (74)is solved for the stresses. If ~.k[VV_J is much
larger than unity, the stresses are determined after the stresses of some of the
1041

neighboring nodes have been calculated.


To apply the solution of the constitutive equation to the finite element method,
the extra stress tensor is represented by a 7-point Gauss quadrature as shown in
Figure 7, which has fourth order accuracy. For the triangular elements, three
points are related to the comers

Figure 7. Seven point Gaussian quadrature for a triangular element.

of the triangle, three points are related to the midsides of the triangle and one
point is related to the centroid of the triangle. The stress tensor from the
Giesekus model is evaluated at the Gauss points using the streamline integration.
Each streamline from an individual Gauss point is traced backwards until the
botmdary of the elements. From this point, the values of the stress tensor are
computed from the adjoining upstream element by interpolating quadratically at
the side of this element. With these values determined, the integration is then
performed along the trace path up to the original point. The elements in the
recirculation zone are determined using equation (76) or the streamline
integration as mentioned previously. At the walls the velocity is zero or defined
by the velocity of the moving boundary and the values of the stress tensor are
found from the constitutuve equation.
The application of appropriate boundary conditions is required in order to
numerically solve the system of equations. Boundary conditions required include
the velocity and stress field at the upstream boundary. Knowing these values, the
velocities and pressure field can be detennined and the streamwise integration
can start. The flow field upstream is assumed to be one dimensional, steady, and
fully developed. The equations ofmotion are written as

8P~ ~ -I- +col '~v3 -co ~r ~v~-~


2 k+. ,~'~ ' (77)
8Y2 9 8Y2 k=l tT~
1042

~9pm ~22~
- ~ (78)
~gy k=l 33'

and

zl,.~ = G,~ 1+ X, - 1 / *'12., = G, 4 1 + X, *~" = G, ~ / l + X , /


- 1 (79)

where Xk = 41 + 4"i'22L[ and ~ = 0 % . Integration of (77) in conjunction with (74)


results in

An' y = + to ~" - to ~"-' + ~'t',2,, (80)


k=l

where A m is defined as the pressure gradient Am=-/)pm//)x. The velocity is


found by assuming no slip at the boundary and integrating the equation

v:(y) = Ji~'dy (81)

In addition to the simple algorithm previously used to solve the viscoelastic


simulations, the numerical scheme implemented in the current simulations
incorporates the transient momentmn equations and the imposition of a moving
boundary. This numerical scheme is shown in Figure 8. First, the time step is set
at t=to. A mesh is then generated for the first flow domain for a fully open
channel. The initial values for the velocities, pressure are assumed, after which a
Newtonian solution is then found for the flow domain. Next, the constitutive
equation is solved using the streamline integration. With the solution of the extra
stress tensor found, the momenama and continuity equations are solved treating
this stress tensor as a known. This iterative procedure is continued until a
converged solution is found for the velocity, pressure and stress componems.
Next, the time and displacement is incremented to the next step. This includes
the movement of the boundary based on the imposed wall velocity for each time
step. A new mesh is generated for this flow domain. Assuming that the
displacement steps of the moving boundary are small between each time step, the
velocity, pressure and stress components from the previous time step are
reinterpolated into this new flow domain. The decoupled algorithm to solve for
1043

Initialization of v_~ pO, ~_oij~,

for first time step

Generatemeshinitial I

I on ions I
solve Newtonian flow field

Reinterpolate i f , pm, _.ffij~,for t~


Solve for_~k from constitutuve [ mesh for t~+~ time step to solve
equation using _v~'~ I transient process
"t
Apply boundary conditions I Generate new mesh for new
wall heil~ht
on v~l I
't
Apply time step tm+l = tm + At and
Find f f and p" from momentum
and continuity equations treating movement
of wall boun m+l = x m + Ax
ff~ as known

Converged i f , pm, ~ij~ for


~iven time step

If tm = t~ , numerical solution
finished

Figure 8. Flow chart of moving boundary.


1044

the field variables is then applied with the appropriate boundary and initial
conditions until a converged solution is found. This process is repeated until the
time and displacement limits have been reached. Then, the velocity components,
stress components and pressure are discretely known at each time step or position
of the moving boundary.

103 -

IMM
102-

model
Giesekus 9
101 ' '" .... I ' '"'"'1 ' '"'"'1 ' '"'"~ ' '"'"'1 ' '"'"'

10-1 10o 10~ 102 103 104 l0 s

7w, S 1
Figure 9. Viscosity versus shear rate at the wall for HDPE at 204.4~ along with
Cross and Giesekus model fits.

4. E X P E R I M E N T A L RESULTS

4.1 Material
The polymer used in the experiments is a commercial grade high density
polyethylene (HDPE) Marlex (Lot # HMN 4550-03) supplied by Phillips. To
rheologically characterize this material for the Cross and Giesekus models,
viscosity measurements were conducted on a Rheometrics mechanical
spectrometer (RMS-800) in the plate-and-plate mode in the low shear rate range
from 10 -1 to 10 2 S-1. For higher shear rates, experiments were conducted on an
Instron capillary rheometer (Model 3211) for the shear rate range from 101 to
3x103 s -1 and on an injection molding machine (Van Dora Demag 170 ton HT
series) fitted with a capillary type nozzle for the shear rate range from 2x103 to
4x104 s ~. Measurements have been made at three different temperatures
including 182.2, 204.4, and 232.2~ A plot of the viscosity curve for 204.4~
against the shear rate at the wall is shown in Figure 9.
1045

4.2 Methods of investigation


To test the numerical schemes implemented, a slit die with a moving boundary
will be used. The experimental setup is shown in Figure 10. The slit die has a
width over height ratio of ten and is attached to a Killion 1-inch extruder, which
will provide the flow of the polymer melt during the closing of the die. A screen
plate (not shown) is placed between the extruder and die to assure that the
pressure increase during the closing of the wall does not effect the flow rate or
screw characteristic curves. The pressure drop across the wall is relatively
insignificant when compared to the pressure drop across the screen plate.

,_ 8,89cm 2,54cm 7,62cm

Ptunger

E
z _j_u
Ft-ow- .N~e~io ~-.------4~ . . . . . . . . ~ CO
/ ,,/3

Figure 10. Experimental Setup for the Slit Die with a Moving Boundary.

To record the experimental data, two pressure transducers (Dynisco PT412-5M-


6/18 and PT435A-5C), as shown, measure the pressure drop across the die and
moving wall as the wall closes. The upstream pressure transducer is located one
inch in front of the entry of the wall region. The downstream pressure transducer
is located one inch from the exit of the wall. The movement of the wall is
controlled with an Instron T-500 tensile test machine and an LVDT (Schaevitz
Eng. Type 300HR S/N 3307) with a maximum displacement of 1.27 cm over a
volt range of 0 to 10 volts. The wall is 2.54 cm in length and 6.35 cm wide. This
wall varies the height of the die across the whole width of the wall. The height of
the flow channel is 0.635 cm with the wall fully open. To control the temperature
of the die, a melt probe is placed next to downstream pressure transducer. The
1046

temperature control components (model E5CS-R1K JX-F) are from Omron Tatesi
Electronics Co. Cartridge heaters (Watlow firerod 9651M G4J33) are placed
symmetrically in the die. In addition, a band heater ( Watlow thinband STB5 1J2
T) is placed around the screen plate and adapter plate, which changes the flow
geometry from the circular geometry of the extruder to the slit die geometry.
Two inch Owens Coming lnsulwool insulates the slit die and the die to extruder
adapter.
Several processing parameters are varied in these experiments including the
flow rate through die and the closing speed of the wall. A data acquisition system
(1 KHz) is used to record data from the pressure transducers and LVDT for wall
movement, which consists of an A/D converter module ADM 12-11 (Quatech)
with in house software and an IBM compatible PC computer. In the experirnemal
testing of the wall closing, the pressure drop across the moving wall is recorded
as the wall moves from a fully open position to 85% of fully closed. The speed
of the moving wall is held at a constant rate for each experiment. The flow rate is
also constant for each experiment. The experimental test conditions are shown in
Table I.

Table I
Extrustion and Die Experimental Conditions
Melt temperature (~ 204.4
Barrel zone temperatures (~ 204.4, 204.4, 204.4
Die Temperature (~ 204.4
Flow Rates (g/min) 20,40,70
Wall Velocity (cm/s) -0.0287, -0.071, -0.154,
-0.317, -0.635
IIII I

4.3 Experimental observations


Experimental results are presented for the transient flow of a polymer in a
channel with a moving boundary. The experiments were conducted for the three
different flow rates. For each flow rate, the velocity of the moving boundary was
given five different values. The only exception is the wall velocity of-0.635 crn/s
for the flow rate of 70 g/min. Results from these experiments show the time
dependency of the pressure drop due to the moving boundary. For all of the
experiments, the pressure increases as the gap height decreases, but some
interesting results are obtained due to the viscoelastic rheology of the HDPE.
The results for the experiments with the given flow rates and wall velocity are
shown in Figure 11, 12 and 13. The experimental pressure measured at the
upstream pressure transducer is plotted versus time on a log-log scale. The
1047

downstream pressure measurement is not included. The pressures measured at


this location are very low and do not add any insight into the understanding of the
effect the moving boundary has on the pressure.
Figures 1 l a, 12a and 13a show the actual velocity of the wall versus time for
each wall speed based on a specified flow rate. This velocity is not constant
during the entire experiment. Typically, the velocity of the wall decreases at the
end of each experiment before the wall reaches the stopping height of 85%
closed. For the faster closing speeds, a variation in the velocity occurs at the start
of the closing of the channel. The actual velocity of the wall is determined by the
slope of the LVDT versus time curves for each experiment.
For the flow rate of 20g/min with different wall speeds shown in Figure 1 l b,
each pressure shows an overshoot above the steady pressure during the closing of
the channel. The overshoot is greater for the higher wall speeds. In addition,
each pressure trace shows an initial jump in the pressure at the start of the closing
of the wall. In Figure 12b, the pressure traces for a flow rate of 40g/min show
no overshoot for closing speeds of-0.0287 and -0.071 cm/s. In addition, the
pressure traces without overshoot reach the steady state pressure after the wall
has stopped moving. This indicates a delay in the development of pressure. The
three fastest closing speeds do exhibit an overshoot in pressure during the closing
of the wall, although, the overshoot is not as pronounced as the overshoot for the
flow rate of 20g/min when comparing the magnitude of the overshoot to the
steady pressure for a given flow rate. These results for 40g/min show an initial
jump in the pressure at the start of the closing of the channel. For the flow rate of
70g/min, experimental pressure is presented for four different wall speeds. The
upstream pressure indicates no overshoot for any of the wall closing speeds. This
indicates a delay in the pressure development. These results for 70g/min also
show an initial jump in the pressure at the start of the closing of the channel.
The results indicate a relation between the closing speed of the wall and the
flow rate of polymer through a slit channel. If the flow rate is low relative to the
velocity of the moving boundary an overshoot in the pressure will occur. As this
flow rate is increased for a given moving boundary velocity, the magnitude of this
overshoot relative to the steady pressure will decrease. If the flow rate is
increased to a larger value, no overshoot in the pressure, over the steady value,
will occur and a delay in the development of the pressure will occur. Thus, the
pressure continues to increase to the steady value after the moving boundary has
stopped.
1048

0.0 - -

-~'-0.0287
-0.1 -
54
-0.2 -

-0.3 - J -0.317
> -0.4 -

-0.5 -

-0.6 -
[ . . . J V w = -0.635 cm/s
-0.7 ! ! | | w i ! | |

2 3 4 5 6 78910 20 30

Logt, s

le+7 - -0.635 J -0 317


t ! " -0.154

ea~
le+6 - ".UJ
]~/] ~ A"0"071 -0.0287

le+5

b
! ! ! ! ! ! ! [ !

3 4 5 6 7 8 910 20 30

L o g t, s

Figure 11. Actual moving wall velocity (a) and experimental (solid lines) and
predicted (dashed lines) upstream pressure (b) as a function of time at the flow
rate of Q=20 grams/min and various imposed moving wall velocities, Vw.
1049

0.0

-0.1 -
--

; __.r'-

-0.0287

-0.2 - ~~.0~17 "0"154


-0.3 -
> -0.4 -
-0.5 -
-0.6 - ~ , ~ Vw= -0.635 cm/s
-0.7 ! ! ! | ! | ! I
!

2 3 4 5 6 78910 20 30
Log t, s

le+7- "0"635 t ~'0"317~ "0"154 4"0-071 -0.0287


. . . . . . . ~. . . . . . . . .

le+6

le+5 b
w I

3 4 5 6 7 8 9 10 20 30
Log t, s

Figure 12. Actual moving wall velocity (a) and experimental (solid lines) and
predicted (dashed lines) upstream pressure (b) as a function of time at the flow
rate of Q=40 grams/min and various imposed moving wall velocities, Vw.
1050

0.0
J -0.0287
-0.1 - / - 5 ,.o.o71
-0.154
-0.2 -
-0.3 -
V w = - 0 . 3 1 7 cm/s
-0.4 -

-0.5 -

-0.6 -

-O.7 u u u u u u i
u

2 3 4 5 6 78910 20 30
Log t, s

le+7 -o.63s d
It I
-0.317
-0.154 -0.071 -o.o 1
/'
I
!
t~
!
/
le+6 ]
,J J

le+5 i "
i i i i i i i | I

2 3 4 5 6 7 8 910 20 30
Logt, s

Figure 13. Actual moving wall velocity (a) and experimental (solid lines) and
predicted (dashed lines) upstream pressure (b) as a function of time at the flow
rate of Q=70 grams/min and various imposed moving wall velocities, Vw.
1051

5. S I M U L A T I O N RESULTS

5.1 Material Parameters


From the rheological experiments for HDPE at 204.4~ the Cross and Giesekus
model were fitted to the viscosity data. The fitted model parameters for the Cross
model are 11o = 1500 Pa.s, x* = 8 x 1 0 4 Pa and n = 0.34. The density of
p=960kg/m 3 is assumed, which is typical for HDPE. For the Giesekus model, the
fitted model parameters are 111 = 775 Pa*s, r12 = 330 Pa*s, 113 = 39 Pa*s, 9~1 =
0.032 s, L2 = 0.0017 s, 9~3 = 0.000081 s, rls = 1.77 Pa*s and the characteristic
relaxation time of the HDPE melt according to equation (6) is 0.0221s. Figure 9
indicates that these parameters fit the experimental viscosity data quite well for a
wide range of shear rates.

5.2 Viscous Numerical Results


The numerical results are presented for the transient flow of a polymer in a
channel with a moving boundary. The computational package FLUENT is used
in the simulations using the Cross model. A quasi-steady approach to accotmt for
the moving boundary is used in this case. The movement of the boundary was
broken up into fourteen displacements. An inlet velocity based on the flow rate,
zero velocity of the stationary boundaries, the velocity of the moving boundary
and a pressure equal to zero at the die exit are the boundary conditions applied
for the simulations along with the appropriate material parameters.
To account for the moving boundary, two types of analysis are employed. The
first is an ideal velocity, where at a time less than zero the moving boundary
velocity is zero. At a time greater than or equal to zero, the moving boundary has
a velocity equal to the specified closing speed. When the moving boundary
reaches 85% closed the boundary motion instantaneously stops. The ideal
velocity is found by determining the first order slope of the LVDT when the wall
is closing the slit channel. These transient boundary conditions are written as

Vw= 0.0 for t < 0 and t > t85~

Vw -- Vclosing for 0< t < t85~

Calculations are conducted for the quasi-steady steps of the fully open position
and 85% closed position for a velocity of zero and with the specified closing
speed velocity. Thus, an instantaneous starting and stopping velocity is assumed.
The second type of analysis employed the actual velocity by determining the local
slope at each quasi-steady position. This variation of wall velocity during the
1052

closing of the slit channel is shown in Figure 1 la, 12a and 13a.
Results of the simulations of pressure traces with the ideal velocity are shown in
Figures 1 lb, 12b and 13b for the flow rates of 20, 40 and 70 g/min, respectively.
These figures indicate the pressure at each position during the closing of the
moving boundary. In all of the graphs, as the flow rate increases for a given
boundary velocity, the pressure increases, but not by a significant amount. As
the boundary velocity increases for a given flow rate, the pressure increase and
peak pressure increases drastically. This indicates that for a viscous quasi-steady
simulation the velocity of the moving boundary has a dominant effect on the
pressure as the moving boundary closes the slit channel. For the start-up and
stopping of the moving boundary, an abrupt change is seen in the value of the
pressure.
Results for viscous simulations with the actual wall velocity are shown in
Figures 14 and 15 for the various closing speeds and the flow rates of 20, 40 and
70 g/min, respectively. The simulated pressures show a sharp increase at the
start of the closing of the slit channel. This increase in pressure is not as drastic
in the case of the ideal velocity simulations since an instantaneous velocity is not
assumed. As the wall closed fiulher, the predicted pressure increases until the
end of the closing of the channel is reached. At this point, the pressure reaches a
maximum value. The location of this peak depends on the value of the wall
velocity at each quasi-steady position. Near the end of the closing simulations,
the actual velocity of the wall decreases rapidly. Thus, lower predicted pressures
are found when compared with the simulations which use the ideal velocity at the
end of the closing of the channel.

5.3 Viscoelastic Numerical Results


Viscoelastic numerical results are presented for the transient flow of a polymer
in a channel with a moving boundary. The simulations use in-house code which
is transient and accounts for the moving boundary. The Giesekus constitutive
equation is applied with the fully transient momentum equations. The simulations
include the elasticity of the polymer, the transient flow of the polymer and the
movement of the wall boundary into the flow domain. The flow rate, zero
velocity of the stationary boundaries, velocity of the moving boundary, steady
viscoelastic entrance and exit conditions and flow exit into a zero pressure region
are applied for the simulations along with the appropriate material parameters.
To account for the moving boundary, the viscoelastic simulations employ the
actual velocity of wall. Results of the viscoelastic simulations, which use the
actual velocity are shown in Figures 14 and 15 for the flow rates of 40, 20 and 70
g/min, respectively. A time step of 0.005s was chosen in the simulations. This
1053

le+7 I
-0.317 ~ r . -0.154 .~ -0.071

t~
le+6 -

, j
- 4 - - Viscous w/Vw= -Vw(t)
7
le+5 - Viscoelastic w/Vw= -Vw(t)
Experimental Data
I I I I I I I i

3 4 5 6 7 8 9 10
Log t, s

Figure 14. Experimental and predicted pressure as a function of time for a flow
rate of 40g/min and various wall velocities.

time step is capable of capturing the time-dependent response of the polymer,


which occurs for the multi-mode Giesekus model. The HDPE used in the
experiments exhibits a low relaxation time compared to the external time of
straining based on the flow rates and closing speeds. Thus, only the steady state
response and not the time dependent response of the second and third modes will
contribute to the predictions of the field variables. In addition, if the second and
third relaxation modes were included in the simulations, the computational time
would become excessively large.
The viscoelastic simulations that use the actual wall velocity are shown in
Figures 14 and 15. This velocity is similar to the velocity used in the viscous
simulations, except that a smoother pressure development occurs since a greater
number of time steps are included in the calculations. The simulations predict a
sharp increase in the pressure at the start of the closing of the wall. The pressure
gradually increases, until a peak pressure is reached near the end of the
simulation. However, the peak pressure is lower than that for the viscous
1054

le+7 I --~'- Viscous w/Vw= -Vw(t)


Viscoelastic w/Vw= -Vw(0
Experimental Data ,A
le+6
tm
O

a
le+5

! I I I I

3 4 5 6 7

Log t, s
le+7 -
- Viscous w/Vw= -Vw(t)
Viscoelastic w/Vw= -Vw(t)
9
_.l(~*,~ . . . .
9

r, le+6

o
-1

le+5 -- b
I I I I I

3 4 5 6 7 8

Log t, s

Figure 15. Experimental and predicted pressure as a function of time for a flow
rate of 20g/min (a) and 70/min (b) for the actual wall velocity with a set velocity
of -0.317 cm/s.

simulations that use the ideal and actual velocity. This pressure then reduces to a
steady value. After the wall has stopped moving the simulated pressure has not
reached the steady value as shown in Figures 14 and 15a. This would indicate
that a small amount of relaxation occurs in the viscoelastic simulations. In Figure
1055

15b, a large peak in the pressure development does not occur. The pressure
development for the viscoelastic simulation almost attain the steady pressure with
no overshoot.

6. COMPARISON OF NUMERICAL AND EXPERIMENTAL RESULTS

Figures 11, 12 and 13 are the comparison of simulated and experimental


pressure traces for the viscous simulations with the ideal velocity. Figures 11a,
12a and 13a are the experimental velocity of the wall during the experiments.
Figures l lb, 12b and 13b are the comparison between the simulated and
experimental pressure during the closing of the die channel.
For the predicted pressure drops using the Cross model with a quasi-steady type
simulation, the simulated pressures do not predict the experimental pressure very
accurately at the beginning or end of the transient flow experiments. In most
cases, for the flow rates in Figure 11, 12 and 13 the simulated pressures are
higher than the experimental pressure at the beginning of the closing of slit die.
At the start of the closing of the slit die, an initial jump is observed in the
experimental pressure. This experimental pressure jump occurs for the three flow
rates and the five closing speeds. The viscous simulation results predict a faster
pressure jump than the experimental pressure. This is true for all simulations
conducted. Then, the simulated pressure shows a much higher overshoot than the
overshoot of the experimental pressure. The simulated pressure instantaneously
goes to the steady pressure when the wall stops moving. However, the
experimental data indicates that the pressure does not reach a steady value
instantaneously, but after a short period of time reaches the steady value. The
lower peak pressure and longer time to reach a steady value after the wall has
stopped moving indicates the viscoelastic response of the HDPE. The lower peak
pressure indicates a delay in the development of pressure due to the viscoelastic
effect. The slow recovery to a steady value of pressure after the wall has stopped
moving indicates the importance of the relaxation of the polymer. This is very
interesting since the polymer has a relatively low relaxation time, but the time
dependent effect of the moving boundary results in the viscoelastic effect
becoming more pronounced.
For the flow rate of 20g/min in Figure 11, the predicted and experimental
pressure traces show overshoot at all closing speeds. For the flow rate of
40g/min in Figure 12, some more interesting fluid behavior is seen. For the wall
closing speeds of-0.0287 and -0.0711 cm/s, no overshoot is seen in the
experimental data, although, an overshoot is numerically predicted for all of the
1056

different closing speeds due to the quasi-steady approach taken and the elastic
effects are ignored in these simulations. When no overshoot occurs, the pressure
development is delayed and the steady value of experimental pressure is not
reached until after the wall has stopped moving. This indicates the viscoelastic
effects of the polymer are relevant in the time dependent experiments, which the
viscous simulations cannot predict. For the other three closing speeds, an
overshoot is seen in the experimental pressure, but the numerical predictions for
the peak pressure are much higher.
For the flow rate of 70g/min in Figure 13, the experimental pressure for the four
closing speeds do not indicate any overshoot. In addition, when the pressure
development is delayed, the steady experimental pressure is not attained until
after the wall has stopped closing the slit channel. In addition, the time to reach
the steady pressure value has a longer time span for the experiments with a
slower wall velocity. Part of this delay observed in the experimental data could
come from the slowing of the moving wall during the end of each experiment. In
addition, the ideal velocity used in the simulations does not take into account the
slowing of the moving wall. The simulated pressures predict an overshoot as in
the previous numerical results. This reinforces the fact that the viscous
simulations cannot predict the pressure accurately for the flow of a viscoelastic
fluid, even though the relaxation modes of the polymer are relatively small. Thus,
the viscoelastic effects appear to be extremely important in the flow of a polymer
through a slit channel with a moving boundary.
For the viscous simulations where the actual velocity of the moving boundary is
applied, the predicted pressures indicate different results when compared with the
experimental values as shown in Figures 14 and 15. The prediction of the
pressure during start-up compares well with the experimental pressure since
enough quasi-steady increments were chosen to approximate this region. The
simulated pressures show good agreement with the experimental pressure after
the start-up region except at the end of the simulations. As with the viscous
simulations with an ideal velocity, the simulations predict a higher pressure even
thought the actual velocity is employed in the simulations. Thus, the simulations
based on the quasi-steady analysis cannot predict the experimental pressure at the
end of the closing.
For the viscoelastic simulations where the actual velocity of the moving
boundary is applied, the predicted pressure shows different results. The
predicted pressure compares well with the experimental pressure at the start of
the closing of the channel. Then, the pressure gradually increases and reaches a
maximum and after that reduces to the steady value of the pressure. The
experimental peak pressure is not predicted accurately, although, the viscoelastic
1057

2 4 5
0.6 --
scale: -
l cm/
0.5 --

0.4 --

~ 0.3-
0.2-
0.1-
0.0 --
'''' '' '' I ' ' ' ' ' ' ''' I'""'''''''' I' '' ''' '' ' I '' ' ' ' ' ' '' I''

8 10 12 14 16
x, cm

Figure 16. Predicted velocity profiles at five different cross sectional areas of the
flow channel at Q=40g/min and Vw= -0.317 cm/s for viscous (solid lines) and
viscoelastic (dashed lines) simulations.

simulations that use the actual wall velocity predict a much lower pressure than
the other simulations. The predicted peak pressure is always higher than the
experimental pressure except in some instances after the wall has stopped closing
the channel. The pressure jump at the start of the closing of the channel is
predicted well since the actual velocity of the moving boundary is employed and
a higher amount of time steps are used in this analysis when compared to the
viscous simulations which used the actual velocity. The viscoelastic simulations
offer the closest prediction to the experimental data. When an overshoot occurs
in the experimental pressure as seen in Figures 14 and 15, an overshoot occurs in
the viscoelastic predicted pressure. In addition, the relaxation of the pressure due
to viscoelasticity of the polymer occurs in this simulation. When a suppression of
the pressure development occurs in the experimental pressure, very little
overshoot occurs in the predicted pressure. Unforttmately, the simulations were
unable to predict this suppression of the pressure development. The predictions
of pressure can be improved by using a smaller time step. This would allow the
simulation to include all three relaxation modes of the Giesekus model.
The velocity vector profiles of the simulations are shown in Figure 16. The v
and u velocity components are shown in Figure 17a and 17b, respectively, for a
flow rate of 40 g/min, and the second fastest closing speed of-0.317 cm/s. The
moving wall is eighty percent closed for these figures. Figure 16 indicates that
lO58

0.6 --
0.5- /
t~ 0 . 4 -
ff~ 0 . 3 -
0.2-
3
O.1- a

-0.50 -0.25 0.00 0.25 0.50 0.75 1.00 1.25 1.50


V, cm/s

0.6 --

0.5 i /5

0.4
if, 0.3 4
0.2
........ /3
0.1-
b
0.0 " '-'' ,;' '"":"' " -i ~ ' '

0 1 2 3 4 5
U, cm/s

Figure 17. Predicted profiles of v- (a) and u- (b) velocity component at different
cross sectional areas of the channel at Q - 40g/min and Vw- -0.317cm/s for
viscous (solid lines) and viscoelastic (dashed lines) simulations.

the variation in the velocity magnitudes across the gap are close for positions 1
and 2 and slight differences occur at positiom 3, 4 and 5 between the viscous and
viscoelastic simulations. It should be noted that the velocity magnitude at
position three in Figure 16 has a value greater than zero at the wall due to the
velocity contribution from the moving boundary. The velocity components in
Figure 17 indicate a good corr~adson between the viscous and viscoelastic
simulations for u-velocity components. However, some differences in the v-
velocity component are shown. This good agreement would indicate that the
1059

elastic effects of the polymer have little effect on the velocity profile of the
flowing polymer during the closing of the slit channel.
Thus, the viscoelastic simulations that use the actual moving boundary velocity
predict pressure better than the other simulations. The viscoelastic simulations
predict a lower pressure which is much closer to the experimental results. When
a suppression of the pressure occurs as seen in Figure 15b, the viscoelastic
simulation predicts a minimal amount of overshoot. When an overshoot occurs,
the viscoelastic simulations predict a lower pressure than the viscous simulations.
For the viscous simulations which implement the ideal velocity of the moving
boundary, the pressure development is poorly predicted. Thus, the variation in
the wall velocity has an important impact on the proper prediction of pressure
development. In addition, the predicted pressure overshoot is drastically higher
than the experimental overshoot of pressure. For the simulations where actual
velocity was employed, the viscoelastic simulations predict a lower pressure than
the viscous simulations. In both simulations, predicted pressures are lower than
in the simulations where the ideal wall velocity is used. The inclusion of the
elasticity of the polymer allows for a much better prediction of the pressure
development. The simulation predicts a pressure much closer to the experimental
data.

7. CONCLUSIONS

In the present study, experiments were conducted to measure the pressure drop
during the flow of a polymeric fluid through a slit die with a boundary moving at
a specified velocity, which is perpendicular to the direction of flow. The polymer
flows through the die as the wall closes the flow passage. The polymer used in
the experiments was HDPE. Experiments were conducted for three different flow
rates and five different moving wall velocities. A data acquisition system was
used to record the transient pressure drop as the wall closes the die channel. The
flow rate was held constant for each experiment. The velocity of the moving wall
was held constant, but was found to actually vary as the wall closes.
Experimental results indicate that for the flow rate of 20g/min an overshoot
occurs in the pressure as the wall closes the channel. For the flow rate of
40g/min, an overshoot does not occur for all of the closing speeds. This
overshoot occurs only for the two fastest closing speeds. For the flow rate of
70g/min no overshoot occurred for any of the wall closing speeds. When no
overshoot occurs, a delay is present in the development of the experimental
pressure. This delay was found to be a combination of the change in the velocity
1060

of the wall and a delay in the development of the pressure due to viscoelastic
effects.
The prediction of the experimental pressure was simulated using three different
types of numerical simulations. These simulations included viscous and
viscoelastic simulations that employed an ideal and actual wall velocity. The
viscous simulations used the generalized Newtonian fluid model according to the
Cross model. The viscoelastic simulations used the Giesekus model with a time-
dependent stress evolution. The viscous simulations were conducted with the
computational code FLUENT and used a quasi-steady approximation for the wall
movement. The viscoelastic simulations were conducted with in-house FEM
code, which is transient and accounts for the moving wall boundary.
The simulations that assumed an ideal wall velocity predicted a sharper increase
in the pressure development at the start-up of the wall velocity. In addition, these
simulations predicted a peak pressure which was drastically higher than the
experimental pressures. All simulations with the ideal wall velocity predicted an
overshoot in the pressure at the end of the closing of the channel. The
simulations which employed the actual wall velocity gave better predictions of
the experimental pressure. The viscous and viscoelastic simulations predicted the
start-up pressure very well. The prediction of the pressure overshoot for the
viscous case was much higher that the viscoelastic predictions. For the
simulations that used the actual wall velocity, a prediction of the pressure during
the start-up of the closing of the channel by the moving wall was found to be
accurate. The simulations, though, did not predict the delay in the development
of pressure. A small amount of overshoot in the pressure was predicted even in
the cases when no overshoot occurs in the experimental pressure traces. The use
of the actual velocity improves the prediction of the peak pressure. The
viscoelastic case was found to predict the transient pressure development better
than the other numerical simulations. The viscoelastic simulations which used the
actual velocity predict a lower pressure than the viscous simulations. This
indicates that the elastic nature of the polymer as described by the Giesekus
model helps in the prediction of pressure. These viscoelastic effects occur even
though the HDPE melt would seem to not exhibit a significant amount of
elasticity. The transient nature of the experiments make this elasticity become
more pronounced. For a better prediction of the pressure smaller time steps must
be implemented in the viscoelastic simulations.
The comparison of the velocity magnitudes and components across the gap
height for several different positions in the die at 80% closed indicate little
difference. Thus, the elasticity does not effect the velocity profiles. For the
simulations conducted, the viscoelastic FEM code can offer significant insight
1061

into the prediction of the experimental pressure. For polymers with higher
amounts of relaxation, this code could possibly help in the prediction of the
suppression of the pressure development. In these instances, a delay in pressure
could possibly be predicted. Research is currently being conducted on the
closing in the absence of flow through the slit or squeeze flow experiments. In
addition, polymers which have much higher relaxation times will be incorporated
into the current research on the channel flow of polymer with moving boundaries.

REFERENCES
1. A. I. Isayev, C. Zook, Y. Zhang and R.J. Scavuzzo, SPE Antec Tech. Papers,
41:1 (1995) 595.
2. M.J. Crochet, A.R. Davies and K. Waiters, Numerical Simulation of Non-
Newtonian Flow, Elsivier Science Publishers B.V., Amsterdam (1984).
3. C.L. Tucker, Computer Modeling for Polymer Processing, Hanser Publishers,
New York (1989).
4. M.G.N. Perera and K. Walters, J. Non-Newtonian Fluid Mech., 2 (1977) 49.
5. M.G.N. Perera and K. Waiters, J. Non-Newtonian Fluid Mech., 2 (1977) 191.
6. M.G.N. Perera and K. Strauss, J. Non-Newtonian Fluid Mech., 5 (1979) 269.
7. H. Court, A.R. Davies and K. Waiters, J. Non-Newtonian Fluid Mech., 8
(1981)95.
8. A.R. Davies, K. Waiters and M.F. Webster, J. Non-Newtonian Fluid Mech., 4
(1979) 325.
9. T.B. Gatski and J.L. Lumley, J. Comp. Physics, 27 (1978) 42.
10. P. Townsend, J. Non-Newtonian Fluid Mech., 14 (1984) 265.
11. P. Townsend, J. Non-Newtonian Fluid Mech., 6 (1980) 219.
12. G. Tiffenbruck and L.G. Leal, J. Non-Newtonian Fluid Mech., 10 (1982) 110.
13. J.H. Song and J.Y. Yoo, J. Non-Newtonian Fluid Mech., 24 (1987) 221.
14. H.C. Choi, H.J. Song, and J.Y. Yoo, J. Non-Newtonian Fluid Mech., 29
(1988) 347.
15. F.R. Phelan, M.F. Malone and H.H. Winter, J. Non-Newtonian Fluid Mech.,
32 (1989) 197.
16. X.F. Yuan, R.C. Ball and S.F. Edwards, J. Non-Newtonian Fluid Mech., 46
(1993) 331.
17. R.K. Upadhyay and A.I. Isayev, Rheol. Acta, 25 (1986) 80.
18. A.I. Isayev and R.K. Upadhyay, J. Non-Newtonian Fluid Mech., 19 (1985)
135.
19. A.I. Isayev and Y.H. Huang, Recent Adv. in Non-Newtonian Flows, 153
(1992) 113.
1062

20. M.B. Bush, J.F. Milthorpe and R.I. Tanner, J. Non-Newtonian Fluid Mech.,
16(1984)37.
21. X.-L. Lou and R.I. Tanner, J. Non-Newtonian Fluid Mech., 21 (1986) 179.
22. X.-L. Lou and R.I. Tanner, J. Non-Newtonian Fluid Mech., 22 (1986) 61.
23. X.-L. Lou and R.I. Tanner, J. Non-Newtonian Fluid Mech., 31 (1989) 143.
24. X.-L. Lou and E. Mitsoulis, J. Rheol., 34:3 (1990) 309.
25. M.A. Hulsen and J. Van Der Zanden, J. Non-Newtonian Fluid Mech., 38
(1991) 183.
26. J. Van Der Zanden and M. Hulsen, J. Non-Newtonian Fluid Mech., 29 (1988)
93.
27. M. Viriyayuthakom and B. Caswell, J. Non-Newtonian Fluid Mech., 6 (1980)
245.
28. S. Dupont, J.M. Marchal and M.J. Crochet, J. Non-Newtonian Fluid Mech.,
17 (1985) 157.
29. X.-L. Lou, J. Non-Newtonian Fluid Mech., 64 (1996) 173.
30. E. Mitsoulis, J. Vlachopoulos and F.A. Mirza, Polymer Eng. Sci., 24:9 (1984)
707.
31. E. Mitsoulis, J. Vlachopoulos. and F.A. Mirza, Polymer Eng. Sci., 25:11
(1985) 677.
32. M. Fortin and A. Fortin, J. Non-Newtonian Fluid Mech., 32 (1989) 295.
33. G. Barakos and E. Mitsoulis, J. Non-Newtonian Fluid Mech., 62 (1996) 55.
34. X.-L. Lou, J. Non-Newtonian Fluid Mech., 63 (1996) 121.
35. J. Sun, N. Phan-Thien and R.I. Tanner, J. Non-Newtonian Fluid Mech., 65
(1996) 75.
36. M.B. Bush and N. Phan-Thien, J. Non-Newtonian Fluid Mech., 16 (1984)
303.
37. M.B. Bush, J. Non-Newtonian Fluid Mech., 31 (1989) 179.
38. M. Kawahara and N. Takeuchi, Computers and Fluids, 5 (1977) 33.
39. M.J. Crochet and M. Bezy, J. Non-Newtonian Fluid Mech., 5 (1979) 201.
40. M.J. Crochet and R. Keunings, J. Non-Newtonian Fluid Mech., 7 (1980) 199.
41. M.J. Crochet and R. Keunings, J. Non-Newtonian Fluid Mech., 10 (1982)
339.
42. R. Keunings and M.J. Crochet, J. Non-Newtonian Fluid Mech., 14 (1984)
279.
43. A.R. Davies, S.J. Lee and M.F. Webster, J. Non-Newtonian Fluid Mech., 16
(1984) 117.
44. F. Dupret, J.M. Marchal and M.J. Crochet, J. Non-Newtonian Fluid Mech.,
18 (1985) 173.
1063

45. B. Debbaut and M.J. Crochet, J. Non-Newtonian Fluid Mech., 20 (1986)


173.
46. R. Keunings, J. Non-Newtonian Fluid Mech., 20 (1986) 209.
47. S. Musarra and R. Ketmings, J. Non-Newtonian Fluid Mech., 32 (1989) 253.
48. M.A. Mendelson, P.-W. Yeh, R.A. Brown and R.C. Armstrong, J. Non-
.Newtonian Fluid Mech., 10 (1982) 31.
49. T. Kajiwara, S. Ninomiya, Y. Kuwano and K. Funatsu, J. Non-Newtonian
Fluid Mech., 48 (1993) 111.
50. P.W. Yeh, M.E. Kim-E, R.C. Armstrong and R.A. Brown, J. Non-Newtonian
Fluid Mech., 16 (1984) 173.
51. R.A. Brown, R.C. Armstrong, A.N. Beris and P.-W. Yeh, Comp. Meth. Appl.
Mech. Eng., 58 (1986) 201.
52. J.M. Marchal and M.J. Crochet, J. Non-Newtonian Fluid Mech., 26 (1987)
77.
53. T.J.R. Hughes and A. Brooks, in Finite Elements in Fluids, eds. R.H.
Gallagher, D.H. Norrie, T. Oden, O.C. Zienkiewicz, Wiley, Chicester, 1982.
54. A.N. Brooks and T.J.R. Hughes, Comp. Meth. Appl. Mech. Eng., 32 (1982)
199.
55. B. Debbaut, J.M. Marchal and M.J. Crochet, J. Non-Newtonian Fluid Mech.,
29 (1988) 119.
56. B. Purnode and M.J. Crochet, J. Non-Newtonian Fluid Mech., 65 (1996) 269.
57. W.H. Hartt and D.G. Baird, J. Non-Newtonian Fluid Mech., 65 (1996) 247.
58. F. Yurun and M.J. Crochet, J. Non-Newtonian Fluid Mech., 57 (1995) 283.
59. R.C. King, M.R. Apelian, R.C. Armstrong and R.A. Brown, J. Non-
Newtonian Fluid Mech., 29, (1988) 147.
60. S.R. Burdette, P.J. Coates, R.C. Armstrong and R.A. Brown, J. Non-
Newtonian Fluid Mech., 33 (1989) 1.
61. P.J. Coates, R.C. Armstrong and R.A. Brown, J. Non-Newtonian Fluid
Mech., 42 (1992) 141.
62. D. Rajagopalan, R.C. Armstrong and R.A. Brown, J. Non-Newtonian Fluid
Mech., 36 (1990) 150.
63. G.P. Sasmal, J. Non-Newtonian Fluid Mech., 56 (1995) 15.
64. A.N. Beris, R.C. Amastrong and R.A. Brown, J. Non-Newtonian Fluid
Mech., 22 (1987) 129.
65. F.P.T. Baaijens, Comp. Meth. Appl. Mech. Eng., 94 (1992) 285.
66. F.P.T. Baaijens, J. Non-Newtonian Fluid Mech., 48 (1993) 147.
67. E.O.A. Carew, P. Townsend and M.F. Webster, J. Non-Newtonian Fluid
Mech., 50 (1993) 253.
1064

68. A. Baloch, P. Townsend and M.F. Webster, J. Non-Newtonian Fluid Mech.,


50 (1993) 253.
69. R. Guenette and M. Fortin, J. Non-Newtonian Fluid Mech., 60 (1995) 27.
70. J. Azaiez,, R. Guenette and A. Ait-Kadi, J. Non-Newtonian Fluid Mech., 62
(1996) 253.
71. A.C. Papanastasiou, L.E. Striven and C.W. Macosko, J. Non-Newtonian
Fluid Mech., 22 (1987) 271.
72. H. Holstein and D.J. Paddon, J. Non-Newtonian Fluid Mech., 8 (1981) 81.
73. G.G. Lipscomb, R. Keunings and M.M. Denn, J. Non-Newtonian Fluid
Mech., 24 (1987) 85.
74. Y. Kwon and A.I. Leonov, J. Rheol., 36:8 (1992) 1515.
75. A.I. Leonov, J. Non-Newtonian Fluid Mech., 42 (1992) 323.
76. D.D. Joseph and J.C. Saut, J. Non-Newtonian Fluid Mech., 20 (1986) 117.
77. D.D. Joseph, Viscoelasticity and Rheology, Academic Press, Inc., N.Y.
(1985).
78. R. Fielder and R.H. Thomas, Rheol. Acta, 6:4 (1967) 306.
79. N.D. Waters and M.J. King, Rheol. Acta, 9 (1970) 345.
80. J.S. Chong and R.G.E. Franks, J. Applied Poly. Sci., 14(1970) 1639.
81. J.S. Chong and D.M. Vezzi, J. Applied Poly. Sci., 14(1970) 17.
82. P. Townsend, Rheol. Acta, 12 (1973) 13.
83. G. Akay, Rheol. Acta, 16 (1977) 598.
84. B.R. Duffy, J. Non-Newtonian Fluid Mech., 4 (1978) 177.
85. B.R. Duffy, J. Non-Newtonian Fluid Mech., 7 (1980) 107.
86. R.T. Balmer and M.A. Fiorina, J. Non-Newtonian fluid Mech., 7 (1980) 189.
87. M.E. Ryan and A. Dutta, J. Rheol., 25:2 (1981) 193.
88. R.K. Upadhyay, A.I. Isayev and S.F. Shen, Rheol. Acta, 20 (1981) 443.
89. A.I. Isayev, J. Rheol., 28:4 (1984) 411.
90. R.W. Kolkka and G.R. Ierley, J. Non-Newtonian Fluid Mech., 33 (1989)
305.
91. R. Ketmings, J. Comp. Physics, 62 (1986) 199.
92. P.J. Northey, R.C. Armstrong, R.A. Brown, J. Non-Newtonian Fluid Mech.,
36 (1990) 109.
93. R.A. Keiller, J. Non-Newtonian Fluid Mech., 43 (1992) 229.
94. F. Olsson, J. Non-Newtonian Fluid Mech., 51 (1994) 309.
95. H.K. Rasmussen, O. Hassager, J. Non-Newtonian Fluid Mech., 56 (1995) 65.
96. S.A. White, A.D. Gotsis and D.G. Baird, J. Non-Newtonian Fluid Mech., 24
(1987) 121.
97. A.I. Isayev and R.K. Upadhyay, in Injection and Compression Molding
Fundamentals, A.I. Isayev Ed., Marcel Dekker, New York (1987).
1065

98. Y.H. Huang, PhD. Dissertation, University of Akron (1992).


99. R.B. Bird, Dynamics of Polymeric Liquids: vol. 1, Wiley, N.Y. (1987).
100. W. Oswald, Kolloid-Z., 36 (1925) 99.
101. A. de Waele, Oil Color Chem. Assoc. J., 6 (1923) 33.
102. C.A. Hieber and S.F. Shen, J. Non-Newtonian Fluid Mech., 7 (1980) 1.
103. C.A. Hieber, in Injection and Compression Molding Fundamentals, A.I.
Isayev Ed., Marcel Dekker, New York (1987).
104. C.A. Hieber, Rheol. Acta, 26 (1987) 92.
105. M.M. Cross, J. Colloid Sci., 20 (1965) 417.
106. M.M. Cross, Rheol. Acta, 18 (1979) 609.
107. C.A Hieber, V.W. Wang, K.K.Wang and B. Chung, SPE ANTEC Tech.
Papers, (1984) 769.
108. M. Sobhanie, J.S. Deng, and A.I. Isayev, J. Applied Polymer Science, Appl.
Polymer Symp., 44 (1989) i 15.
109. S. Deng and A.I. Isayev, Rubber Chemistry and Technology, 64:2 (1991)
296.
110. M. Sobbanie and A.I. Isayev, in Modeling of Polymer Processing, A.I.
Isayev Ed., Hanser, Munich (1991).
111. J.S. Deng, Ph.D. Dissertation, Univ. of Akron (1989).
112. J.L. Duda and J.S. Vrentras, Trans. of the Society of Rheology, 17:1 (1973)
89.
113. M.E. Kim-E, R.A. Brown and R.C. Armstrong, J. Non-Newtonian Fluid
Mech., 13, (1983) 363.
114. J.C. Maxwell, Phil. Trans. Roy. Soc., A157 (1867) 49.
115. R.G. Larson, Constitutive Equations for Polymer Melts and Solutions,
Butterwirth Publishers, Stoneham, MA, (1988).
116. J.G. Oldroyd, Proc. Roy. Soc., A200 (1950) 523.
117. H. Jeffreys, The Earth, Cambridge University Press, (1929) 265.
118. H. Giesekus, Rheol. Acta, 5:1 (1966) 29.
119. H. Giesekus, Rheol.Acta, 21 (1982) 366.
120. H. Giesekus, J. Non-Newtonian Fluid Mech., 11 (1982) 69.
121. H. Giesekus, J. Non-Newtonian Fluid Mech., 12 (1983) 367.
122. H. Giesekus, J. Non. Equilib Thermo., 11 (1986) 157.
123. J.Y. Yoo and H. Ch. Choi, Rheol. Acta, 28 (1989) 13.
124. A. Oztekin, R.A. Brown and G.H. McKinley, J. Non-Newtonian Fluid
Mech., 54 (1994) 351.
125. G. Schleiniger and R.J. Weinacht, J. Non-Newtonian Fluid Mech., 40
(1991)79.
1066

126. A.N. Beris, M. Avgousti and A. Souvaliotis, J. Non-Newtonian Fluid Mech.,


44 (1992) 197.
127. D. Vlassopoulos and S.G. Hatzikiriakos, J. Non-Newtonian Fluid Mech., 57
(1995) 119.
128. J.D. Schieber and J.M. Wiest, J. Rheol., 33:6 (1989) 979.
129. S.A. Khan and R.G. Larson, J. Rheol., 31:3 (1994) 207.
130. L.M. Quinzani, G.H. McKinley, R.A. Bronw and R.C. Armstrong, J. Rheol.,
34:5 (1990) 705.
131. T. Isake, M. Takahashi, T. Takigawa and T. Masuda, Rheol. Acta, 30
(1991) 530.
132. R.G. Larson, J. Non-Newtonian Fluid Mech., 23 (1987) 249.
133. M. Grmela and P.J. Carreau, J. Non-Newtonian Fluid Mech., 23 (1987) 249.
134. M.A. Hulsen, J. Non-Newtonian Fluid Mech., 30:1 (1988) 85.
135. M.A. Hulsen, J. Non-Newtonian Fluid Mech., 38:1 (1990) 93.
136. A.I. Leonov, Rheol. Acta, 15:2 (1976) 6.
137. A.I. Leonov, J. Non-Newtonian Fluid Mech., 25 (1987) 1.
138. N. Phan-Thien and R.J. Tanner, J. Non-Newtonian Fluid Mech., 2 (1977)
353.
139. N. Phan-Thien, J. Rheol., 22:3 (1978) 259.
140. R.B. Bird and R.J. DeAguiar, J. Non-Newtonian Fluid Mech., 12 (1983)
149.
141. J.R. DeAguiar, J. Non-Newtonian Fluid Mech., 13 (1983) 161.
142. M. Doi and S.F. Edwards, J. Chem. Soc. Faraday Trans. II, 74 (1978) 1789.
143. E.C. Bingham, Fluidity and Plasticity, McGraw Hill Book Company, New
York, N.Y. (1922).
144. W. Prager and K. Hohenemser, Uber die Ansatze der Mechanik der
Continua Zeitschrit~ fur Angewandte Mathematik und Mechanik 12 (1932)
216.
145. A. Slibar and P.R. Parsley, J. Appl Mech., 29 (1959) 107.
146. A. Slibar and P.R. Parsley, in Elasticity, Plasticity and Fluid Dynamics,
Haifa ( 1962).
147. J. Harris, Rheol. Acta, 6 (1967) 6.
148. T. Schwedoff, Congres de Physique, 1 (1900) 478.
149. J.F. Hutton, Rheol. Acta, 14 (1975) 979.
150. J.L. White, J. Non-Newtonian Fluid Mech., 5 (1979) 177.
151. J.L. White, J. Non-Newtonian Fluid Mech., 8 (1981) 195.
152. Y. Suetsugu and J.L. White, J. Non-Newtonian Fluid Mech., 14 (1984) 121.
153. S. Montes and J.L. White, J. Non-Newtonian Fluid Mech., 49 (1993) 277.
154. A.I. Isayev and X. Fan, J. Rheol., 34 (1994) 35.
1067

155. D. C-H Cheng and F. Evans, Br J. Applied Physics, 16 (1965) 1599.


156. D. C-H Cheng, Br J. Applied Physics, 17 (1966) 253.
157. Z. Kemblowski and J. Perera, Rheol. Acta, 18 (1979) 702.
158. A.I. Leonov, J. Rheol., 34 (1990) 1039.
159. M. Simhambhatla, PhD Dissertation, The University of Akron (1994).
160. P. Coussot, A.I. Leonov and J.M. Piau, J. Non-Newtonian Fluid Mech., 46
(1993) 179.
161. M. Sobhanie, A.I. Isayev. and Y. Fan, Rheol. Acta, 36 (1997) 66.
162. FLUENT User's Guide, Fluent Inc., Lebanon, NH, V4.2 (1994) 854.
163. S.V. Patankar, Numerical Heat Transfer and Fluid Flow, HcGraw-Hill, New
York (1980).
164. J.P. Maruszewski, Fluent Inc. Technical Memo, TM-049, 1991.
165. B.P. Leonard, Draft for First National Fluid Dynamics Conference,
Cincinnati, Ohio, July, 1988.
166. S.F. Shen, Int. J. Num. Meth. Fluids, 4 (1984) 171.
1069

FREE SURFACE VISCOELASTIC AND LIQUID


CRYSTALLINE POLYMER FIBERS AND JETS

S t e p h e n E. B e c h t e P , M. Gregory Forest b, Qi W a n g r
and H o n g Zhou b

Department of Aerospace Engineering, Applied Mechanics, and Aviation,


The Ohio State University, Columbus, Ohio ~3210
b Department of Mathematics, University of North Carolina, Chapel Hill,
North Carolina 27599-3250
r Department of Mathematical Sciences, Indiana University-Purdue Uni-
versity at Indianapolis, Indianapolis, Indiana ~6202

1. I N T R O D U C T I O N

We provide a cursory history of thin filament modeling of fibers and


fiber manufacturing processes (referred to as fiber spinning in the indus-
try); extensive references can be found in our published articles. Our goal
here is to identify those aspects of fibers and filaments which are accessi-
ble from slender I-D models, and which have been addressed in our own
work. As such, this chapter will be a biased account of viscoelastic and
liquid crystalline polymer fiber developments.
An important issue is a unified derivation of I-D fiber models. Different
industrial processes vary widely in the dominant competing physical effects.
For example, [1] analyzed the competition between inertia and Newtonian
viscosity; [2] considered viscosity and elastic relaxation. It is a fact of life
that one does not know a priori what the dominant physics in the slender
flow is, and moreover, it is highly likely that a slight change in the physical
competition may be either beneficial or detrimental. For this reason, we
have focused on a unified derivation together with a comparison of all the
various leading order physical regimes; a similar point of view has been
taken in [3] for Newtonian fibers where the number of possible dominant
balances is manageable.
For a wide range of inviscid [4], Newtonian viscous [5, 6, 7, 8, 9, 10],
1070

viscoelastic [11, 12, 13], and liquid crystalline polymer [14, 15] constitu-
tive relations, together with surface tension, gravity, and inertia, we have
developed a slender fiber perturbation algorithm that produces all one-
dimensional slender, axisymmetric, isothermal and thermal models in the
literature, as well as several relevant new models. This approach posits
a separable radial and axial structure at every order in perturbation, and
therefore has one major advantage and disadvantage. The advantage is
that the method is algorithmic, can be implemented with symbolic ma-
nipulations, and works whenever there exist separable slender flows. The
disadvantage is that the method does not address whether the separable
flows it describes are uniquely specified by the slender longwave approx-
imation. For Newtonian flows this issue is resolved (private notes). But
for more complex constitutive laws of the types considered here, it is an
open problem to construct slender fiber flows which are not captured by
our posited radial/asymptotic expansions. We have also addressed how to
select the specific regimes and corresponding model which best fits a par-
ticular experiment or process. This modeling feature is essential both for
direct simulations of a given experiment/process, and for inverse material
characterization [16, 17, 18]. We refer to our original published work for
details of these developments over the past decade.
For this volume we survey a selection of leading order models for increas-
ingly complex physical regimes. Similar to work by [19] and [20, 21], we
have exploited the mathematical structure of transient models to deduce
which quantities can, or must, be specified at upstream and downstream
boundaries of the fiber spinline. In viscoelastic [13, 22] and liquid crys-
talline polymer [14] fiber models we have discovered a novel space-time
change-of-type in which the governing equations for the fiber flow may
experience a transition from a well-posed hyperbolic structure to a catas-
trophic elliptic structure. We have also identified this transient change-
of-type with predictions for the onset of polymeric fiber failure [23]. The
mathematical structure of the I-D models for our selection of physical
regimes will further be shown below to reflect on the stability of two very
different classes of steady states" cylindrical jets and noncylindrical fibers
under tension.
1071

2. S L E N D E R F I B E R M O D E L S F O R I N C O M P R E S S I B L E I N V I S C I D , N E W -
T O N I A N , V I S C O E L A S T I C , A N D LIQUID C R Y S T A L L I N E P O L Y M E R F I B E R
FLOWS

We refer the reader to our journal articles for the 3-D, incompress-
ible axisymmetric free surface formulation and slender fiber perturbation
theory. The fluid rheology is modeled in the 3-D equations by the consti-
tutive (i.e. stress vs. rate-of-strain) relation one posits. The I-D models
we survey below arise from the following constitutive assumptions for the
stress t e n s o r T:

Inviscid Fluids:

T -- - p I , (1)

where p is the scalar pressure maintaining the incompressibility constraint,


I is the identity tensor.

Newtonian Viscous Fluids:

T = - p I + 2~7D, (2)

where r/is the zero strain rate viscosity, D is the rate-of-strain tensor given
by

[Vv + (3)
with V denoting the Eulerian gradient and v the fluid velocity.

Viscoelastic Fluids:

T - - p I + q?, (4)
where

+A~DT=2y D+A2 . (5)


z)t z)t
Here '~1 is the relaxation time, A2 is the retardation-time constant, which
is also called "polymeric viscosity", and

- (~-~ + v . V ) ( . ) + Ft(.) - (.)f~ - a ( D ( - ) + ( . ) D ) , (6)


1072

where a is a slip parameter ( - 1 _< a <_ 1) (a ~ 1 is considered the most


physical value, a = 1 is called the upper-convected Maxwell rate, a = 0 is
the co-rotational rate), and f~ is the vorticity tensor given by

[Vv - (Vv) ]. (7)

Liquid Crystalline Polymers:

T = - p I + 2r/D + To~, (8)


where
N I
To,. - 3ckO[(1 - --~-)q - N ( Q . Q ) + N ( Q " q ) ( q + ~)

I
+ 2A(Vv 9 Q)(Q + 5) ]. (9)

The orientational contribution to stress, To,., depends on flow and the


macroscopic orientation tensor, Q; a discussion of Q is given later. Here r/is
the solvent (Newtonian) viscosity; c is the number of polymer molecules per
unit volume; k is the Boltzmann constant; 0 is absolute temperature: N is
a dimensionless measure of the LCP density and characterizes the strength
of the intermolecular potential; and the parameter )~ is the relaxation time
of the LCP molecules associated with rotation of the dumbbell molecules.
The LCP constitutive law (8), (9) requires a dynamical equation for Q
as it couples to flow :
~ q - ( V v . Q + Q . Vv T) - F ( Q ) + G(Q. Vv),

F ( Q ) - - ~ { ( 1 - -~)q - N ( Q . Q ) + N ( Q " Q)(Q + 5)},


I
(lo)
I
G(Q, Vv) _ ~2 D - 2(Vv" Q)(Q + ~)
where a a is the drag or friction parameter taking on values between zero
and one; Crd = 1 is the isotropic case where the friction tensor is propor-
tional to the identity. As a d decreases, this corresponds to increasing the
ratio of resistance encountered perpendicular to the dumbbell axis versus
resistance along the axis; the limit a a = 0 is the most anisotropic limit
in which the friction is zero along the molecular axis. In equation (10)
note that F characterizes the orientation dynamics independent of flow,
1073

whereas G describes the flow-orientation interaction. These macroscopic


constitutive and dynamical equations are derived in [24] in the Doi closure
approximation [25] which allows one to decouple the molecular-scale prob-
ability distribution function. Later we recall an efficient representation for
Q.
We return to the general development: As we vary all possible com-
binations of physical effects which enter at leading order, for all regimes
which are self-consistent a typical structure emerges in, the leading order
I - D models in the axial coordinate z and time t:

Ndiag qt + Mf~u qz - f(q, qzz, b), (11)

where q is a d-dimensional vector of unknown functions of (z, t) which de-


couple at leading order, e.g. the leading order axial velocity v, free surface
radius r axial normal stresses J~zz, T~, and orientation order parameter
s; Ndiag is a d x d diagonal matrix; Mf~u is a d x d matrix; b denotes
body forces. In all I-D models which follow from this approach, the phys-
ical unknowns which decouple are those which are uniform in the fiber
cross-section to leading order.
In all the models presented below, the following dimensionless parame-
ters are used throughout:

R- pzg
r#o
(Reynolds number)

F - gt20z~ (Froude number)

W- proz~to
2 (Weber number)

A1 - ~1
to
(Weissenberg number)
(12)
A2 - &
to
(dimensionless retardation time)

Ca - ~,to =
rFo
R/W (capillary number)

Pe- pCz~
Kto
(Peclet number)

Bi - hro
K (Biot number),
where a is the surface tension of the material interface, p is the density
of the fluid, r0 and z0 are respectively the characteristic scMes in radiM
1074

and axial directions, to is a characteristic time, C is the specific heat per


unit mass, I( is the thermal conductivity, and h is the heat loss coefficient.
The ratio e = r o / z o is the s l e n d e r n e s s r a t i o , 0 < e < < 1, which is the
perturbation parameter in the asymptotic expansion.
In this section we present the governing equations and, where possible,
compute characteristic speeds. Solutions to these equations are presented
in later sections.

2.1 Model 1" Inviscid fibers dominated by surface tension and inertia
In the inviscid model [4, 6] the axial velocity v and the free surface ra-
dius r are decoupled at leading order, based on the inviscid constitutive
assumption (1)-
1 o v ~ r 0
v, 1-[o] (13)

The characteristics are given by


dz i
d t = 81,2 - v Jr v / 2 W ( ~ , i - x/~. (14)

Hence the model equations are elliptic, and ill-posed as an initial value
problem. This model has classical (Hadamard) catastrophic instability; it
results fi'om the longwave approximation of the free surface Euler equa-
tions and, of course, is not a property of the full 3-D equations. Refer
to [5, 13, 22, 23] for physical interpretations and cautionary remarks on
ill-posedness in slender models. Refer to [4] for similarity solutions which
yield predictions for the conical shape of an inviscid drop tip.

2.2 Model 2: Viscous fibers dominated by viscosity, surface tension and inertia
In this model, we promote viscosity and gravity to leading order. The
model equations are derived with the constitutive law (2):

0 r ~
__ 1
w
r
v
][v l I z ~
r 0
~zz+~r162
1 02 ] .(15)

This model, together with physically-motivated ad hoc regularizations, has


been applied extensively to the study of viscous drop formation [7, 8, 9,
10].
1075

2.3 Model 3: Viscoelastic fibers dominated by surface tension, inertia, viscos-


ity and relaxation (a Maxwell fluid)
In this model we promote elastic relaxation to leading order, so that
now surface tension, inertia, viscosity and relaxation are dominant effects.
The model equations are deduced from the constitutive law (4) and (5)
with A2 - 0. The axial (~'zz) and radial (Try) extra stress components now
couple to the free surface radius (r and axial velocity (v) at leading order.
Letting u - (r v, Tzz, T~), we find
N ut + M Uz - f ( u ) , (16)
where
N -- diag[0, 0 , - T z z , - J ] , . ] ,

f - (0, O, -Tzz, - T ~ ) , (18)


v -~
2 0 0
1 (/)2
w v -B B
M- 0 AI(1 - 2a)Tzz - 2Z r A1 v 0 ' (19)
0 AI(1 -[- a)~'rr -]- Z r 0 A1 v

where
B- fot (20)

Z = qr~
tofo' (21)

f0 is a characteristic force, and B-k2is the Reynolds number.


The characteristics of the quasi-linear first order system in Model 3 can
be calculated explicitly:
dz [ v j-1,2
dt = s j - / v-4-~ j - 3,4. (22)

Here A is given by
1
/ k _ A I R 3 r q_ B (J'zz + 2 rrr) - - ~ r (23)

in which the first term (AI-~ r is the "gain" from high viscosity, the second
term (B (Tzz + 2 2~r)) is the "penalty" for compressive stresses (for which
1076

it is negative) , and the third term ( T~


1 oF) is the "penalty" for high surface
tension.
The system (16) is genuinely hyperbolic if and only if A > 0, indicating
well-posed evolution. In particular, the system (16) is:

9 Hyperbolic type (3, 1) (meaning 3 positive and 1 negative character-


istics) if and only if A > (Ov) 2, where s'l, s2, s3 > 0; 5"4 ~ 0.

In this case, there are exactly 3 upstream boundary conditions and 1 down-
stream boundary condition. Therefore, the system is well-posed for mod-
eling fiber spinning processes as long as this inequality on the solution
is maintained. Moreover, the Riemann variables prescribe what process
quantities may be imposed upstream and downstream.

9 Hyperbolic type (4, 0) if and only if 0 < A < (~v)2 where all four
characteristic speeds sj are positive.

In this case all information moves downstream and one can only impose
upstream boundary conditions. This same system therefore describes ex-
trusion flows as long as the above constraints are satisfied.

9 Mixed elliptic/hyperbolic type if and only if A < 0, indicating ill-


posedness.

The equations are not applicable for dynamical evolution in this regime.
The boundary between well-posed hyperbolic evolution and ill-posed
mixed-type behavior is A = 0, which is not an invariant condition (i.e. the
sign of A may change along exact solutions). [22] studies families of exact
model solutions which yield all varieties of hyperbolic to elliptic/hyperbolic
transitions. In Figure 1 we show one possible space-time boundary of this
well-posed/ill-posed transition, corresponding to the exact model solution:

(~ - - r V -- VO~

Tzz(t, z ) - e-t/AiAo(z- vot), (24)

T~(t, z) -- e-t/a~[Ao(z -- vot) + A~


where A ~ qSo, vo are constants and the function Ao(s) is arbitrary except
that Ao(s) =/=O , - A ~
1077

10-

61

2 4 6 8 10

Figure 1" The regions in the (z,t) plane of hyperbolic and mixed ellip-
tic/hyperbolic type which evolve from the Cauchy d a t a with the discrimi-
nant A - A ( e c ) + e-88 v0t), where v0 - A1 - 1, A(z) - 2 + sin(z) _>
0, ~ >- 0, z x ( ~ ) - 2wA~
~o ( ~- A
o 1) _ - 0 . 5 , so t h a t the system is hyperbolic
at time t - 0 for all z _> 0. A(ec) - - 0 . 5 implies t h a t the system becomes
mixed elliptic/hyperbolic as t --+ oc for all z > t. From [22].
1078

2.4 Model 4" Viscoelastic fibers dominated by surface tension, inertia, viscos-
ity, relaxation and retardation (a Johnson-Segalman fluid)
When an elastic retardation effect is included, the extra stress can be
decomposed into two parts,
A2
rr _ Tp + 2r/~ll D (25)

where Tp is the polymer part of the stress, A2 is the retardation time and
2 q ~ D is the Newtonian part of the stress. Substituting the decomposition
into the generalized Oldroyd Fluid-B model, the constitutive equation (5)
becomes a Maxwell-type equation for Tp,

rrp 4- )kl - ~ = 27/(1 - ~11 D. (26)

This is the Johnson-Segalman formulation for viscoelastic flows. From this


formulation, the leading order model is a 4 x 4 degenerate parabolic system,
involving the same unknowns as in Model 3:

ut + C(u)u~ = f(u) + Duzz, (27)


where u -- (4), v, 7"zz, ~,.,.)t,

v ~-
2
0 0

1
+ Tzz) -B B
C(u) - (28)
0 -2a~z - 2Z(1-~)_A~ v 0 '

Z(1-A1) 0 V
0 a Z , r + " A1

0 0 0 0 0

0 3 B Z ~A1 0 0 T1 + 6 BZ-~11dz
A,,
Vz
D f(u) - . (29)
_T_~
O0 O0 A1

0 0 0 0 T~
A1
1079

All the physical parameters are the same as before. Due to the decompo-
sition of the extra stress, the range of A2 is restricted to 0 _< A2/A1 _< 1.

2.5 M o d e l 5: L C P inviscid fibers d o m i n a t e d by i n e r t i a , s u r f a c e t e n s i o n , gravity,


a n d m o l e c u l a r - s c a l e o r i e n t a t i o n effects" p o l y m e r k i n e t i c e n e r g y , an e x c l u d e d -
volume potential, and anisotropic drag
From [14, 15] the uniform leading order behavior in the LCP fiber cross-
section is given by the special uniaxial representation for the orientation
tensor Q:
1 1 2
Q-sdiag[ 3' 3' 3 ]. (30)
where s is the scalar order parameter, which is restricted to the range
1 < s < 1 From this representation, the slender fiber equations for this
2 ~ ~ "

regime take the form [14]:

ut + C ( u ) U z - f, (31)

where u - (0, v, s) t,, f - (0, 1/F, -ad/AU(s)) t, and


v r2 0
6"~.(U ~ __ 1 2aU(s) V -~U'(~) (32)
'~.w't ]
we ~ r
o -h(~) v

The eigenvalues of the coefficient matrix, which determine the characteris-


tic speeds, are
~1 m V~

~,3 - v + qA(~, r ~, w, m),


(33)
1
A(,. r ~. w. - ~[h(.)U'(.)
N) - u(.)] 2Wr
- ~[h(~)U'(~)- u ( ~ ) - c_~],
~<~

where f U(s)ds is an effective intermolecular excluded-volume potential,


with

U ( s ) - s[1 - --~ N(1-s)(2s+l)], (34)

h(s)-(1-s)(2s+l), (35)
1080

3ck0
(36)

which parametrizes the molecular kinetic energy per unit volume relative
to inertial energy per unit volume; and Catcp is the LCP capillary number,
(7
= 3ck07----- (37)

The discriminant A(s, ~b; a, W, N) may be positive, zero, or negative


depending on the solution variables s and ~, and depending on the param-
eter values a, W, N. In particular, this regime has the possibility to evolve
from a well-posed hyperbolic initial-boundary value problem into an ill-
posed elliptic-hyperbolic region at a critical space-time location, signaling
a rapid departure from slender fiber behavior. Note the strong similarity
between this LCP model system (31)-(37) and the Maxwell model system
(16)-(23).
When A > 0 the quasilinear system is strictly hyperbolic. W h e n / k - 0
the system is degenerate, with three identical characteristics and only two
independent eigenvectors. When A < 0, then the system is of mixed
elliptic-hyperbolic type, and is therefore ill-posed as an evolutionary sys-
tem.
In the context of general solutions, the sign of A varies through a com-
petition of the solution variables s, ~; at particular values of s, ~ the hy-
perbolic versus elliptic behavior is governed by the parameters N and Cal~p.

2.6 M o d e l 6: T h e highly a n i s o t r o p i c d r a g limit of M o d e l 5


This model is formally obtained fi'om Model 5 by letting ad = 0 in
the leading order equations; refer to [14]. The nonlinear classification in
Model 5 is unchanged, since the terms proportional to ad only contribute
to the nonlinear lower order terms in the equation for s. Thus, the criteria
for hyperbolic versus mixed hyperbolic-elliptic type are as given in Model
5. In this model one loses the selection of discrete branches of equilibria
determined from the zeroes of U(s); the intermolecular potential is effec-
tively constant in this limit so all values of s are allowable equilibria. The
parameter values for which A(s0) < 0 correspond once again to ill-posed
equilibria, as discussed in Model 5.
1081

2.7 Model 7" Viscous generalization of Model 6


This model couples viscous terms, both Newtonian and orientation-
dependent, to Model 6, retaining its highly anisotropic drag limit (ad = 0)
[14]
The leading order system is given by
r + V~z + ~Vz/2 = o,
_ 1(~2 1 (38)
(02~), + (02~)z ~ + ~r + ((,7~sf<~ + ~u(~))r
st + VSz -- Vz(1 - s)(2s + 1),
where fluff(s) is a I-D effective viscosity consisting of a Newtonian elonga-
tional viscosity (gr/) and an orientation-dependent contribution (2a)~s2),

~ss(~) - a~ + 2~,~ ~. (ag)


2.8 Model 8" The most general L C P fiber regime, with a r b i t r a r y drag, L C P
relaxation, inertia, viscosity, surface tension, gravity, and L C P kinetic energy
We now couple arbitrary anisotropic drag effects to Model 7.
Let u - (r v, s) t. The leading order equations can be written as
N d u t + Muz - G(u) + (Uz, Cuz)(0, 1,0) t + 'q~ffr 1,0) t, (40)
where
Nd -- diag[1, O2, 1], (41)

G(u) - (0, s (42)

o r o

C Cr/~SS(,) 0 2(~Asr 2 , (43)

0 2a)~sr 2 0

v r2 0

M 1
W
2ogU(8) ~ V(~2 - o z U ' (8)(~ 2 (44)

0 - ( 1 - s ) ( 2 s + 1) v

The form (40) amplifies various properties of the equations:


1082

i) The Nd, M and G terms correspond to quasilinear 1-D behavior along


dz -- aj, where aj are the eigenvalues of N d l M .
characteristics -3-i

ii) An indefinite nonlinearity in derivatives of u appears in the second


(axial momentum) equation, in the way of a quadratic form with sym-
metric off-diagonal C. If this quadratic form is diagonalized, one finds
the sign of these nonlinearities in derivatives is indefinite.

iii) A degenerate viscous regularization term, rl~llr appears only in


the axial momentum equation. This yields a viscous Burgers equa-
tion for the axial velocity v with additional nonlinear derivative terms,
coupled to quasilinear equations for the free surface r and order pa-
rameter s.

2.9 Model 9: Thermotropic LCP fibers


In this model we couple temperature (0) through an energy equation.
We refer to [26] for the derivation, and a discussion of the application of this
physical regime. The leading order equations in the melt-phase including
all effects in Model 8, are:
(r + (v62)z _ 0,

(r + (O2 2)z - + 1 + ( ~2 0)v: +


(45)
st + VSz - Vz(1 - s)(2s + 1) - Ac0)U(s),

Pc(Or + vOz) __ -1~ (~2 2Bi ( 0 - 1 + A0)


O z ) z - --~
where A(0) is the scaled LCP relaxation time,
A ( 0 ) - Ae~(a/~ (46)

R~ff(s, O) is an effective 1-D flow-orientation Reynolds number,


R e ; f ( 8 , O) -- 3 egCl/~ -~- 2 ol 0 A(0) 82, (47)

consisting of a Newtonian contribution and an orientation contribution


(2o~OA(O) s2); and U ( s ) - s ( 1 - N / 3 ( 1 - s)(2s + 1))defines the uniaxial
bulk free energy, f U(s)ds.
The dimensionless parameter a characterizes the molecular stress rel-
ative to inertial stress; ad/A parametrizes anisotropic drag on polymer
1083

molecular motion relative to the solvent; N is a polymer density parameter


that enters prominently in the intermolecular excluded-volume potential;
Pe is the Peclet number which is a measure of specific heat relative to
thermal conductivity; Bi is the Biot number characterizing the heat loss
relative to thermal conductivity; g is the Griffith number quantifying the
degree of viscosity variation with temperature. The nonisothermal param-
eter A0 is chosen as

AO -- 1 -- Oa/Omelt , (48)
which is a measure of the degree of nonisothermality. When Oa -- Omelt,
then A0 = 0 and our model reduces to the isothermal case. In [26] we
have varied this parameter to study the energy effects on our previous
isothermal LCP spin model predictions [15].
One can also formulate (45) into the general vector form (11) (see [26]).

3. L I N E A R STABILITY OF CYLINDRICAL JETS

Insight into the structure of the 1-D slender fiber models of the pre-
vious section can be deduced by linearizing the models about constant
cylindrical jet solutions when they exist (i.e. when gravity is negligible).
This calculation can then be compared against Lord Rayleigh's analysis of
the capillary instability of circular cylindrical jets [27]. The comparison
clarifies the nature of the slender, longwave model for each selection of
dominant physics.
First we recall exact 3-D results. Another discussion along these lines
may be found in [28]. Rayleigh's [27] linear stability analysis of a cylindrical
inviscid jet with radius r0, constant axial velocity v0, surface tension a, and
constant mass density p shows that an infinitesimal perturbation of the jet
free surface of wavelength 27r/k grows exponentially with rate

(49)
7(k)- pr~ Io(kro) '
where I0 and I1 are the modified Bessel functions of the first kind of order
0 and 1, respectively, indicating a longwave instability for all wavenumbers
k such that 0 < k < 1/ro, with a finite cutoff at the wavelength of the jet
circumference 27~r0.
1084

To compare this dimensional result with slender asymptotic growth


rates, we nondimensionalize the wavenumber k and growth rate 7(k) as
follows"

k--, ~(k) - ~ , (50)


zo to
which yields

I1 k (1 - ~'2d)I1(~)
~n,(k)- W e Io(k'e) ' (51)

where

- ro/zo (52)
represents the radial to axial aspect ratio and W is the Weber number.
In nondimensional form, the exact linearized growth rate a/vi~c(k) for a
viscous fluid jet is described by the transcendental equation [29, 30]

(7~visc)2 ~(]r Jr- ")/visc~~2 [2 r - 1]


(53)
--- 2T4/(1 -(kc) 2) - ~2~'4 7(~;~)- f(~,~) ,

where
7 ( : ) - : i0 (~-)
/l(Z) ' (54)

(~5)2 __ (~6)2 -t- ZYvisc /~ ~2. (55)

Looking forward to a comparison of slender 1-D models and these exact


3-D formulas, consider the slender longwave limit Ik,el < < 1. The exact
inviscid rate (49) has the expansion

%,~,(~) - v/g ~ + o(~2/~2), (56)

while the viscous rate has the expansion [30]

3~:2 / [3~2, 2 ~2
#visc(k) -- 2R + ~ [ ) - ~ + ~ + ~ (57)
1085

| . . . . | .

3.5 Inviscid,..E'~

xact
e2.5
c-
2
o
c,1.5 er . . . . . . . . . . . .

0.5

Oo 2
I

4
I I

6
I

8 10 12
k

Figure 2: Nondimensionalized Rayleigh growth rate versus wavenumber k


for inviscid or viscous jets.

The exact inviscid Rayleigh growth rate, together with the exact viscous
Rayleigh growth rate, the leading order approximate inviscid growth rate
(56) and the leading order approximate viscous growth rate (57) where
W - 1.0, e - 0.1, R - 10, is provided in Figure 2. As shown in Figure 2,
the inviscid and viscous growth rate share the same cutoff.
Now we investigate linear stability of constant cylinder equilibrium so-
lutions to all leading order slender fiber models presented in the previ-
ous section (with ~1 __ O) and display how these models approximate the
Rayleigh instability. The constant solutions always have radius r - 1,
axial velocity v0 - 1, with the remaining constant unknowns specific to
the model. For simplicity, we drop the superscripts. We posit that the
perturbation 5u of the constant solution is of the form
(~U ~ r
c, (5s)
where c is a constant vector. The linearized growth rate is given by the
real part of u, 7 - Re(u). The I - D linearized growth rate of each model
is given as follows (recall all the tilde overbars have been dropped)"
1086

Model 1" Inviscid fibers dominated by surface tension and inertia:

,7 nv(k) - k ( 2 w ) (59)

Model 2" Viscous fibers dominated by viscosity, surface tension and inertia:

~/visc (~) -- 3k2


2R +
I k?-R]
(3k2~2 k2 (60)

Note that these slender model growth rates agree, as they must, with
the leading order slender longwave approximation of the exact 3-D inviscid
and viscous growth rates (56) and (57) (Figure 2).

Model 3: Viscoelastic fibers with Maxwell fluids:

The growth rate of Model 3 is given by a fourth-order algebraic equation


[22] for the four values 7j(k) (j - 1 , . . - , 4). There is one trivial factor of
this characteristic polynomial, 71(k) - -1/A1, corresponding to linearized
exponential decay at the relaxation rate (A1)-1. The three remaining rates,
7j(k), k - 2, 3, 4, are the roots of the following cubic polynomial:
1 2 k2 k2
7 3 + ~117 + 2WAI - =0, (61)
2WA1
where
6
A= A1. (62)
Ca
Note that the sign of the linear term in 7 fixes the characteristic type of
the constant solution. Application of Descartes's rule of signs shows there is
always one positive real root, which we denote 72(k), of (61), independent
of the sign of the linear term, while the two remaining roots V3,4(k) are
complex with negative real part. A simple asymptotic balance for k ~ ec
determines the behavior of 72 (k) :
A -1 A>O as k---~ oo,
/A 6w
72(k) ~ k~/2w~ A < 0 as k ~ oo, (63)
k2/3( i2w
n 2) 1/3 A - 0 as k ~ o o .
1087

Model 4" Viscoelastic fibers with Johnson-Segalman fluids:

As in Model 3, there is one linear factor:

,7,~odd4(k ) _ _ 1 . (64)
A1
The three remaining growth rates satisfy the cubic polynomial equation:
1 3k2A2 k2 6 k2
~3 _[_ (All -+- AIR )3/2 4- 2WA1 (~aa - A1)~/- 2WA1 -- 0. (65)
The growth rates in the range of short waves are given by

~/Js "
( i
(WA1) -1 m -~- A 2 -~- RWA2 -t-O(]r A ~_ 0 for ]~-1 e,o 0
(66)
a~ ( - A
6RA2 + j A2 + RWA~ + O(k- 1) A < 0 for k -1

whereA-- 3 1 andR- 1
RA1 2W BZ"
In Figure 3 we plot the growth rates for the leading order inviscid, vis-
cous Rayleigh results and the growth rates for Maxwell, Johnson-Segalman
models with small (Ca = 1.0) and large (Ca = 10) capillary numbers Ca.
The remaining parameters are A1 = 2.0, A2 = 1.0, R = 10.0, e = 0.1,
and the Weber number is determined by the Reynolds number and the
capillary number through W = R/Ca.

Model 5: LCP inviscid fibers with anisotropic drag"

With a nonzero anisotropic drag parameter present at leading order, there


are three branches of constant solutions parametrized as in Figure 4. The
linearized growth rates are given by the roots of the cubic equation to
follow, which we present in a suggestive form with the drag terms isolated
on the right hand side:

,),3 + / k ( s j ) k 2 ~ _ ad U' (sj)(k2(2W)-i _ .y2),


A (67)

where A(sj) for equilibrium solutions is given by (where for consistency


r 1)
Calcp
A(sj) -- c~h(sj)U'(sj)- (2W) - 1 - ct h(sj)U'(sj) 2 ' (68)
1088

u m . m i. m m 9

Inviscid Leading Order,'


3.5 ." / 9 /
9 /

9
..'" ,'Maxwell (large Ca)
/
9 /

~2.5
-I~ ."
-'/"/ "
9 /

x: 2
o
o~1.5

0.5
m m m i i m
O0 2 4 6 8 10 12
k

Figure 3: Nondimensionalized growth rate versus wavenumber k for lead-


ing order inviscid or viscous, Maxwell or Johnson-Segalman jets. The
lower part of the figure corresponds to Maxwell model with small capillary
number (dash-dotted curve), Johnson-Segalman model with small capil-
lary number (solid curve), and the leading order viscous Rayleigh result
(dotted curve), respectively.

and sj (j - 1, 2, 3) are zeros of U(s)given by


s~-0, s2,3-~ 1+3 1-~-~ . (69)

The longwave and shortwave asymptotic behavior of (67) for Model 5


are"
3'21 ~ -i-k(2W) _ ~1 - ~ A k2 for k ,,~ 0
' ~ ' ' (70)
"~3 ~ - ~ u ' ( s j ) aAh(sj)k2 for k ~ O,
+ ~
and
I ")'3,1 e,o -t-~q--/k(Sj)- ~.~)h(sj)U,2(sj) for k-1 ~
2AA(sj '
0, (71)

~,~ ,.., "2WA


--~-U'(~;)(A(~;)) -1 - F (9(k-2), for k -1 ,-., o.
1089

0.5
U/>0
i
i

~ U ~ <0
$1 Ok
o

U'>O 8/3 \ 3 U'< 0

-0.5
U'>O

, , , , I , , J , I , , , , I , , , ' I , ,' , ,

-10 2 4 6 8 10

Figure 4: Real order parameter equilibria s versus LCP density N. The


solid curve indicate where U(s) = 0 and U'(s) is positive while dashed
curves are where U(s) = 0 and U'(s) is negative. There is a turning point
at N = 8/3 where two real equilibrium branches s2 and sa are born. At
N = 3, s2 collides with Sl and U'(sl) and U'(s2) change sign, which shall be
responsible for transitions in the orientation-dominated linearized modes
as we analyze stability of these equilibria.
1090
Model 6: LCP inviscid fibers with highly anisotropic drag:

In this model with Od = 0, any equilibrium order parameter so is allowed


in - ~1 <_ so <_ 1. However, for comparison with all other models where the
critical points of f U(s)ds are the selected equilibria, we only analyze these
discrete equilibria.
The growth rates are given by (67) where ad = 0:
(72)
"/2,1-Ar-kRe(~/-A(8j)) 9 (73)
Clearly, when U'(sj) < 0 (see Figure 4), then A(sj) < 0 and the equi-
librium is catastrophically unstable. This is, of course, consistent with
the classification earlier that shows the full nonlinear equations are elliptic
when A < 0.
When U'(sj) > 0, then A(sj) > 0 only if Catcp is sufficiently small,
Catcp < 2h(sj)g'(sj). In such cases (in particular for the prolate phase,
s2 (1 + 3x/1- 8/(3N))/4 ), when LCP surface energy dominates sur-
-

face tension, then the growth rates are identically zero, indicating neutral
stability.

Model 7" LCP viscous fiber with highly anisotropic drag"

Now this model couples Newtonian and effective LCP viscosity to the pre-
vious model. Intuitively, one expects a "viscous regularization", i.e., the
previous model growth rates should be lowered.
The linearized growth rates 71,2,3 are easily calculated in closed form:

~1 ,2 - - -~(--r]eff(80) k2 -j" ~ff(so)


2 k4 - 4A(so)k2), (74)

"73 - O, 74 - O, "/5 - O. (75)


Again we evaluate so at the critical points of the 1-D Maier-Saupe poten-
tial, sj.
The vanishing of 73 reveals the orientation instability, realized through
growth in the uniaxial order parameter s, is also suppressed when ad -- 0
at leading order, i.e., the highly anisotropic drag limit.
1091

We turn now to the Rayleigh-dominated rates. Note that ~/1 + ~2 -


-k2rl~ff(sj) < 0, and
~1~/2 - - k2m(sj). With regard to stability, the sign of
A(sj) is critical:

If A(sj) > 0 then Re(")/1,2) a r e both negative for all k, with the remark-
able result that orientation contributions to the free surface energy, in the
limit of hyper-anisotropic drag ( a d - 0), completely stabilize the surface-
tension-driven Rayleigh instability.

If A(sj) < 0 then the growth rate Re(~/2), which reflects the Rayleigh
instability, is positive and bounded for all k, with qualitative behavior
similar to the viscous Newtonian 1-D Model 2 with surface tension and
inertia. Therefore, for such equilibria with sufficiently large Calcp, the
Rayleigh instability survives independent of the drag anisotropy.
This regime suggests that the degree of drag anisotropy plays a critical
role in the orientation-flow coupling, in particular with regard to the possi-
bility that orientation contributions to free surface energy may suppress the
Rayleigh capillary instability. This remarkable suggestion is made in the
mathematical limit, aa = 0, corresponding to zero drag along the dumbbell
axis relative to the drag orthogonal to the molecular axis.

Model 8" LCP fibers with all effects coupled at leading order:

This is the most general regime of LCP fibers, with all effects coupled at
leading order. The linearized dispersion relation for Model 8 is

dU'(sj)k 2
=0. (76)
2W~
In Figure 5 we compare the growth rates for all LCP slender models.

4. FIBER SPINNING STEADY STATE SOLUTIONS AND THEIR STA-


BILITY

In Figure 6 we show the numerical solutions of the viscous Model 2 with


1092

3.5 L ' ' ' .... ' ' /" ' ' ' ' -i

2.5

1.5

0.5

0 ~ N
Model 6 with s m a l l C a
\

-0.5 - \. \ I "- " "-" " -- " ~ Model-7 with- s m a l l C - a -


!
\ '

'/
--1 -

-1.5 , , I .... , I , , I J
0 1 2 3 4 5 6 7 8 9 10
k

Figure 5" Growth rate as a function of wavenumber k for different models.

1 W'
different values of the parameters ~, 1 and T"
1 The upstream boundary
condition is chosen as

v ( O ) - 1, (77)

while the downstream boundary condition is

v ( 1 ) - 10. (78)

Figure 7 shows a typical steady state profile for a Maxwell liquid filament
(Model 3), while Figure 8 gives a typical steady state profile for a Johnson-
Segalman liquid filament (Model 4).
In Figure 9 we present a family of steady state solutions of the liquid
crystalline Model 8 reflecting variations due to changes in the initial degree
of orientation, s(0) = So. Note that there are uniform responses in 0, v, s to
1093

changes in so. Physically, we predict that lower so yields steady processes in


which the filaments are thinner and faster at each interior spinline location.
Figure 10 shows the qualitative dependence of the steady spinning solution
on the draw ratio, which in our nondimensional model coincides with the
take-up velocity. We have considered a range from slow spinning (Dr = 2)
to fast spinning (Dr = 40). From 10 one observes that a higher take-
up speed leads to higher axial velocity, thinner fiber radius, and higher
degree of orientation for all z > 0. In [15] we have also studied how
the steady spinning solution responds to variations in Reynolds number
(R), Weber number (W), and Froude number (F). Our numerical results
indicate negligible quantitative changes in the free surface r axial velocity
v, and uniaxial order parameter s due to variations in W and F, and
still very small quantitative changes due to variations in R. Sensitivity
to these hydrodynamic parameters becomes evident, however, when we
analyze stability of these steady states [15]. For further studies on the
quantitative influence of LCP parameters (e.g., anisotropic drag coefficient,
LCP kinetic energy relative to inertial energy, and dimensionless relaxation
time), refer to [15].
In Figure 11 we present linearized stability results of the steady isother-
mal LCP spinlines of Figures 9 and 10. Figure 11a displays the maximum
growth rate curve for the s0-parametrized family of Figure 9, from which we
conclude that the entire Figure 9 family of steady states is linearly stable.
Figure l l a further indicates that a weaker upstream degree of orientation
leads to more stable steady states.
The maximum linearized growth rates as a function of Dr, for the steady
state family of Figure 10 is plotted in Figure llb, from which we de-
duce that faster processes are less stable in this region of parameter space.
This might seem intuitively obvious, but we refer to a study [5] of slen-
der Maxwell fluid models where the maximum growth rate curve oscillates
about zero with varying Dr, creating alternating windows of stable and
unstable draw speeds.
Figure 11c provides the full two-parameter neutral stability curve: the
critical draw ratio, Dr*(so) for each so, above which the steady state is
linearly unstable. We deduce that the critical draw ratio for all steady
states of Figure 9 is between Dr - 27.5 and Dr - 30.5. Most notable,
however, is the shape of this neutral stability curve, indicating there is
preferred initial degree of orientation (around so = 0.25) associated with
1094

the maximum stable drawing speed (around Dr - 30.4) for this two-
parameter family of steady states.
Now we present steady fiber spinning solutions to Model 9. Consistent
with our nondimensionalization, the upstream conditions on fiber radius,
velocity, and temperature are fixed:
r 1, v ( 0 ) - 1,0(0)- 1. (79)
The downstream boundary condition is selected by the assumption that
axial thermal conduction is negligible downstream, i.e.,
0(r
Oz (1) - 0 . (80)
The remaining boundary conditions are free processing parameters to
be specified/varied in the simulations below"
s(0), v(1). (81)
The upstream degree of orientation (s(0)) is a function of spinneret design,
whereas the take-up speed (v(1) - draw ratio - Dr) is a measure of
process speed and throughput.
In Figure 12 we present typical steady state solutions of Model 9 due
to changes of A0. Here a two-phase model has been posited [26], where
the governing equations for the temperature above the glass transition
temperature (0 > 0g) are given in Model 9, and in the solid phase (0 _ 0g)
we employ a rigid cooling LCP fiber model so the velocity is constant
(fixed by the take-up speed) whereas the orientation and energy equations
are maintained. (The parameters used here are for academic illustrations,
and are not taken from experiments.)
Figure 13 depicts the critical draw ratio Dr* as a function of the non-
isothermal parameter A0. Three different forms of N have been used,
the constant value N - 4 and then two temperature-dependent forms
[26]. These forms are chosen so that the intermolecular potential f U(s)ds
has only the isotropic critical point for sufficiently high temperature, then
passes to a double-well potential for lower temperature.
As shown in Figure 13, the critical draw ratio grows with A0. As A0
increases from 0 to 0.4, there is 9.3% gain in Dr* for N c~ 9.7% gain
for N linear, and 9.5% gain for N quadratic. These predictions clearly indicate
that the cooling process increases stable spinning speeds.
Note that for A0 - 0, N c~ -" 4, N linear : 2.1, and N quadratic - - 2.52,

the qualitative differences in these isothermal steady states are negligible,


1095

1,...
"'111111/I/1[[
. . . .
12

~
10
0.8

0.6
> 6

0.4

0.,~
Ill
..................... II iii|I |1|

1 0.2 0.4 z 0.6 0.8 1 (3.,'} % 0.2 0.4 0.6 0.8

.., ~ 9 o , 12

. ."""/. ////////,//////////,,,,:i~
0.8

0.{

O.z

0.~ 1 ..............
...... 11111111111111|I

1 i ! i I
(b) ~ o'.2 o:4 o:6 o'.8
CO 0.2 0.4 0.6 0.8 1
z z
12

0.!
10
/
/-.:y

0.6 if::
> 6
.../.! i /
0.4
........;:;r i ....

0.2
........... .,.,.,.,~.,~,-r~ . . . . .

Oo 0.2 0.4 0.6 0.8 1


o
-0 0.2 0.4 0.6 0.8
z z

Figure 6: The variation of fiber radius 4) and axial velocity v due to the
changes of: (a) l / F , where 1IF goes from 0 to 5 in increments of 0.5, with
1 / W - 1 / R - 1.0; (b) l / W , where 1/W goes from 0 to 5 in increments
of 0.5, with 1 / F - 1 / R - 1.0; (c) l / R , where 1 / R goes from 1.0 to 5 in
increments of 0.5, with 1/F - 1/W - 1.0.
1096

20

0.8 15
>,,
,,,,,,,,,,

0
~50.6 o10
L
>

0.4

0.2 ! ,,,

0 I

0 0.5 0 0.5 1
axial coordinate axial coordinate

30

25
-0.5
03
20
-I
9-~
X 15
L

-I .5
10

!
-2
0 0.5 1 0 0.5 1
axial coordinate axial coordinate

F i g u r e 7: A t y p i c a l s t e a d y s t a t e profile for a M a x w e l l l i q u i d f i l a m e n t in a
fiber spinning process. The parameter values are B - W - F - Z - a -
1, A1 - 0.05, 0 ( 0 ) - v(0) - 1, T,.,.(0) - 0, D r - 20.
1097

20 |

0.8 15
>.,
to,) ,,,..,.,

rj
~0.6 o10
(1)
>

0.4

0.2 I
0 I

0 0.5 0 0.5 1
axial coordinate axial coordinate

300

r,/) -2
200
L
(l)

,,,,,... ~-4
..,,,
~
"0
100 L

-6

0 ~ ~ -- --8 "' ~ '"

0 0.5 1 0 0.5
axial coordinate axial coordinate

Figure 8" A typical steady state profile for a Johnson-Segalman liquid fila-
ment in a fiber spinning process.
1098

Figure 9: Steady solutions of LCP fiber spinlines reflecting variations due


to changes in the initial degree of orientation, s(0). We vary s(0) = so from
.05 to .95 in increments of .05. Arrows indicate the direction of increasing
so. Figures 9a,b,c display the family of solutions, fiber radius r so), axial
velocity v ( z ; so), orientation s ( z ; so), respectively. All parameter values are
fixed at order one values" 1 / R - c~ - 5, 1 / W - 1, 1 I F - 1 , N - 4, A -
1, a = 0.5. Boundary conditions are: 4)(0) = v(0) = 1, s(0) = .5, D r =
10 = v(1).
1099

Figure 10: Variations in the steady spinning state due to changes in draw
ratio, Dr. All other parameter values and initial data are the same as
Figure 9. D r - v(1) is varied from 2 to 40, in unit increments.
1100

//i
-1.6 /
J I
1. /

// //
/ i
/
-1.7 /
t
/ /
/ /
/ /
/
/
-1.8 /
/ /
j/
l/max l/max /
//
/
/
-1.9 /
/J I
/
i
//
I
i
// i
l
/ I
j/
/ (b) l
l

-2.1 ...................... 9 , , , ~ . . . . : . . . . : ~ .

0.2 0.4 0.6 0.8 i0 2O 3O 40


.
80 D?

/ - ....... \
/' \,
3o / \
:
/ \ \

/ \
29.5 ',
'\
Dr*
29 \

28.5

(c)
0 0.2 0.4 0.6 0.8
80

Figure 11: Linearized stability information. Figure l l a concerns the s0-


parametrized family of Figure 9; Figure 11b concerns the Dr-parametrized
family of Figure 10; Figure 11c summarizes the neutral stability bound-
ary for a full two-parameter variation of the spinning steady state. Each
data point in Figures l la,b depicts the maximum real part, Uma~ (vs. so
in Figure 11a, vs. D r in Figure l lb), of all linearized growth rates for
that respective steady state; a negative value of U,ia~ indicates linearized
stability.
1101

1
10 : ""--:----:; ....-.....:...... -.... ;
0.9

0.8
B J/#I/Y ....... ................

0.7

0.6
:i!~i i
0.5

0.4
~) .... , ~ :i!;i:i:!:~:~:~''
0-30 0:2 ....0:4 0:6 0:8 ~ 0'2 o14 0'6 0'8 1
Z z
(b)

0.8 ..,~iii~!~

0.6 ~ii~ o:i~~~,,,~ ~ii!i!:~il!i,i:i~:~ .,


0.4

0.2

O0 012 014 016 018 1


0.65

06;
f 0:2 0:4 0:6 0'8
. . . 11
z z
(d)

Figure 12" Steady state variations with respect to the nonisothermal pa-
rameter A0, 0 <_ A0 <_ 0.9, in increments of .045. All parameter values
are fixed at order one values" c~ - 5, R - 0.2, 1 / W - 1 / F - A - 1,
N- 4, a d - - 0 . 5 , P e - ~ c - c v - 1, 0 g - 0 . 8 . Boundary conditions are:
r v(0)- 0(0)- 1, s(0)- 0.5, v(1)- 10.
1102

43 9 9 n i

42

41

~
I,,--

D
.1~i" 5~
40 J ...

~ .. : 1~:'~
~... ""

39 ...

...
I ...
.~
......
.....

8 , i l I I

0 0.1 0.2 0.3 0.4


Ae

Figure 13" Critical draw ratio as a function of the nonisothermal parameter


A0: effects of a temperature-dependent excluded-volume potential. The
solid curve corresponds to N c~ = 4.0, the dotted curve corresponds
to N linear = - 3 -t- 5 . 1 / 0 , and the dashed curve corresponds to N quadratic =
25 - 43.9/0 + 21.42/02. Here the material parameters are the same as in
Figure 12. The A on the left margin of the solid curve corresponds to the
critical draw ratio, Dr*, for the isothermal steady state of Figure 12.
1103

whereas the effect on process stability is on the order of 2%. Therefore, the
dimensionless LCP density parameter N has a bearing even in isothermal
stability of spinning states.

5. A P P L I C A T I O N S OF S L E N D E R F I L A M E N T M O D E L S

We now discuss several applications of slender filament models, beyond


the modeling of steady fiber manufacturing processes [1, 2, 3] and drop
formation [7, 8, 9].

5.1 O n s e t of p o l y m e r i c filament failure


In this section we use slender, one-dimensional asymptotic balance equa-
tions to model stretching and the onset of failure in a polymeric liquid
filament. In one experiment designed by J. Matta to measure elongational
flow properties of dilute polymer solutions, the fluid is slowly extruded
from a vertical nozzle and forms a drop at the nozzle tip. As the drop
reaches a critical mass it begins to fall, pulling a filament behind it. This
experiment is described in reference [31]. Markowich and Renardy [32]
modeled the Matta filament/drop experiment with simulations of a slen-
der 1-D quasilinear parabolic integro-differential equation, derived from
the 3-D Jeffreys (or Johnson-Segalman) constitutive law, which combines
elastic relaxation and retardation. Their numerical simulations predict a
qualitatively similar filament behavior for all parameter values reported.
That is, an initial uniform relaxed filament always fails at the ceiling, as
indicated in their model by an excessive localized growth in axial strain at
the ceiling. Further, they report that enhanced elasticity delays the time
of failure, while increased surface tension advances the failure time.
The model we analyze here derives from a Maxwell viscoelastic con-
stitutive law, which isolates the effect of elastic relaxation from elastic
retardation. The Eulerian forms of both the Maxwell and Jeffreys 1-D
slender axisymmetric approximations have been given in previous sections.
The key distinction between the 1-D Jeffreys model and the 1-D Maxwell
model is that the Maxwell equations may be either genuinely hyperbolic or
mixed elliptic-hyperbolic, whereas the retardation terms from the Jeffreys
constitutive assumption yield a quasilinear parabolic system. As a result
the Maxwell model predicts a wider variety of filament behavior than the
Jeffreys model, both in the location and nature of approach to failure, for
1104

the same class of initial/boundary data.


The onset of filament failure is indicated by the 1-D Maxwell filament
model through a hyperbolic to mixed elliptic-hyperbolic change-of-type.
Our rationale for this interpretation is that a filament break must be ac-
companied by arbitrarily small axial length scales to resolve the rupture;
coincidentally, the change-of-type in our model corresponds to a localized
rapid generation of small lengthscales. Since the radius of the filament
does not simultaneously vanish at the change-of-type location, the aspect
ratio of radial to axial lengthscales is no longer small, in violation of a fun-
damental assumption of the I-D model. Thus, although our model breaks
down at the change-of-type location, the way in which the model becomes
invalid coincides with behavior associated with filament failure. Depending
on the particular process conditions and material properties, the change-
of-type location may occur anywhere from the ceiling to the drop and does
not necessarily occur at the location of minimum radius.
The diagnostic for change-of-type in the mathematical model is the dis-
criminant defined in equation (23) which is positive for well-posed evolution
and which vanishes at the change-of-type space-time location (z*, t*). This
discriminant function consists of a sum of three terms, each with clear
physical importance. There is an inertia/relaxation term which is strictly
positive, a stress term which is positive in tension and negative in com-
pression (the sum of these first two terms must be nonnegative [33)], and a
surface tension term which is strictly negative. The sign of the discriminant
may change as the initial/boundary data evolve, and the evolution toward
change-of-type involves a delicate balance among these competing physical
expressions. But there is no simple, consistent pattern in the transition
away from a slender flow. Rather, a variety of mechanisms is possible and
the model predicts various approaches to failure. This is in contrast to
the results of Markowich and Renardy [32] for the Jeffreys model whereby
the filament always fails at the ceiling due to an unbounded axial strain
accompanied by a vanishing radius.
In this study, we convert the Eulerian I-D Maxwell filament model to
Lagrangian form and adapt it to model the filament break experiment of
Matta. We always pose uniform initial data such that our I-D Maxwell
model is a well-posed genuinely hyperbolic system for finite nonzero time,
and satisfies a slenderness condition for finite nonzero time. Thus, our
model allows the study of the competition of the included physical effects
1105

0 243.29 486.57

o
O~-'J

.,..4
M

o
,=;
~, .0 243.29 ' 486.57
,v--.q 0
'It' --.4

9,,.4 o

.,=4
r

.,,.~
"0

o .....

o ~.29 ~88.~7
axial c o o r d i n a t e

Figure 14: Onset of failure preceded by uniform thinning and uniform


change-of-type: radius ~b and axial force r - T,~) - ~ / W of the fila-
ment, and discriminant A at the time of failure t* = 9.8470. The param-
eter values are B - A - a - 1, W 1 - 1 , ~1 _ 10, Z - 1, vol- 100 . The
number of mesh points n - 800 and the time step A t - 0.625 10 -4.
Hyperbolic to elliptic-hyperbolic change-of-type occurs at the drop inter-
face z* - L ( t * ) - 486 957, with -~-z(
d~ z*-, t* ) -- --2.5199 10 -6 very small, so
that the change-of-type is uniform. At the moment of change-of-type the
centroid of the filament is located at Zero(t*) = 358.68.
1106

0 0.5002 1.0004

0 0.5002 1.0004
o

0 0.5002 1.0004
axial coordinate

Figure 15: A very brittle onset of failure: radius, axial force, and discrimi-
nant at the time of model failure t* - 0.951 x 10 -2 9 B - A - a - 1, W 1
=
114, T1 _ 20, Z - 20, v o l - 10/~; the number of mesh points n - 3200 and
the time step At -- 0.5 x 10 -5. Change-of-type occurs at z* - 0.78181 x 10 -3
near the ceiling; note T d A( z *-~ t* ) - - 9 3 . 2 9 8 and ~ ( z *+ t*) - 1070.1 are
large and of opposite sign, indicating a cusp-like behavior in A near (z*, t*).
At the m o m e n t of change-of-type the centroid of the filament is located at
Zcm(t*) = 0.50100. The suggested failure is brittle, with only 0.04 percent
strain at the m o m e n t of change-of-type.
1107

during the evolution from a slender uniform filament until the slenderness
assumption breaks down. In particular, the model fails whenever there is
a hyperbolic to elliptic transition, since at that time the model experiences
exponential growth in all lengthscales. The filament continues to stretch
up to and even after the change-of-type, and we conjecture this transition
equates to the onset of filament failure. Thus our model captures the
evolution into non-slender flow.
As long as the slenderness approximation is maintained and our assump-
tion as to the competing physical effects remains valid, we can confidently
predict details of the onset of failure: whether failure is approached through
uniform flow and profile or through a sharp, localized gradient; where the
onset of small scales occurs; whether the onset of failure is ductile (i.e.
the filament at onset of failure has stretched significantly from the orig-
inal length) or brittle (i.e. there is essentially no elongation before onset
of failure); and whether there is little mass redistribution or concentration
of mass. These are the aspects of polymeric filament behavior observed
in the model. Figure 14 depicts a ductile failure with significant mass
redistribution; whereas Figure 15 shows a brittle failure with little mass
redistribution.
Of importance to the application of this study, after the critical time of
ill-posed change-@ type our model is invalid. That is, if we apply the model
at all lengthscales, the full pdes are ill-posed. Effectively, the characteristic
axial scale goes to zero immediately after the elliptic change-of-type with
no argument for the radial scale converging to zero, so the slenderness con-
straint is immediately violated. As a result, we make no claims to capture
the detailed structure of breaks.

5.2 N e w t o n i a n fluid characterization for dynamic surface tension


To address overspray of agricultural pesticides (due to spray drift and
drop rebound from leaf surfaces) leading to groundwater contamination,
the U.S. Department of Agriculture (USDA) needs to measure the physical
properties of the spray solution which control droplet formation and drop
rebound, namely surface tension and elongational viscosity; regulations
are based on these properties. Due to the presence of surfactant addi-
tives, significant decay in surface tension of the pesticide solution, from
an initial value near that of pure water (72 dyne cm -1) to as low as 30
dyne cm -1, occurs in the first 2-5 ms from the creation of the free sur-
1108

face, which coincides with the timescale of drop formation, atomization,


and drop rebound. Since surface tension dominates the hydrodynamic in-
stability leading to droplets and drop rebound, accurate methods for the
determination of surface tension on the millisecond timescale are neces-
sary; slow or static techniques such as the Du-Nouy ring effectively yield
the equilibrium surface tension, which may be as low as 20 dyne cm -1.
Similarly, the presence of the surfactant also introduces a flow dependence
in the elongational viscosity which must be understood. With USDA re-
searchers in Wooster, Ohio, we, in work described in [16], combine an
inverse formulation of their model for an oscillating free surface jet with
elliptical cross-section [12, 39] with USDA experiments of oscillating jets
to measure the dynamic surface tension and elongational viscosity of jets
of surfactant solutions with ages on the order of 1 ms.
The oscillating jet technique is the only technique which measures the
surface tension of a very new surface on the time scale of milliseconds. The
oscillating jet is a liquid/gas system in which new free surface is created
constantly, as in spray formation and drop rebound. The surface in the
oscillating jet over the axial domain where the measurements are made
has existed much less than the time necessary for the equilibrium surface
tension value to be reached. To briefly describe the oscillating jet phe-
nomenon, when a fluid exits an elliptical orifice the resulting free jet has
a chain-like appearance, with the jet cross section oscillating in an elonga-
tional flow between perpendicular ellipses as it goes down the jet axis (see
Figure 16). Essentially, the jet is an effective mass/spring oscillator; the
tension of the free surface (the "spring") pulls the elliptical cross section
to a circle to minimize area, but the inertia of fluid motion in the cross
section (the "mass") causes an overshoot in the perpendicular direction,
and so on. The wavelength of the oscillation is related thereby to the
average surface tension over that length, and the amplitude decay to the
average fluid viscosity. The particular contour of the free surface between
the maximum amplitudes is indicative of the evolution of surface tension
and viscosity within the wavelength, and can be exploited in the inverse
problem. In the oscillating jet, the elongational motion is transverse to the
filament axis, with high strain rates (on the order of 103 [16)], but small
strains. This is in contrast to the flow under tension of [18] exploited in
the next section, in which the elongational flow is in the direction of the
filament axis, with much larger strains at lower strain rates.
1109

The USDA apparatus utilized in [16] records on video tape simultaneous


perpendicular views of the oscillating jet free-surface profile, and hence af-
fords the seven measurements shown in Figure 17. These measurements, in
combination with an analytical model for the oscillating jet which assumes
the surface tension and viscosity are constant over the wavelength over
which the measurements were taken, produce in [16] the average values of
surface tension and viscosity over the corresponding surface age interval.
In [36], free-surface measurements made with the USDA apparatus and the
model of [16] are utilized to characterize the surface tension of a battery
of fluids.

Figure 16" The oscillating jet phenomenon; from [40].

In [17] equations are derived for oscillating jets which allow for surface
tension and elongational viscosity to vary in space and time. Three specific
relations for the decay of surface tension with surface age are investigated"
exponential decay, the thin-film diffusion model derived by [35], and the al-
gebraic form proposed by [41]. For each form the direct problem is solved,
i.e. the governing integro-differential equations are numerically solved with
the particular specified function for the decay of surface tension with age
to obtain the free surface profile. In all forms initial and equilibrium values
of surface tension and decay rates were selected consistent with experimen-
tal values reported in [35] and [41]. See Figure 18. Important qualitative
differences in the oscillating jet behavior predicted under the three forms
are discovered, with a striking similarity: in all forms the nearfield decay
of surface tension, occurring in the first few wavelengths of oscillation or
over a fraction of the first wavelength determines the amplitude and wave-
lll0

length of the farfield oscillation. This has powerful consequences in the


inverse formulation of the problem, since rapid surface tension decay can
be deduced from measurements of the jet profile far downstream of where
this decay takes place.
B
R rain

B
-iV- R max

M
R mtt(

~, Ru
rain

E
R min

E
_~t_ R mcu

Figure 17: USDA measurements taken of the oscillating jet" from [16].

The inverse formulation of the analysis of [17] provides the means to


resolve dynamic surface tension and elongational viscosity within a wave-
length of oscillation and on millisecond timescales.

5.3 N o n - N e w t o n i a n fluid c h a r a c t e r i z a t i o n for elongational viscosity and relax-


ation
In [18] we combine recent advances in spinline measurements with math-
ematical analysis to produce a new capability in material characterization
in elongational processes. Three experimental ~dvances documented for
the first time in [18] are (i) the ability to measure the kinematics (filament
radius and slope, velocity, and velocity gradient) pointwise along the fila-
ment accurately enough that one can treat them as known, differentiable
functions in the inverse problem for material properties, (ii) quantitative
1111

75
\

\~
,,,
60 \~

~ 55
E
\x 9
gso \X
x
g '\\ .
9~ 45 \
\
\
.
.
\ \\
4o \

.~ 35
..... Z.~_~.~ ....

25 i i i i i i i
0 0,005 0.01 0,015 0,02 0.025 0,03 0,035 0,04
surface age (s)

1,8

1,6

1,4

_1,2 i t
A /I,, I. /,' I, /,, /~,
/?',i,,
/,,/,', /,~ ,, t
-e-

. 1
\ I I I,' I', /'," t',,,Ii,'x,, I /,'t ,,,/
~u}0 , 8 \ I t I; ~',l',
\
\,,/.','
'\ ~
\,',/,,'
\
\,I, ~. '

.~_ 0,6

9~" 0 , 4

0.2

01 i i i i /

0 2 4 6 8 10 12
dimensionless axial distance z

Figure 18: (a). Surface tension vs. surface age. For all three decaying forms
- 72.8 dyne cm -1 at 0 ms and cr - 31.4 dyne cm -1 at 0.031ms. The
constant case is for a - 72.8 dyne cm -1 ( ); exponential decay ( - -
- - ) ; Brazee model ( . . . . . . ); Hua model ( . . . ) . (b) Free surface profiles
~1 with the variable surface tension given (a) and ~ ( 0 ) - 1.5, 01,z(0) - 0 .
1112

measurement of the axial force in the filament at the die exit, and (iii) an
accurate steady flow-rate control. In the analysis we admit many possible
physical effects. In the forward problem the known material properties and
boundary conditions dictate the dominant physical effects. In the inverse
problem considered in [18], the determination of which physical effects are
dominant is part of the solution.
The apparatus employed in [18], shown in Figure 19, was constructed
in collaboration with Prof. K.W. Koelling of the Department of Chemical
Engineering at the Ohio State. An infilsion/withdrawal syringe pump is
used to pump the fluid through a stainless steel delivery tube, controlled
to a specified value which ranges from 1.5 x 10 -s to 2 x 10 -1 ml/s. This
system avoids the variability in the volumetric flow rate of other types of
pumps, due to for instance the difficulty in controlling the pressure level in a
reservoir with falling fluid level, or the periodicity of a reciprocating pump.
The fluid is ejected from the delivery tube in the form of a vertical filament
which is taken up tangentially by a rotating drum of 5 cm diameter. The
take-up drum is mounted on a movable assembly which allows the fiber
length to be varied fi'om 0-19.6 cm. The drum is also connected to a flexible
torque transmission cable which allows the driving motor to be placed away
from the apparatus in order to minimize vibrations. A scraper is attached
to the bottom of the drum to remove excess fluid.

Figure 19" The experimental apparatus.

The diameter of the filament as a function of axial position is recorded


1113

using a charged coupled device video camera positioned on a constant speed


vertically traveling stage. A limitation to resolution in edge detection, es-
pecially for transparent fluids, is the variability in apparent edge location
due to non-uniform illumination; the advance here in the resolution of the
free surface is due in large part to the use of a diffuse background illumina-
tion system (Lumitex Inc.). Images are recorded using a Super VHS video
cassette recorder. Two other techniques to measure the fiber diameter were
also evaluated: still photography, and using a laser based diameter mea-
suring device; the video setup was found to provide comparable resolution
coupled with greater ease in recording measurements. The video camera
captures an image every 1/60 of a second. Each digitized frame captures
448 pixels in the axial direction; to avoid possible aberration in the camera
lens only the central 400 pixels of the frame are utilized. The digitized
image files are processed using an Optimal Zero Crossing edge detection
algorithm. The velocity of the camera translation is computed by tracking
an identifiable fixed point in successive images. Enough frames are stored
so that each axial location on the free surface appears in the core 400 pixels
in three different frames. Using the position of each successive image and
the velocity of travel of the camera, the images are concatenated to give
the entire profile of the jet. The raw data in terms of pixels is converted
to length units with the calibration factor 0.001254 cm/pixel. The error
in image capture and edge detection procedure is no more than one pixel,
or 12.54 #m. This resolution enables the apparatus to measure on-line
the kinematics of the deforming filament as a function of axial position
accurately enough so that they and their derivatives can be considered as
known functions in the inverse problem for material properties.
The delivery tube deflects downward due to the tensile force in the
filament. This deflection, measured by a displacement transducer, is used,
together with analysis, to deduce spinline force at the tube exit (a filament
under take-up conditions is referred to as a spinline in the terminology of
the textile and reinforcing fiber manufacturing industry). There are several
possibilities for the zero force state in the calibration of the transducer, such
as the delivery tube full of stationary fluid, and the state with a container
placed just below the issuing jet. These two choices have drawbacks" the
first does not take into account dynamic and pressure effects in the tube
during filament formation, and the second fails due to the elasticity of the
fluid studied in [18], which leads to an upward thrust exerted on the tube.
lll4

To avoid these errors, two sets of surface profiles and tube deflections are
recorded for each experiment, one for the filament with the take-up force,
and one, for the calibration, without windup falling only under the effect
of gravity. Analysis contained in [18] allows one to deduce the axial force
in the filament with windup at the tube exit from comparison of the two
sets of measurements.
Reference [18] demonstrates how the measurements provided by the ap-
paratus can be exploited to give quantitative information about rheological
material properties in isothermal elongational flows under applied tension.
The profile and force measurements fl'om a particular experiment are cou-
pled with a mathematical model formulated as an inverse problem, to de-
duce material properties. Referring to [18] for details, with the fiber profile
and upstream axial force experimentally known, the momentum and con-
stitutive equations decouple: the momentum balances reveal the evolution
of stresses that must be present in the jet to balance the effects of inertia,
gravity, and surface tension and produce the experimentally observed mea-
surements; and the constitutive equations deduce what stresses would be
produced by the measured kinematics and upstream boundary conditions
for a proposed rheological model of the fluid. The inverse problem consists
of searching through parameter space in the proposed rheological model
until the stresses generated by that model match those that must be there
to satisfy momentum considerations. Specifically, in [18] the elongational
rheology of a Boger test fluid is characterized within the assumption of a
single-relaxation-mode Giesekus constitutive model. Significantly different
values of the relaxation time and mobility parameter were obtained in the
elongational characterization than were obtained via shear rheometry: the
shear rheometer produced a relaxation time of 0.9148 s and a mobility pa-
rameter of 0.00035, whereas in the elongational flow the relaxation time
was found to be 1.323 s with one inversion technique and 1.303 s with
another, and the mobility factor was found to be zero, so that in the elon-
gational flow (but not shear flow) the proposed Giesekus model collapsed
to an Oldroyd-B fluid. The solvent and polymer viscosities deduced from
the elongational and shear experiments were the same.

ACKNOWLEDGEMENT

This work was funded in part by the Air Force Office of Scientific Re-
1115

search, Air Force Materials Command, USAF, under Grants F49620-97-1-


0001 and F49620-97-1-0003, and the National Science Foundation, under
Grants CTS-9319128 and CTS-9711109. The US Government is authorized
to reproduce and distribute reprints for governmental purposes notwith-
standing any copyright notation thereon. The views and conclusions con-
tained herein are those of the authors and should not be interpreted as
necessarily representing the official policies or endorsements, either ex-
pressed or implied, of the Air Force Office of Scientific Research or the
US Government.

REFERENCES

o M. Matovich and J. R. Pearson, Ind. Eng. Chem. Fundam. 8 (1968)


512.
. M.M. Denn, C.J.S. Petrie, and P. Avenas, AIChE. J. 21 (1975) 791.
3. W.W. Schultz and S.H. Davis, J. Rheology 26 (1982) 331.
4. L. Ting and J.B. Keller, SIAM J. Appl. Math. 50 (1990) 1533.
5. M.G. Forest and Q. Wang, Siam J. Appl. Math. 54 (4) (1994) 996.
6. S.E. Bechtel, M.G. Forest, and K.J. Lin, SAACM 2 (1992) 59.
7. J. Eggers and T. Dupont, J. Fluid Mech. 262 (1994) 205.
8. J. Eggers, Phys. Rev. Lett. 71 (1993)3458.
9. X.D. Shi, M.P. Brenner, and S.R. Nagel, Science 265 (1994) 219.
10. D.T. Papageorgiou, ICASE Report 93-45 (1993).
11. S.E. Bechtel, J.Z. Cao: and M. G. Forest, J. of Non-Newtonian Fluid
Mech. 41 (1992) 201.
12. S.E. Bechtel, M.G. Forest, D.D. Holm, and K.J. Lin, J. Fluid Mech.
196 ( 988) 241.
13. S.E. Bechtel, K.D. Bolinger, J.Z. Cao, and M.G. Forest, Siam J. Appl.
Math. 55 (1) (1995) 58.
14. M.6. Forest, Q. Wang, and S.E. Bechtel, Physica D 99 (4) (1997) 527.
15. M.G. Forest, Q. Wang, and S.E. Bechtel, J. Rheology 41 (4) (1997)
821.
16. S.E. Bechtel, J.A. Cooper, M.G. Forest, N.A. Petersson, D.L. Re-
ichard, A. Saleh, and V. Venkataramanan, J. Fluid Mech. 293 (1995)
379.
17. S.E. Bechtel, M.G. Forest, N.T. Youssef, and H. Zhou, accepted for
publication in J. Appl. Mech.
18. V.V. Ramanan, S.E. Bechtel, V. Gauri, K.W. Koelling, and M.G.
Forest, J. Rheolo~v 41 (1997~ 283.
1116

19. D.D Joseph, Fluid Dynamics of Viscoelastic Liquids, Springer-Verlag,


New York (1990).
20. A.N. Beris and B. Liu, J. Non-Newtonian Fluid Mech. 26 (1988) 341.
21. B. Liu and A.N. Beris, J. Non-Newtonian Fluid Mech. 26 (1988) 363.
22. M.G. Forest and Q. Wang, J. Theor. Comp. Fluid Dyn. 2 (1990) 1.
23. Q. Wang, M.G. Forest, and S.E. Bechtel, J. Non-Newtonian Fluid
Mech. 58 (1994) 97.
24. A.V. Bhave, R.K. Menon, R.C. Armstrong, and R.A. Brown, J. Rhe-
ology 37 (1993) 413.
25. M. Doi, J. Polym. Sci. Polym. Phys. Ed. 19 (1981) 229.
26. M.G. Forest, H. Zhou, and Q. Wang, Univ. of North Carolina at
Chapel Hill, Dept. of Math., Preprint Series # 97-11.
27. L. Rayleigh, Proc. Lond. Math. Soc. 10 (1879)4.
28. F.J. Garcia and A. Castellanos, Phys. Fluids 6 (1994) 2676.
29. S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, (Claren-
don Press, Oxford, 1961).
30. S.E. Bechtel, C.D. Carlson, and M.G. Forest, Phys. Fluids 7 (1995)
2956.
31. J. Matta and R.P. Tytus, J. Appl. Polym. Sci. 27 (1982) 397.
32. P. Markowich and M. Renardy, J. Non-Newtonian Fluid Mech. 17
(1985) 13.
33. D.D. Joseph and J.C. Saut, J. Theor. Comput. Fluid Dynamics 1
(1990) 191.
34. W.D.E. Thomas and L. Potter, J. Fluid Mech. 76 (1975) 625.
35. R.D. Brazee, M.J. Bukovac~ J.A. Cooper, H. Zhu, D.L. Reichard, and
R.D. Fox, Transactions of the ASAE 37 (1994) 51.
36. D.L. Reichard, J.A. Cooper, S.E. Bechtel~ and R.D. Fox, Atomization
and Sprays 7 (1997) 219..
37. R.S. Hansen, M.E. Wallace, R.C. Woody, J. Phys. Chem. 62 (1958)
210.
38. R. Defay and J.R. Hommelen, J. Coll. Sci. 13 (1958) 553.
39. S.E. Bechtel, J. Applied Mech. 56 (1989) 968.
40. L. Rayleigh, Proc. R. Soc. Lond. 41 (1890) 281.
41. X.Y. Hua and M.J. Rosen, J. Colloid and Interface Science 141 (1991)
180.
1117

NUMERICAL SIMULATION OF MELT SPINNING OF


POLYETHYLENE TEREPHTHALATE FIBERS

Kyoung Woo Kim a, Sang Yong Kim, and Youngdon Kwon b

Department of Fiber and Polymer Science,


Seoul National University, Seoul 151-742, Korea

aFiber Research Center,


Sunkyong Industries, Su Won, 440-745, Korea

bDepartment of Textile Engineering,


Sung Kyun Kwan University, Su Won, 440-746, Korea

1. INTRODUCTION

Fiber spinning is one of the major polymer processing operations, by


which fine and uniform filaments with required modulus and strength can
be produced. There are three types of spinning processes such as melt,
wet and dry spinning. Of all these, the melt spinning is the most widely
applied for the production of synthetic fibers, where severe elongational
deformation is usually accompanied. Thermally stable polymers such as
polyesters, polyamides and polyolefins are employed for manufacturing
fibers via melt spinning. Especially, the melt spinning of polyethylene
terephthalate(PET) has been intensively investigated due to its commercial
importance.
In the typical melt spinning operations, undrawn filament yarns (UDY)
are made at the spinning speed of 1---2 km/min. Then, the process of
drawing in the solid state by 3---5 times of the original length follows in
order to enhance the mechanical properties, and thus, results in fully
drawn yarns (FDY). Herein, the melt spinning process is performed in
1118

two steps; one is spinning and the other is drawing. Later, for the
improvement of productivity, spin-draw process was developed, which
yields so called partially-oriented yarns (POY) at the spinning speed of
3 --- 3.5 km/min. They constitutes draw-textured yarns (DTY).
If one further increases the spinning speed, the drawing process hardly
affects mechanical properties of the fibers, and then the melt spinning can
be carried out in one step to get the final products. This high speed
spinning process was developed in the 1970's by modifying the
functionality of winders. In the 1980's, high speed spinning of PET, PP
and Nylon has been thoroughly investigated, and commercial products
manufactured at the speed of 6---8 km/min[1] become available. Under
such high speed spinning, one may occasionally observe some instability
like melt fracture and draw resonance, which deteriorates the filament
quality and reduces its productivity[2-3].
One can also notice neck-like deformation as the spinning speed
increases, and vast amount of research has been carried out to clarify the
origin of such phenomenon[4-5].
One of other peculiarities in high speed spinning is the crystallization
in the spinline, where according to the order of their appearance three
regions such as flow deformation, orientation crystallization and plastic
deformation can be defined[6].
Flow deformation occurs immediately after the exit of the spinneret,
and molecular orientation formed in the spinneret channel becomes
relaxed. Below this is the orientation crystallization region, where abrupt
increase in the orientation of amorphous region induces crystallization and
the filaments become thinner and necking may occur. Finally, in the
plastic deformation region, filaments are frozen and cold drawing under
the additional effect of air drag due to the increase in the filament speed
is also possible.

1.1. Structure and Properties of High Speed Spun Filaments


1.1.1. Morphology anal Physical Properties of Filaments
As was explained in the previous section, the main objective of the
high speed spinning is to produce fibers with the required high molecular
orientation and crystallinity in one step process, that is, the spinning
without the drawing.
1119

In the case of PET fibers spun at the speed of 3---4 kin/rain, even
though molecular orientation is high, the fibers are in low crystalline or
amorphous state[7,8]. Tensile strength and elasticity of the fibers are also
very low, but the breaking elongation is high, thus the fibers cannot be
applied for practical use without some additional processing like drawing
followed by heat treatment. However, if the spinning speed reaches 5---6
km/min, such produced filaments possess as high as 30---40% of
crystallinity and reasonable mechanical characteristics.
In general, with the increase of spinning speed, higher the tensile
strength of the filaments becomes and lower the breaking strain. When
the speed exceeds 5 km/min, necking in the cold-drawing process after
spinning disappears and the first yield point in the stress-strain curve
becomes obscure. Hence, the stress-strain curve becomes similar to that of
FDY. The tensile strength achieves maximum at the spinning speed of 6
---7 km/min and then decreases if it is further raised[9].
The high speed spun fibers have higher crystallinity and crystal
orientation than the low speed spun fibers with the treatment of cold or
hot drawing. They possess low total orientation due to lower molecular
orientation in the non-crystalline region and thus exhibit poorer
mechanical properties than the low speed spun and drawn fibers.
Thermal stability of high speed spun fibers is very high, and the
shrinkage in the boiling water is usually very low. Dye uptake of high
speed spun fibers is better than that of low speed spun and drawn
fibers[ 10].

1.1.2. Structure Distribution in the Fiber Cross-section


The cross-section of fibers spun at low speed and drawn are relatively
uniform in its morphological structure, but sometimes the birefringence in
the core region is somewhat higher than that in the sheath region[l 1].
This phenomenon may result from the thermal effect on the surface
during drawing. On the other hand, the birefringence in the skin region
of fibers spun at high speed is higher than that in the core region, and
thus, the high speed spun fibers have a distinct skin-core structure.
When observing the cross-section of high speed spun fibers under the
interference microscope[12], one may examine that the distribution of
birefringence along the fiber cross-section is uniform for the fibers spun
1120

at the speed below 6 km/min, however, the birefringence is higher in the


skin than in the core for the fibers spun at the speed above 7 km/min.
From the dyeing test, it has been also found that the skin region shows a
darker image than the core region[13].
From these investigations, one can conclude that the cross-section of
high speed spun fibers are not homogeneous, but exhibits the distinct
skin-core structure.

1.2. Structure Formation in High Speed Spinning


1.2.1. Attenuation during Spinning
The shape of molten polymers extruded from the spinneret nozzle
attenuates under the influence of take-up tension. The attenuation is very
slow in the low speed spinning, but in the case of high speed, the
diameter of fiber decreases rapidly until it reaches the point of 50---100
cm below the spinneret[4]. This behavior is similar to plastic deformation
of metal or neck formation during the cold drawing of solid polymers.
Therefore, it is often called neck-like deformation or necking. The
appearance of necking in the spinline is reported in the high speed
spinning of PET, PA and polyolefin.
Necking seems to result from the high tensile stress concentrated in the
narrow region of spinline. This is one of the unique characteristics of
high speed spinning which may play an important role in determining the
regime of internal structure formation.
Study on the condition of the neck formation in the solid material was
initiated via the application of the Consid~re construction[14,15], and a
number of numerical studies[16,17] followed. The theory states that the
position of neck formation is usually located at the point where the
following equation based on the Consid~re construction from the true
stress-strain curve holds"
dGt 6t dGt r (1)
d/l - /1 or de - l+e
where at is the true stress, /1 the draw ratio, and e the strain.
If the condition
dat at
dA)A
is valid in the entire region, smooth attenuation should be expected;
1121

however, if equation (1) is satisfied at one point, fracture may occur after
necking. On the other hand, if it holds at two points, the spinline
stabilizes and cold drawing will follow[18].
Necking phenomena at high speed spinning have been investigated since
Perez's study[19]. It has been reported that crystallization starts near the
neck point, which was confirmed through examining the data of X-ray
scattering for the PET filaments.
Necking has been explained in terms of the spinning instability and
draw resonance[20]. On the other hand, it has been also studied in view
of the structure formation, and it is claimed that necking occurs at the
point where the viscosity drops abruptly. Thus, necking point moves
upward along the spinline as the take-up speed increases[21].
Haberkorn et a1.[22] showed that crystallization rate is a function of
take-up speed for Nylon 6 and 66 in the high speed spinning. According
to their explanation, neck-like deformation occurs in the region above the
solidification point of the fiber, and then the polymer crystallizes after
neck formation. Thus, it is asserted that crystallization does not induce
neck-like deformation and the decrease of local viscosity due to the heat
of crystallization cannot be the cause of the necking.
A theoretical study of the neck-like deformation has been conducted
after the hypothetical division of the neck deformation into qualitative and
quantitative deformations.
Ziabicki[23] explained that an inflection point on the curve of velocity
variation along the spinline becomes a criterion for neck formation and
the cause of necking is the viscosity increase due to cooling of the melt.
However, the viscosity increase is only a necessary condition for the
stable spinning but not a sufficient condition for necking.
Even the existence of inflection point in the velocity distribution curve
may not be sufficient for the occurrence of necking, thus it can be said
that necking occurs if the radius variation exhibits two inflection
points[24]. The condition of neck formation is, thus, the appearance of
two inflection points, where
R"-- d2R
dz2 -- 0 (2).
Here R is the filament radius and z the axial coordinate.
1122

1.2.2. Orientation Crystallization Behavior


Since the fibers crystallize within 5---10 ms under high tensile stress in
high speed spinning, they show structure and properties quite different
from those of low speed spun and drawn fibers.
Orientation of polymer molecules starts to develop in the amorphous
region under high take-up forces, and then this oriented region
crystallizes. Thus, crystalline region in the fibers possesses high degree of
orientation, and amorphous region remains in the state of low orientation.
The crystals are in the form of bundle-like fibrils rather than chain-folded
lamellae[25-27], and the fibrils grow with the increase in take-up speed.
Heuvel and Huisman[28] demonstrated the dependence of molecular
orientation on the take-up speed. The results are as follows: below the
speed of 2 kin/rain, the low oriented molecules in the fibers do not
crystallize, at 3.5 km/min the molecules are oriented but do not
crystallize, and above 5 km/min the distance between molecular chains
becomes close enough to form crystalline state.
Necking also appears in the spinline under the condition whereby the
molecules are able to crystallize. In other words, necking is not probable
for the noncrystallizable polymers such as atactic polystyrene or even for
crystallizable polymers when the crystallization is suppressed by rapid
cooling or low tensile stress.
Effect of crystallization on necking has been discussed in view of heat
transfer and deformation mechanism for some crystallizable Newtonian
materials[29]. Crystallization is responsible for the direct increase of
viscosity, and hence, increases necking intensity. On the other hand,
crystallization heat lowers the local viscosity indirectly, and decreases
necking intensity[23].

1.3. Numerical Simulation


An attempt to numerical simulation of the melt spinning process was
first made by Ziabicki [3 0] and Kase et a1.[31,32], and they were followed
by Prastaro[33], Yasuda[34,35], and Denn[36]. In their approaches, the
analysis neglected the effect of non-isothermal crystallization and
molecular orientation. Therefore, even though the results describe quite
satisfactorily the experimental data in the case of low speed spinning
process without crystallization, the numerical scheme in high speed
1123

spinning with high degree of crystallization necessitates some modification.


Shimizu et al.[4] and Katayama and Yoon[37] constructed a model with
the energy equation including the effect of crystallization and applied it to
numerical simulation of the high speed spinning.
Spruiell et a1.[38-40] extended this procedure in the melt spinning
process of Nylon. They included in their consideration the effects of
temperature and orientation to the crystallization kinetics.
In the high speed spinning, there exists viscosity distribution in the
fiber cross-section and it results in the distinct skin-core structure. Since
such viscosity variation may become a cause of necking, it should be
also taken into account in the numerical simulation.
Shimizu[4] calculated the temperature distribution in the radial direction
using 1-dimensional numerical modeling. Hutchenson et al.[41] calculated
the radial temperature distribution by solving energy and continuity
equations in the cylindrical coordinate. Bell and Edie[42] performed the
same work with the application of a finite element method.

1.4. Objective of This Work


To investigate the high speed spinning process in terms of the
numerical analysis is the main objective of this work.
1-dimensional numerical analysis enables one to predict the diameter
and temperature variation along the spinline, but with this simplified
method, it is not possible to describe the occurrence of discontinuous
peak in the temperature distribution caused by the crystal formation and
the formation of skin-core structure.
In this work, the spinning process is assumed to be in cylindrical
symmetry and the 2-dimensional numerical analysis has been carried out
with a finite element method, neglecting the variation in the angular
direction. Solving the energy and momentum balance equations
simultaneously, we could estimate the temperature distribution in the fiber
and describe the skin-core formation when calculating the crystallinity
distribution.
With the method of finding free surface employed in the typical finite
element method, it is hardly expected to describe the neck formation in
the high speed spinning and it is not easy to include in the formulation
the effect of cooling, convection and air drag on the free surface.
1124

Therefore, in this study, the temperature and fiber diameter distributions


are estimated by the relatively simple 1-dimensional numerical analysis.
Then, with the boundary conditions specified by this simple calculation,
solutions of the 2-dimensional problem are obtained, which give
information on the fiber morphology and variation of some related
properties.
Effect of take-up speed, cooling air velocity and flow rate on the fiber
structure and on the skin-core formation is also estimated.

2. NUMERICAL SIMULATION

2.1. One-Dimensional Numerical Analysis


2.1.1. Governing Equations
Generally, there exists transfer of heat as well as momentum with
possible variation of the material density in the polymer processing
operations. Thus for detailed analysis, it is necessary to take all of
continuity, momentum, energy and rheological equations into consideration.
In the mathematical formulation, the followings are assumed:
(1) The cross-section of a fiber extruded from the circular orifice remains
circular.
(2) The fiber spinning is performed in a vertical direction.
(3) Flow is in a steady state.
(4) Property variation in the fiber cross-section is neglected.
(5) Surface tension of the fiber surface is neglected.
(6) Heat generation due to viscous flow is neglected.
(7) Heat conduction in the spinning direction and radiation is neglected.
With these assumptions, a set of governing equations in the fiber axis
direction(z-axis) is as follows[4,25,26,43]:
Continuity equation: W- xD2
4 Ov (3)
where W is the mass flow rate, D the fiber diameter, 0 the fiber
density and v the fiber velocity.
Momentum equation: dF
de - ~ + _ ~ Oai,,C/v 2 __ Wg
v (4)

where F is the tensile force acting along the fiber, par is the air density,
1125

of which the value is 0.815x10 3 g/cm3[20], Cf is the surface friction


coefficient and g is the gravity acceleration.
Energy equation is reduced to the following including the effect of
crystallization heat during the high speed spinning, which is shown as the
second term in the right hand side of equation:
Energy equation: dT zcDh( T - T a i r) AHI dX
de- + C, de (5)
where T is the temperature, h the heat transfer coefficient, lair the air
temperature, Cp the specific heat of the polymer, A Hf the heat of fusion
of polymer and X the crystallinity. For PET, A HT = 29 cal/g[4]. PET
shows the behavior of constant viscosity at low shear rate(<100 s l) in
the temperature range of 250---290~ but at higher shear rate(>100 s-~) it
shows the shear thinning behavior within the same temperature range. To
accommodate this non-Newtonian behavior, the power law model shown
in equation (6) is adopted for this analysis.
Rheological equation of state: a = r/~ ~ (6)
where a is the tensile stress, ~e the elongational viscosity, n the
exponent of the power law equation and the value of 0.9 is chosen in
this work.

2.1.2. Crystallization Kinetics


For the analysis of crystallization in the high speed spinning process,
the following Avrami equation with the modification introduced by
Nakamura et a1.[44] is employed:

xo~X - 1 - e x p _[ - ( f_ ~tK( T)dt),o] (7)


where Xoo is the maximum crystallinity, K(T) the crystallization rate
constant, t the time, na the Avrami exponent. The Avrami equation is
thought to describe the initial crystallization stage only, but in this
research we assume that it holds in the emire range of crystallization
process with na = 1 specified[37].
The transformation of the time derivative of the cry stal l inity to the
space derivative yields:
1126

d X - X~176n vflf { xooX } (8).

Half time of crystallization(tv2) defined by Ziabicki [3 0] becomes:


1. = 1 e x p [ - ( 4 1 n 2 ) ( T - Zmax )2]
tl/2 t~/2 Do (9)
where Tmax is the temperature at the maximum crystallization rate, Do is
the half-width obtained from the crystallization rate-temperature curve, and
t~/2 is the minimum value of tl/2 which coincides with the half time at
Tmax.
The crystallization of polymers may depend not only on the thermal
condition but also on the molecular orientation in the amorphous region,
and when we include this effect, tl/2 becomes[30]:
1 _ 1 {Af2 a
tl/2(fa) -- tin(0) exp + Bfa + ""} (10)
where fo is the orientation factor in the amorphous region and A, B, .-.
are parameters which are in general functions of temperature. When the
orientation is very low, i.e. focal, the second and higher order terms in
the parenthesis of the right hand side of equation (10) may be neglected.
Combining equation (10) with (9) with the following relation taken into
account:
K( T, f~) - ln2 (11)
tl/2
results in the final form for the crystallization rate constant:

Do + Aft2 (12).
The values of parameters in the above equations have been specified by
Ziabicki[44] and Alfonso et a1.[46] as:
A = C1 - C2zIT
(~T) 3 (13)
AT: TOm- T (14)
where C1=3.09 l~ C 2 = - 1 . 5 5 t~/z=42s, T~176 D0=64~

2.1.3. Boundary Conditions and Supplementary Relations


Boundary conditions for extrudate velocity, temperature and crystallinity
at the spinneret exit( z = 0) and the take-up point( z = L) are given as:
1127

v=v0, T= To, X = 0 at z = 0
v-- VL at z = L (15)
where v0 and To are the output velocity and the temperature at the die
exit, respectively, and vL is the take-up velocity.
The heat transfer coefficient was modeled by Kase and Matsuo[32] for
the forced convection by cooling air as:
0 D
[1 + ( 8 a;r) ]0
u (16)
where k~g~ is the thermal conductivity of air and its value is 5.50 10 -4
cal/cms~ v~ the velocity of cooling air, ReD is the Reynolds
number for the fiber diameter defined as:
vD
ReD = 0.29 (17).
The elongational viscosity ~e is in general a function of temperature
and crystallinity as assumed by equation (18) and the surface friction
coefficient is expressed in terms of Reynolds number[7] :

7e = 0.73exp T + 273 exp a ~ (18)

Cs= 0.5R@ ~ (19)


where the constants, a and b are assigned to be 4 and 2, respectively[38].
Such empirical formulae as the following equations are also employed
in this study[33-37]:
o = 1.356 - 5x 10-4T (20)
C~ = 0.3 + 6x 10-4T (21).
Birefringence zJna manifested by the molecular orientation in the
amorphous region may be calculated by the formula[47,48]:
Aria = Coba (22)
where Cop is the stress-optical coefficient, the value of which is 7.8 10 -9
mZ/N. The orientation factor fa in the amorphous region and total
birefringence A n can be estimated using the equations:
f_ Aria (23)

An = 1 xooX )Aria + x~o f c A o (24).


1128
Here fc is the orientation factor in the crystalline region and z/~ and z/~
are the intrinsic birefringences of the amorphous and crystalline regions,
respectively. Their values for PET obtained by Dumbleton et a1.[49,50]
are ,d o = 0.275 and ,d o = 0.22.
Equation (24) fits rather well with the experimental data for the low
speed spun fibers, but some deviation from the data can be observed for
the high speed spun fibers. Therefore, for its remedy Katayama and
Yoon[37] proposed the following relation, derived on the basis of the
rubber elasticity theory:

z/n =0.211 - e x p ( - 1.65x 10 -6 T+a273 )] (25).


Numerical solutions for each process condition are then obtained with
the application of above relations and material constants to the
1-dimensional formulation.

2.2. Two-Dimensional Finite Element Analysis


2.2.1. Governing Equations
Equations for the 2-dimensional finite element calculation of the high
speed spinning are formulated in the cylindrical coordinate system (r, 8,
z) under such assumption that the flow in this spinning process is
axisymmetric. Figure 1 shows the melt spinning procedure and coordinate
system employed in this analysis.

RO

Die i z:0

Godet '~_ z=L


Roller

Figure 1. Schematic representation of the fiber spinning process.


1129

Following approximations are assumed to be valid in constructing the


goveming equations:
(1) The cross-section of spun fiber is always circular.
(2) The fiber spinline is in a vertical direction.
(3) The flow is in a steady state.
(4) Heat generation due to viscous flow is neglected.
Let u, v and w denote the components of a velocity vector v in r,
0 and z-directions, respectively, assuming that the flow is axisymmetric.
Then v vanishes, and u and w become functions only of r and z,
that is,
v = (u, v, w) = [u(r, z),O, w(r,z)] (26)
which yields the following expression for the strain rate tensor D:

bl,, r 0 89 ( U, z -[- W, r)

D = 0 _u 0 (27)
l
(u,~+w,~) 0 w,~

where u , is defined as - Ou
~, etc.
The second invariant of D is defined as:
lid = u, 2~ + ( uT) 2 + w , 2~+ 1 ( u , ~ + w , ~ ) 2 (28)
and thus, the total set of equations in the 2-dimensional formulation can
be expressed as the following.
Momentum equation:

-- D, r "[- rrr , r+ Z'rz,Z _+_ ~'rr--~r r O0 Af_ f r : P( UZr ?"-+- WU, z) in r-direction (29a)

__ 1), _~_ rrz r_~_ ~'rz _~_ rzz z_~_f z = P( 1,1W, -~- WW, z) in z-direction (29b)
where p is the pressure, fr and fz are r and z-components of gravity,
respectively, r0 the components of the stress tensor in the cylindrical
coordinate, and the second assumption assigns f~ = 0 and fz = p g.
Energy equation: pC~(uT, r+ wT, z ) - k(T, r~ + + T,r-'~" T , zz) -[- (/V (30)
where k is the thermal conductivity of polymer and I;V is the
crystallization heat.
1130

Continuity equation: u,~ + __u + w,~ = 0 (31)


Y
n-1
Rheological equation of state: 7(IID) = rj~(4IID) 2 (32)
where n is the exponent of the power law equation and 77 is the
viscosity function.
Other relations for the crystallization kinetics and the temperature
dependence of the elongational viscosity are given by those used in the
1-dimensional numerical analysis.

2.2.2. Finite Element Method


In solving the above equations by the Galerkin method, the following
weak form is made with the biquadratic polynomial for the interpolating
functions ~. [51,52]:

fx2r~J'i [ -- P, r + rrr, r + r,~ + r,, -r too _ O( uu, ~ + wu, ~) d~ = 0 (33a)

fs~rgri [ - P' + r'~ '~ + r'~


r + r= '~ + O g - p(uw' ~ + ww' ~)] d~ = (33b).
Here ~ is the domain in the (r,z) plane in which this problem is
defined. Momentum equations are then modified with the application of
the product rule and the divergence theorem:
fs~ [rgi. ~ ( - P + rrr) + rgi, zrrz + ~[l'i(--P + Z'00) (34a)
+ or!if'i( uu, r -F WU z) ]df2 -- jo Q r~i~J r &

fQ [ 7"~J'i, rrrz q- 7"~i,z( -- p Jr rzz ) Ji- pT"~l"i( ZlW, r -t- WW, z ) ] d ~ (34b)
= f r~ipgds foar~i~Szds
where s denotes the boundary 0~2 of the domain ~ . ~r and ~z are the
components of air drag acting on the fiber surface in r- and z
-directions, respectively, and the related boundary conditions are expressed
as:

r'~r = (__ p_Jf_ rrr)nrq- rrznz "-- tYrrnr'+" crr~nz (34c).

Here a0 is the total stress tensor component and n~ the component of a


unit vector normal to the fiber surface. Substitution of the rheological
relation into the above equations yields:
1131

f, I"

(35a)
+ r+WU, Lo
f Y[ 7]~i, r( U, z "~- W, r) -~- ~i, z ( -- P "~ 2 7]W,z) -lt- i0 ~J'i( UW, r ~- WW, z)]d,~
f~
(35b).
= J r~, og d ~ + Jo,.car~i ~ z ds
Applying the product rule and the divergence theorem results in the
following form for the energy equation:
f~ [oCprgi( uT ~+ wT,~) + kr( ~ri,rT, r+ ~i, zT, z)]d~
f'l l'~
(36)
= J,.c2
[ r g i f V d f 2 + k [J O.(2
Y~i T , n ds
where

T,~ = T,~ n r = hk ( Z - T air) (37)


and the quantity T,n refers to the convective heat transfer.
Employing the bilinear interpolating function r we can express the
continuity equation as"
~ r~i( U, r+ Zt w,~)dO = 0 (38).
Here each approximating solutions are in the form of:

~( r, z) = z ~ bli~i( ~, ~) , ~( r, z) -- z~= Wi~i( ~, ~) (39)


1,#" I B

where ~e and ~" are coordinates in so called the master element and M
is the number of nodal points.
Substitution of the above relations for the governing equations gives us
a system of nonlinear algebraic equations. In order to solve this set of
nonlinear equations, we employed the simple Newton iteration method. In
this numerical computation, the iteration does not stop until the obtained
values lie within a desired tolerance limit.

2.2.3. Boundary Conditions


Boundaries of a fiber in the spinline assigned in this finite element
1132

method are illustrated in Figure 2. The line D-A in the figure denotes the
center line of the fiber, i.e. the line of symmetry axis.

Die Exit
A B

Fiber Free
Center , , ~ Surface
Line

C
Godet
Roller

Figure 2. Definition of boundaries in the flow domain.

The boundary conditions can be stated as follows:


At the nozzle exit: z - - 0 (A-B) (40)
where w0 is the volumetric flow rate divided by the nozzle cross-section
and To is the temperature at the exit.
Free surface : r = R(z) (B-C)
u(r,z)=O, w ( r , z ) = wo, T ( r , z ) = To
On the fiber surface, normal components of both velocity and stress
vanish and heat transfer results only from convection, thus, the boundary
conditions become"
ni r O (41)
n i Vi ~- 0 (42)
h (43)
Yl i T , i = - -~ ( T - Tai,,)
where n~ is the unit normal vector on the free boundary and v, is the
component of velocity vector. Equation (42) is the necessary condition for
the boundary surface, and in this case the function for the fiber diameter
determined from the 1-dimensional numerical analysis is used instead.
Temperature variation obtained from the 1-dimensional computation is
1133

also assumed as the boundary condition instead of equation (43) for


2-dimensional calculation.
Take-up p o i m : z = L (C-D)
w( r, L ) = wr~ ' OT
Oz - 0 (44)
where wL is the take-up velocity.
Fiber center line : r = 0 (D-A)
Since there is neither flow nor heat transfer in the radial direction on this
center line, the following conditions are valid:
u=0, aT
Or - 0 (45t.

3. RESULTS AND DISCUSSION

3.1 Behavior of Power Law Fluid


Since the power law fluid assumption is used in this study of the high
speed spinning, the dependence of fiber diameter, velocity, temperature
and crystallinity on the power law index is first estimated. Figures 3-6
show the results of calculation when the tensile stress acts only on the
amorphous region, which is the only deformable region in the fiber. As
shown in Figure 3, when the index n is small, the crystallinity becomes
very low and thus, the crystallization process is slow and the onset point
is located rather far from the spinneret. Figures 4 and 5 demonstrate that
diameter variation becomes slow and the neck-like deformation and small
plateau caused by exothermic crystallization disappear as the index n
decreases.
Velocity distribution depicted in Figure 6 shows some singular point (a
point of unsmoothness) when n is small, as if it was under a vigorous
solidification condition. Hereinafter, n - 0.9 is assigned to the whole
numerical analysis, since this value gives the best fit for the behavior of
PET.
1134

50
1.0 (Newtonian) 6 km/min
4O

~
~ 30

~
r~
20

10 n=0.7

I I I

0 50 100 150 200


Axial distance (cm)
Figure 3. Crystallinity along the spinline at various power-law indices.

0.1
6 km/min

0.7
0.01 0.8
0.9
?5 (Newtonian)

0.001 I I I

0 50 100 150 200


Axial distance (cm)

Figure 4. Fiber diameter along the spinline at various power-law indices.


1135

300
6 km/min

n shift(cm)
200 //~ 0.7 30
~D

.+.a
0.8 20
0.9 10
._
100

I I [

0 50 100 150 200


Axial distance (cm)

Figure 5. Temperature distribution along the spinline at various power-law


indices.

6 km/min

6
n

0.7
4 0.8
o
0.9
CD
1.0 (Newtonian)
> 2

I I [

0 50 100 150 200


Axial distance (cm)

Figure6. Velocity distribution along the spinline at various power-law


indices.
1136

3.2. Attenuation along the Spinline


Of all factors affecting fiber dimension and properties, such spinning
conditions as take-up velocity, quench air velocity and flow rate seem to
be the most important.
Figure 7 shows that fiber diameter attenuates slowly at such low
take-up speed as 4 km/min. However, as one increases the spinning
speed, it decreases rapidly to its final value, as if a neck-like deformation
occurs in the spinline. This is because the crystallization proceeds rapidly
under the high speed and the extrudate solidifies quickly. Hence, the
deformation is concentrated on the narrow region in the amorphous state.
Figure 8 demonstrates the effect of quench air speed on the fiber
diameter at the spinning speed of 6 km/min. From this result, we can
draw a conclusion that the quench air speed does not have a significant
influence on the fiber diameter.
Effect of mass flow rate on the fiber diameter is illustrated in Figure
9. It is evident that the fiber diameter, and hence, the distance of the
onset point of crystallization from the spinneret increases with the
increase in the flow rate.
0.1
V ai r --'- 2 0 m / m i n

W = 1.5 g/min, hole

0.01 ~ 4 km/min
9 v,,,,q

"Q~ 6

\ ~ ~ _ :__,__:, . . . . . . . . . . . . . . . . . . . . . . . . ,

0.001 I I. . . . . I

0 50 100 150 200


Axial distance (cm)
Figure 7. Diameter distribution along the spinline at various take-up speeds
(symbols are for experimental results).
[] ~ 4 km/min A ... 5 km/min o--- 6 km/min
1137

0.1
v = 6 km/min
W = 1.5 g/min, hole

0
~-- 0 m/min
o 0.01 20
cO

3O
?5

0.001
0 50 100 150 200
Axial distance (cm)
Figure 8. Diameter distribution along the spinline at various quench air
speeds (symbols are for experimental results).
[] 0 m/min 9 20m/min o--- 30 m/min

0.1
v = 6 km/min
Uair = 20 m/min

(D \~"~'~
\
2.0 g/rain, hole
r 0.01 ',,\ ~_1.5
"\ 1.0
?5

.o _ ----- . . . . . . . . . . . . . _ . . . . . . . . . . . . . . . . . .

0.001 I I I

0 50 100 150 200


Axial distance (cm)
Figure9. Diameter distribution along the spinline at various mass flow
rates(symbols are for experimental results).
n 1.0 g/rain 9 hole A ... 1.5 g/min, hole o--- 2.0 g/min 9 hole
1138

In Figures 10-12, we can observe the temperature profile along the


spinline as a function of take-up speed, quench air velocity and mass
flow rate, respectively. The temperature decreases slowly (slow in a sense
not of time but of distance), as the spinning speed increases. Similarly to
the previous result on the fiber diameter, the change of quench air speed
rarely alters the temperature profile as shown in Figure 11. We can
examine the significant effect of flow rate in Figure 12. As the flow rate
increases, cooling becomes very slow due to the increase of fiber
diameter, and therefore quick temperature variation is suppressed.
Figures 13-15 are the results for crystallinity variation. Figure 13
clearly shows that the rate of crystallization and the final value of
crystallinity increase as the spinning speed increases. Again, the effect of
quench air speed on the crystallinity may be regarded as negligible.
Concerning the effect of flow rate, even though the crystallization starts
later, the equilibrium value of crystallinity increases with its increase.

300
- v air = 20 m/min
', \ W = 1.5 g/min, hole

~ 200 shift(cm)
"- x\ 4km/min 0
"",, \ \ / / - - 5 10
" ~,</--- 6 20
100
% . %~
".o 9

I, I I

0 50 100 150 200


Axial distance (cm)
Figure 10. Temperature distribution along the spinline at various take-up
speeds.
4 km/min ... 5m/min --- 6 km/min
1139

300
- v = 6 km/min
_ ~ W = 1.5 g/min, hole

~ 200
0 m/min

100

0 50 100 150 200


Axial distance (cm)
Figure 11. Temperature distribution along the spinline at various quench air
speeds.
0 m/min .-. 20 m/min --- 30 m/min

300
-', v = 6 km/min
~ v~i~ = 20 m/min
\,,,
200

\, \ \

100 -
~ "'",,%"",,2.0
~
g/min, hole

I I I

0 50 100 150 200


Axial distance (cm)
Figure 12. Temperature distribution along the spinline at various mass
flow rates.
1.0 g/min 9 hole ..- 1.5 g/min- hole --- 2.0 g/min 9 hole
1140

50
6 km/min
tl'"
40 #

!
I
I

o~..i
30 I
I
I
o~..~ I
/ ,;~
I
t~ /I ..~
20 I '
I :
r,.)
10 ,i"U
// Vair~" 2 0 m / m i n

W = 1.5 g/min, hole

0 50 100 150 200


Axial distance (cm)
Figure 13. Crystallinity along the spinline at various take-up speeds
(symbols are for experimental results).
[ ] ~ 4 km/min 9 5km/min o--- 6 km/min

50

40

o,.-i
.~.,i

t~
ra~

r,.)
30

20
! f 0 m/min

30

10 v = 6 km/min
W = 1.5 g/min, hole
1 I

0 50 100 150 200


Axial distance (cm)
Figure 14. Crystallinity along the spinline at various quench air speeds
(symbols are for experimental results).
[] 0 m/min 9 ..- 20 m/min o--- 30 m/min
1141

50
2.0 g/min, hole
,'7 ..... ]'13- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
40
'~ 1.0

30
///
4~
~ 20
'///
10
'J;,/
!,
;. ,
... i
v = 6 km/min
U ai r =

i
20 m/min

0 50 100 150 200


Axial distance (cm)

Figure 15. Crystallinity along the spinline at various mass flow rates
(symbols are for experimental results).
[] 1.0 g / m i n - h o l e 9 1.5 g/min, hole o--- 2.0 g/min- hole

3.3. S k i n - C o r e F o r m a t i o n in the Fiber Cross-Section


One of the most important characteristics of high speed spinning is the
distinct structural difference formed along the fiber cross-section due to
different thermal and stress history applied during the spinning process.
This is so-called the skin-core structure, and its formation mechanism can
be explained with the help of results obtained in the 2-dimensional
simulation.
Figure 16 shows the temperature change at the skin and core regions
of the fiber cross-section. Fibers extruded from the spinneret are cooled
by the ambient quench air. Fiber skin has direct contact with the cooling
air, and thus, it is cooled down more rapidly than the fiber core. This
temperature difference between the skin and core induces the difference in
crystallization process and as a result the skin-core structure forms. Figure
17 demonstrates the difference between the crystallinities in skin and core.
1142

Calculated results on radial distribution of the temperature in the fiber


cross-section at various axial positions, are also presented in Figure 18,
which simply illustrates monotonic decrease of temperature with radius.
However, the temperature difference between the skin and the core
becomes insignificant at the point far from the spinneret, where the
temperature of the fiber becomes close to that of the surrounding air.
Figure 19 shows the radial distribution of crystallinity at various points
in the axial direction. The crystallization starts at the skin near the
spinneret exit, then it proceeds slowly along the spinline, and crystallinity
finally reaches the equilibrium distribution. It may be interpreted as that
the crystallization process is initiated by the cooling and also the
following stress development in the skin, and thus the crystallization
propagates from the skin to the core.

300
v = 6 km/min

200
O

100 skin

1 I _ l

0 50 100 150 200


Axial distance (cm)

Figure 16. Temperature variation along the spinline in the skin and core of
the PET fiber.
1143

skin

core

.l..,a

~" 0.5
(D
~ 1,,,,i

v = 6 km/min
. I I
0
0 50 100 150 200
Axial distance (cm)
Figure 17. Relative crystallinity along the spinline in the skin and core of
the PET fiber.

300
v = 6 km/min
z = 20 cm
30
200
45

~' 100
75

115
I I I [

0 0.2 0.4 0.6 0.8 1


Normalized radius (r/R)

Figure 18. Radial distribution of temperature at various positions in the


spinline.
1144

~ v

z =60 cm
v = 6 km/min

o
0.5

35

33 -----------
i i -----------'-~ -~

0 0.2 0.4 0.6 0.8 1


Normalized radius (r/R)

Figure 19. Radial distribution of relative crystallinity at various positions


in the spinline.

3.4. Orientation Distribution in the Fiber Cross-section


At the spinning speed of 6 km/min, it is thought that the molecular
orientation develops mainly in the skin region before the point of
necking, then after the completion of necking the orientation in the fiber
core increases to the amount comparable to that in the skin. However, at
the speed of 8 km/min or above, the stress concentration in the skin
region prevents the formation of high degree of orientation in the
amorphous region of fiber core, hence the distinct skin-core structure is
formed.
In Figure 20, one can see the calculated results of radial distribution of
refractive indices in both parallel (n/l) and perpendicular (n directions.
While n does not vary much along the radius, n// exhibits strong
dependence on the radial position.
1145

The birefringence obtained from such data in Figure 20 as is plotted


on Figures 21-23. At low take-up speed it does not show much variation
along the radial direction, but at 6 km/min its dependence becomes
noticeable (Figure 21). From Figure 22, we can observe somewhat
significant effect of quench air speed on birefringence. As can be seen in
Figure 23, the increase of the mass flow rate, the difference between the
birefringence in the skin and the core becomes larger, since the increase
of fiber diameter magnifies the structural difference between those regions.

1.68 1.54

1.67 g/

x~
9 ,,,,-i

~ 1.66 1.535
9 ,-,,i

~ 1.65 v = 6 km/min
v~i~ = 30 m/min
W = 2.0 g/min, hole
I q q
1.64 1.53
0 0.2 0.4 0.6 0.8 1
Axial distance (cm)

Figure 20. Refractive index profile across the fiber cross-section for PET
fibers.
parallel (n/l) --- perpendicular ( n l )
1146

140
Pair = 30 m/min

W = 1.5 g/min, hole


. . . . . . . . . . . . . . . . .

_
x 120

I I 1 I
80
0 0.2 0.4 0.6 0.8 1
Axial distance (cm)
Figure21. Birefringence profile across the fiber cross-section for PET
fibers at various take-up speeds.
4 km/min ... 5 km/min --- 6 km/min

140
0 rn/min

30
x 120

em

100
cD
o~.~

v = 6 km/min
W = 1.5 g/min, hole
1 I I
80
0 0.2 0.4 0.6 0.8 1
Axial distance (cm)
Figure22. Birefringence profile across the fiber cross-section for PET
fibers at various quench air speeds.
0 m/min ... 20 m/min --- 30 m/min
1147

140

1.5

""
X
120 10 ...................

2.0 g/min, hole

N loo
.~,,i
v = 6 km/min
Uair-- 3 0 m/min
I
80 I I

0 0.2 0.4 0.6 0.8 1


Axial distance (cm)

Figure23. Birefringence profile across the fiber cross-section for PET


fibers at various mass flow rates.
1.0 g / m i n - h o l e ... 1.5 g/min, hole --- 2.0 g/min- hole

4. CONCLUSIONS

In this study, we have performed the analysis of high speed spinning


process, employing some hybrid numerical methods combining both 1-
and 2-dimensional formulations. The radius of the extrudate and the
temperature on its surface are calculated under the 1-dimensional flow
assumption, and then with these results specified as boundary conditions,
2-dimensional simulation is conducted to estimate the temperature profile
and the morphological structure development along the fiber cross-section.
Including the effect of crystallinity in the elongational viscosity and
assuming the tensile stress acts only on the amorphous region enabled us
to observe the neck-like deformation in the spinline.
From the result of numerical simulation, it can be concluded that the
take-up speed plays the most determining role in the fiber structure
formation. Even though the flow rate shows strong effect on diameter and
1148

temperature variation along the spinline, its role diminishes in the case of
crystallinity and birefringence of the final products. The formation of the
fiber skin-core structure is simulated in this numerical procedure.
Effect of spinning conditions on the formation of skin-core structure is
estimated by observing the birefringence distribution in the cross-section.
The distinct skin-core structure builds up due to the combined effect of
high speed take-up and quench air. As the quench air velocity increases
at high speed spinning and the fiber diameter becomes thicker, this
structural inhomogeneity becomes severer.

REFERENCES

1. T. Kawaguchi, High-Speed Fiber Spinning, Chapter 1, John Wiley and


Sons, 1985.
2. R. J. Fisher and M. M. Denn, AIChE J., 22 (1976), 236.
3. J. C. Hyun, AIChE J., 26 (1980), 294.
4. J. Shimizu, N. Okui and T. Kikutani, High-Speed Fiber Spinning,
Chapter 7, John Wiley and Sons, 1985.
5. H. Yamada, T. Kikutani, A. Takaku and J. Shimizu, Sen-I Gakkaishi,
44 (1988), 177.
6. M. Matsui, Sen-I Gakkaishi, 38 (1982), P-508.
7. A. Ziabicki and K. Kedzierska, J. Appl. Polymer Sci., 6 (1962), 111.
8. K. Nakamura, T. Watanabe, K. Katayama and T. Amano, J. Appl.
Polymer Sci., 16 (1972), 1077.
9. T. Kikutani, Ph.D. Thesis, Tokyo Institute of Technology, 1982.
10. K. Kamide and T. Kuriki, Sen-I Kikai Gakkaishi, 38 (1985), P-268.
11. G. Vassilatos, B. H. Knox and H. R. E. Frankfort, High-Speed Fiber
Spinning, Chapter 14, John Wiley and Sons, 1985.
12. J. Shimizu, N. Okui and T. Kikutani, Sen-I Gakkaishi, 37 (1981),
T-135.
13. J. Shimizu, Sen-I Kikai Gakkaishi, 38 (1985), P-243.
14. S. A. Silling, J. Appl. Mechanics, 55 (1988), 530.
15. B. D. Coleman and D. C. Newman, Polymer Eng. Sci., 30 (1990),
1299.
16. Y. Tomita and K. Hayashi, Int. J. Solids Structure, 30 (1993), 225.
1149

17. Y. Tomita, Appl. Mech. Review, 47 (1994), 171.


18. S. I. Krishnamachari, Applied Stress Analysis of Plastics, Chapter 3,
Van Nostrand Reinhold, 1993.
19. G. Perez, High-Speed Fiber Spinning, Chapter 12, John Wiley and
Sons, 1985.
20. S. Kase, High-Speed Fiber Spinning, Chapter 3, John Wiley and
Sons, 1985.
21. K. Fujimoto, K. Iohara, L. Owaki and Y. Murase, Sen-I Gakkaishi,
44 (1988), 53.
22. H. Haberkorn, K. Hahn, H. D. Dorrer and P. Matthies, J. Appl.
Polymer Sci., 47 (1993), 1551.
23. A. Ziabicki, J. Non-Newtonian Fluid Mech., 30 (1988), 141.
24. A. Ziabicki and J. Tian, J. Non-Newtonian Fluid Mech., 47 (1993),
57.
25. A. Ziabicki and L. Jarecki, High-Speed Fiber Spinning, Chapter 9,
John Wiley and Sons, 1985.
26. J. M. Schultz, Polymer Eng. Sci., 31 (1991), 661.
27. J. M. Schultz, Polymer, 32 (1991), 3268.
28. H. M. Heuvel and R. Huisman, High-Speed Fiber Spinning, Chapter
11, John Wiley and Sons, 1985.
29. A. Ziabicki, J. Non-Newtonian Fluid Mech., 30 (1988), 157.
30. A. Ziabicki, Fundamentals of Fibre Formation, Chapter 2---3, John
Wiley and Sons, 1976.
31. S. Kase and T. Matsuo, J. Polym. Sci., A3 (1965), 2541.
32. S. Kase and T. Matsuo, J. Appl. Polymer Sci., 11 (1967), 251.
33. A. Prastaro and P. Parrini, Textile Research J., 45 (1975), 118.
34. Y. Yasuda, H. Sugiyama and H. Yanagawa, Sen-I Gakkaishi, 40
(1984), T-370.
35. Y. Yasuda, H. Sugiyama and H. Hayashi, Sen-I Gakkaishi, 40 (1984),
T-227.
36. M. M. Denn and D. K. Gagon, Polymer Eng. Sci., 21 (1981), 844.
37. K. Katayama and M. G. Yoon, High-Speed Fiber Spinning, Chapter
8, John Wiley and Sons, 1985.
38. R. M. Patel, J. H. Bheda and J. E. Spruiell, J. Appl. Polymer Sci.,
42 (1991), 1671.
39. K. F. Zieminski and J. E. Spruiell, J. Appl. Polymer Sci., 35 (1988),
1150

2223.
40. J. H. Bheda and J. E. Spruiell, J. Appl. Polymer Sci., 39 (1990),
447.
41. K. W. Hutchenson, D. D. Edie and D. M. Riggs, J. Appl. Polymer
Sci., 29 (1984), 3621.
42. W. P. Bell and D. D. Edie, J. Appl. Polymer Sci., 33 (1987), 1073.
43. A. Dutta, Text. Res. J., 57 (1987), 13.
44. K. Nakamura, K. Katayama and T. Amano, J. Appl. Polymer Sci., 17
(1993), 1031.
45. A. Ziabicki, Kolloid Z., 175 (1961), 14.
46. G. C. Alfonso, M. P. Verdona and A, Wasiak, Polymer, 19 (1978),
711.
47. I. Hamana, M. Matsui and S. Kato, Melliand Textilberichte, 50
(1969), 382.
48. I. Hamana, M. Matsui and S. Kato, Melliand Textilberichte, 50
(1969), 499.
49. J. H. Dumbleton, J. Polymer Sci. A-2, 6 (1968), 795.
50. T. Murayama, J. H. Dumbleton and M. L. Williams, J. Polymer Sci.,
A-2, 6 (1968), 787.
51. E. B. Becker, G. F. Carey and J. T. Oden, Finite Elements : An
Instruction Vol. 1, Prentice-Hall, 1981.
52. M. J. Crochet, A. R. Davies and K. Walters, Numerical Simulation of
Non-Newtonian Flow, Elsevier, 1984.
1151

PHYSICAL MECHANISMS O F F O A M F L O W IN P O R O U S
MEDIA

K o n s t a n t i n G . K o r n e v ~ , A l e x a n d e r V . N e i m a r k b , a n d A l e k s e y N. R o z h k o v ~

Institute for Problems in Mechanics, Russian Academy of Sciences,


101(1) Prospect Vernadskogo, Moscow 117526, Russia
kornev@ipmnet, ru, rozhkov@ipmnet, ru

b TRI/Princeton, 601 Prospect Ave., Princeton, NJ 08542-0625, USA


aneimark@triprinceton, org

1. INTRODUCTION

The cell structure of foams can be observed in many daily situations. We


encounter many examples of foam irregularity, instability, chaos and the like in
everyday practice. The specific properties of foams have been attracted the
attention of researchers over the past few centuries. At the present time, there
exists a substantial collection of books and reviews which cover a wide range of
fimdamental questions. Books and reviews written by Boys [1 ], Mysels, Shinoda
and Frankel [2], Bikerman [3], Ivanov [4], Kraynik [5], Wilson [6], Kruglyakov
and Exerowa [7], Stavans [8] and Prud'homme and Khan [9] are only some of
them. In confining systems such as porous materials, foam has been recognized as
a fluid with a tmique hierarchical structure and rheological properties. In recent
decades, due to novel applications, foams in porous media have become an
intriguing subject of fimdamental, systematic studies by physical scientists. In
particular, because of their extremely efficient blocking action, foams open wide
horizons for applications in problems of groundwater/soil remediation, selective
blockage and other processes of sweeping out a given porous material of either
oil or another liquid. Foams found various applications in industrial technologies
dealing with fiber structures, such as dyeing and finishing of textile fabrics, paper
coating, resin-impregnation of fibrous mats and fabrics. However, while
technologically effective, foams distributed in porous media or residing in contact
with highly dispersed materials remain enigmatic. In this chapter, we discuss
1152

some physical problems associated with foams in porous systems. We do not


attempt a complete review. Previous reviews of Raza [10], Marsden [11 ], Heller
and Ktmtamukkula [12], Schramm [13] and Rossen [14] give a useful guideline in
this branch of studies. Many important problems remain beyond the scope of this
survey. Among them are mechanisms of foam generation in pores, interactions
between foam lamellae and the wetting films clinging to the pore walls (Kheifetz
and Neimark [15]), a role of disjoining pressure acting in thin films (Derjaguin, et
al [16]; Neimark and Vignes-Adler [17]), transformations of lens into lamella
(Ivanov [4]; Kruglyakov and Exerowa [7], bubble transport through smooth
capillaries and physics of dynamic wetting (de Gennes [18]), mechanistic models
of foam evolution in the bulk and in porous media, which are based on the
population balance method (Falls, et al [19]), and also percolation models of foam
flow in pore networks (see the latest review of Rossen [14] and references
therein). We focus on the physical mechanisms which govern foam flow in
confining systems.

2. FOAMS IN POROUS MEDIA

Although bulk foam, as a matter for scientific inquiry, has a great history, the
problem of foam patterning in confining systems, such as porous media, drew the
attention of scientists only recently. Apparently Fried [20] was the first to show
that foam, because of its unique structuxe, reduces gas flow in porous media. In
recent decades, it has been recognized that foam within porous media indeed
possesses a unique structure and rheological properties. In particular, in both pre-
and in situ-generated foams, pore sizes impose some constraints on the foam
texture. If the characteristic bubble size is much smaller than the characteristic
pore size, the foam confined in pores does not differ from bulk foam. In the
opposite case, the foam in a porous medium is a system of thin liquid films
(lamellae) spanning the pore channels.
The transition from the bulk phase to the new "pore-confined" regime of foam
patterning is very complex, and is discussed, as a rule, for ideal homogeneous
capillaries or bead packs (Hirasaky and Lawson [21]; Nutt and Burley [22]).
However, a more reasonable conceptual model of natural porous media is a
network of interconnected capillaries of different sizes, which may contain
constrictions and enlargements. Recent experiments (Bazilevsky, et al [23]) show
that the motion of pregenerated foam through such a system demonstrates specific
characteristic patterns.
1153

Figure 1. The sequence of frames depicts the displacements of the lamellae in the
enlargement of the channel. The flow direction is from the bottom upwards. The
white and black arrows indicate the positions of the pertinent lamellae at different
moments of time. The channel consisted of 10 cubic cell elements, an 8 x 8 m m 2
cross-section, and 10 narrow elements 8 r a m long, and a 2.5 x 2 . 5 r a m 2 cross
section.

The main peculiarities of the phenomenon are as follows (Figure 1). Just after
the beginning of their injection, the lamellae move freely along the narrow part of
the channel. If the broad part is free of foam, the lamella, reaching this point,
stops for the present, not being able to overcome the enlargement of the channel.
Under the action of pressure supplied by the gas, the lamella swells, forming a
bubble. The bubble either disappears, and the bubble swelling process restarts
when the next lamella reaches this point; or, after touching the wall of the
enlargement, the lamella jumps to the free zone and stops, clinging to the walls of
the wide pore somewhere. The following lamella behaves similarly. The
advancing bubbles jump sequentially from the narrow to the wide part of the
channel and come to a standstill in the engagement with the walls or other
lamellae. This process continues until the wide part of the channel is filled
entirely with foam, whereupon the lamellae are free to pass through it. In the next
enlargement, the process of foam filling repeats itself. And all the wide parts of
the channel are gradually filled with bulk foam. In doing so, the specific film
channels of approximately the size of the narrow part of the original channel
arise. With all the broad parts filled, the lamella motion proceeds as follows. The
drift of lamellae along the narrow zone alternates with periods of motion through
1154

the effective channels in the enlargements. In the last case, lamellae are subjected
to transformation, and their motion is jumpy. (It is related to the tendency of the
lamella system to minimize its surface energy.) No processes of rupture or birth
of the moving lamellae are observed. Due to the manner of the foam generation,
the distance between the drifting lamellae is of the same order as the channel
period. When lamellae are passing each other, the bulk foam in the enlargements
also undergoes deformations. However, its structure is not changed. It seems to
be the same before and after each lamella passes.
Thus, the experiments show that, during the motion of foam through the channel
of an abruptly varying cross-section, a portion of it is used to form the effective
channel. Such a channel plays the role of a transport path for another portion of
foam. In the process, the foam lamellae demonstrate an astonishing stability.
This and other visual observations (Owete and Brigham [24]; Chambers and
Radke [25]) speak in favor of a crucial role of the solid skeleton in foam
patterning in porous media. Even excluding for the moment the hydrodynamic
features, foams in porous media or in contact with highly dispersed materials,
have a unique structure by virtue of the thermodynamic reasons. Indeed, in the
absence of external restraints, an interface between the gas and liquid phases
tends to minimize its surface free energy by adopting a particular ordered
structure. In the presence of solid surfaces, however, the interface between the
two fluids takes up configurations whose equilibrium and stability depend upon
the areas of contact and the interfaeial free energies of the surfaces separating the
various phases.
The solid skeleton imposes to the bubble system a multilevel hierarchical
structure, in which the surface and capillary forces are self-consistently
interconnected. The wetting films covering the pore walls interact with the
lamellae, while the general conditions of foam equilibrium include primarily the
conditions of equilibrium between these two type of thin films. In many instances,
the thermodynamic properties of lamellae and the wetting films may dominate the
foam behavior in porous media. This type of interactions especially concerns the
strong foams, in which all bubbles are bounded by the lamellae (bubble-bubble
interfaces) and the wetting films (bubble-solid interfaces) (Ettinger and Radke
[26]; Kovscek and Radke [27]; Rossen [14]) (see Figure 2). Namely the strong
foams play a major role in applications due to their strong blocking action.
1155

9 iii
Figure 2. The sketch of the gas and liquid distribution within a strong foam in a
granular medium.

3. STICK-SLIP MOTION OF FOAM L A M E L L A

Pore channels in real materials are not uniform. Alternating constrictions (pore
throats) and enlargements cause variations of the Laplacian pressure across a
lamella moving through the channel. Under these conditions, the resulting motion
of the lamella will be irregular. Indeed, just after the lamella has left the pore
throat, it bulges forward and resists the movement. But as it reaches a convergent
part of the pore, the lamella bulges backward and accelerates. Visual experiments
(Chambers and Radke [25]; Bazilevsky, et al [23]) do reveal that, when pushed at
a constant pressure drop, very often the lamellae do not move steadily, but by an
alternation of slides and periods of rest. This type of motion is called the stick-
slip motion (Rabinowicz [28]; Bowden and Tabor [29]).
Specific features of foam patterning within a porous medium can be revealed on
the simplest yet expository model of a one-dimensional foam, referred to as a
bubble chain or train, immersed into a wavy channel (Komev [30]; Komev, et al
[31 ]; Dautov, et al [32]). Imagine such a channel as a rigid capillary with a radius

(2~x
r=ro+~C~ 2 ~' (1)
1156

where the x - axis coincides with the axis of the symmetry of a capillary, to, ,~,
and J are some characteristic scales of the porous medium. Assume, for
simplicity, that the pore aspect ratio is a low ,~/ro << 1. This assumption,
however, can be altered without change in most of the physical conclusions.

I ~I a)

90 ~ Plateau border

b)

X
----~+i ~. ai
2

Figure 3. a) Scheme of lamella distribution in a bamboo-like channel. Dashed


lines are attributed to the initial positions of lamellae, and bold faced lines
represent the lamellae under a load. b) Specification of the input parameters
needed for calculation of the bubble volume variation and capillary force.
1157

Consider the train consisting of N lamellae, whose centers of chords are


prescribed as points x = a ~ , a 2 , . . . a ~ , and assume that in the initial undeformed
state, the foam is perfectly ordered. In such an ordered foam, the lamellae are
connected by links (bubbles, or train carriages) of length K2 with an integer
number K (K2 is the "carriage length").
Thus, the capillary forces tend to fix the lamellae at the pore throats. However,
the elastic forces, caused by gas compressibility, compel the lamellae to shift into
new equilibrium positions. Competition between these forces does result in the
equilibrium states, which are observed in the experiments with foams in porous
systems. The situation resembles 'commensurate - incommensurate' phase
transitions in solid state physics (Pokrovsky and Yalapov [33]; Bak [34]) and
related problems, in which a struggle between the binding and pinning forces
occurs (Frenkel and Kontorova [35]; Josephson [36]; Kulik [37]; Seeger and
Schiller [38]; Barone and Paterno [39]; Lichtenberg and Lieberman [40]).
Under a load, the i -th bubble is deformed. And its length, with an accuracy of
O ( 6 / r o ) , becomes a , - a,_, (Kornev [30]; Kornev, et al [31]; Dautov, et al [32]).
The resulting elastic force acting upon the i - th lamella can be written with the
same accuracy as

f e = xr21Pi+l - Pi ) (2)

Here P, is the gas pressure in the i-th bubble. Making use of the equation of the
state of the ideal gas, we have with the same accuracy the following equation

Pi +1 (ai +1 - ai ) = Pi (ai - ai-1 ) ="= PgK2 , (3)

where P~ is the initial gas pressure in an individual bubble. It is more convenient


to rewrite equation (3) by introducing a new unknown fimction, p, - displacement
of the i - th lamella from its initial position at the throat (Figure 3). Then
a~ = i K ~ + p, + 3./2 and equation (3) takes the form

PgK2
Pi+l-Pi = p 2K , (4)
i+1

Note that the capillary force has the form, similar to equation (2), but, instead of
the pressure differential, the Laplacian pressure drop,
1158

R-1 (P'+I - P i )
ii+l = 4y (5)

should be inserted. Here 2y is the lamella tension, /~.,+,is the radius of curvature
of a lamella (Figure 3). Then the capillary pressure drop can be written with the
same accuracy of O(6/r o) as

Pi+l-Pi =
8Jr~
2r 0
sinl2~r;i)/ (6)

The stick-slip motion of the bubble tram can be qualitatively described by using
the following model

d2p i
m~ = ~rrgAP (7)
Of 2

where m is an effective mass of the lamella, i is the dimensional time, and AP is


the pressure drop across the lamella. The latter is originated from i ) the 'elastic'
pressure drop, equations (2)-(4), ii) the Laplacian pressure drop, equation (6),
iii ) the dynamic pressure drop, iiii ) and the extemal pressure drop. For clarity,
consider only a quasi-static motion so that all the deviations from the Newtonian
friction are assumed to be small.
The main peculiarities of the stick-slip phenomenon can be elucidated by
analyzing the motion of an individual lamella. In dimensionless variables,
equation (7) written for a single lamella takes the form

e 2 d2pi dPi = Ap-2n-psin (8)


dt---2- + -~- Pi '

2
where e2= 2mh~Pg/9r/;ppr o2 2 and a dimensionless time is introduced as follows
t-2~Pgi/32rL~. Here rl.,~ is an apparent viscosity of a lamella and h~ is the
wetting film thickness. The parameter

4y6 (9)
p = Pgro;~
1159

serves as a measure of the intensity of the capillary forces with respect to the
driving forces (e.g., in the case of a bubble chain, these forces will be the elastic
forces). The parameter zXp denotes an external pressure drop sealed by P~.
Equation (9) has a very broad range of interpretations. It is widely used and
describes various physical systems, depending on the choice of the parameters
(Landau and Lifshitz [41]; Barone and Paterno [39]; Lichtenberg and Lieberman
[40]; Zaslavsky and Sagdeev [42]). In application to the lamella motion, the
parameter G is usually small. Therefore, the first term in the left hand side of
equation (9) dominates the second only for initial instants of time. For an analysis
of the lamella dynamics during a long time interval, the inertial term may be
dropped and, as a result, equation (9) takes the form

dt = A p - 2~r/.t cjn P i ' A p = C o n s t . (10)

This equation can be completely analyzed, because it has the explicit solution
(Aslamasov and Larkin [43]; see also Barone and Patemo [39])

, X~-~ . tan + 2~/~


Ap ][' (11)

where the period T is expressed as

27/"
r = . (12)
IJAp 2 - 4~r2p 2

Notably, the characteristic time of translation over a period of the channel


depends upon the applied pressure drop.
Thus, the periodic motion represented by equation (11) looks like a stick-slip
motion. In the vicinity of a threshold, i.e., when the applied pressure drop is of
the same order as the Laplacian barrier, 6p ~ 2z/~, the lamella spends the main
portion of the time in a slow creep. After overcoming the maximum pinning force
(i.e., after reaching the point p - rc/2 ), the lamella suddenly jumps an enlargement
and then the process of lamella motion repeats itself. As the pressure drop
increases, the jump time becomes negligibly small, and the lamella moves almost
1160

steadily. The lamella velocity, averaged over the period T, becomes the following
nonlinear function of the external pressure drop

0.12

2~./,t~p=0.99
...... -2~Z~ =0.5
- - - 2~/gp =0.1

0.4

--. . . . . . . _ . . . . . . . . . . . - . . . . . . --- ~ 9. . . . . . .

, I , i , i , I , ,

O.O 0.4 0.8


tff

Figure 4. Transition from a stick-slip motion to a sliding motion ( ~ = 10 -3 ).

1 T d ~r 2n = ~Ap 2 - 4n. 2/z 2 (13)


<,b> =~ 0 r= T

with the expected threshold zXp~ p . Formula (13) displays the increase of the
apparent viscosity of the lamella along with the increase of the rate of shear, as
determined by the lamella speed (Figure 5). This is the so-called 'shear-
thickening' effect (Barnes, et al [44]). The averaging procedure which leads to
equation (13) somewhat disguises the cause of the 'shear-thickening' effect. The
physics of such a rheological behavior is as follows. In the vicinity of the critical
pressure drop, the capillary forces drive the lamella almost entirely. Figuratively
speaking, in this critical regime, a 'bare' external pressure drop is needed only to
compel the lamella to shift from its equilibrium position. Once shifted, the lamella
drifts almost autonomously. Only in a high-speed regime, does the apparent
viscosity of the lamella tend to its bare value. Therefore, this 'shear thickening'
behavior of a lamella is caused solely by irregularities of the channel. If we
increase the pressure drop ft~her and further, we will inevitably arrive at the
range of the validity of a Non-Newtonian friction law. The Bretherton friction law
1161

[45] is one of the possible candidates for description of the lamella friction in a
Non-Newtonian range of flow.
We recall that Bretherton [45] analyzed the motion of long bubbles in a
cylindrical capillary of the radius Rand computed the thickness of the film
between the bubble and the capillary wall h~ and the dynamic pressure drop
across the length of the bubble. Both depend on the capillary number
Ca (3Ur//cr), where U is the bubble velocity, 1/ is the viscosity of the wetting
=

film, and cr is the surface tension. Bretherton pointed out that the dynamic
pressure drop across such a bubble, additional to the Laplacian pressure drop, is
given by

(3r/U)2/3 cr
B =3.58.

0.05
J
<~> f

J
0.02
/-
J
0.01

, I I , I j , i ,,, 9 I

GO 0.01 0.02 0.03


Ap

Figure 5. 'Shear-thickening' behavior of an individual lamella. The increase of


viscosity with the increasing rate of shear in an averaged steady flow (p = 10-~).

This result helps to explain the lamella mobility. In particular, Hirasaki and
Lawson [21] automatically spread Bretherton's analysis to a chain of bubbles
separated by lamellae. It is assumed that the bubble train moves through a
1162

uniform capillary in a piston-like manner so that the distance between the


lamellae remains the same. The Plateau border, in their analysis, is modelled as a
region of constant curvature, i.e., the Plateau border is similar to the liquid ahead
of and behind an isolated bubble, except that the radius of the curvature of a
meniscus can be less than the tube radius. Hereby, the characteristic length of the
hydrodynamic perturbations becomes of an order of the size of the Plateau
border, i.e., it is much smaller than the length of the bubble. Within the regions of
uniformity, the wetting film does not move yet resides at the corresponding gas
pressure P, . Because of the difference in the gas pressures within the adjacent
bubbles, there exists a transition region, the Plateau border, in which the liquid
undergoes an action of the respective pressure drop. So, the liquid in the vicinity
of the Plateau borders flows from one bubble to another, 'beating' the lamella. As
a result, the total dynamic pressure drop per lamella becomes proportional to
(Ca)2'3cr/r~ , where r~ is the radius of curvature of the interface at the Plateau
border.
It should be stressed that the theory which deals with a uniform capillary and
which diminishes the effect of gas compressibility underestimates the foam
viscosity under all experimental conditions (Figure 6). In addition, even the
tendency in dependence of the gas-phase velocity is predicted incorrectly.
To resolve the contradictions, Falls, et al [46] modified the Hirasaki-Lawson
theory to account for the contribution of pore constrictions to the apparent
viscosity. The Laplacian contribution to the dynamic pressure gradient was
approximated as v P ~ n14o-/r,, where n~ is the number of lamellae per unit
length and r, is the radius of the pore throat. As a result, the Hirasaki-Lawson
apparent viscosity is modified by augmenting the Laplacian contribution. The
latter is defined through the Hagen-Poiseuille law as, rico,= 4onI/Vgrt, where vg is
the gas-phase velocity. Thus, the expected result is as follows: the apparent
viscosity must vary with the -1 power of the gas velocity at low speeds. Falls
and coworkers studied two regimes of foam generation: the first is the so-called
controlled-bubble-size regime, where the bubble size was unchanged in situ, and
the liquid was transported within the lamellae; the second is the so-called pack-
generated-bubble-size regime, where one bead pack was used to generate the
foam for a second. It has been demonstrated for both flow regimes that the
apparent viscosity measured at low speeds varies as -1 power of the gas-phase
velocity. Moreover, involving the Laplacian contribution of the apparent
viscosity, the authors managed to fit the Hirasaki-Lawson theory to the
experimental data even at higher speeds. They showed that, at higher rates, the
viscosity varies as - 1 ! 3 power of the velocity in the first flow regime, and
1163

changes to - 2 / 3 power of the velocity in the second. The latter result means
that the effect of the surface tension gradient governs the foam resistance in the
second flow regime (Hirasaki and Lawson [21 ]).

10

o / / /
8 ZX C17:~Pt=0 / / / _

0 2 4 6 8 10 12

Figure 6. Typical rheological curves. Foams from 0.5% Hostapur SAS carbon
number fractions. The length of the pack- 100 cm, the permeability-8 darcy, the
pressure at the exiting end -7 bar. AP, denotes the threshold pressure drop (after
Hanssen and Dalland [47]).

Thus, the key assumption in treating the experimental data concerns the form of
the contribution of the pore constrictions to lamella resistance. This term dictates
the tendency in the total changes of the apparent viscosity as a function of the gas
velocity. In the Bretherton flow regime, the 'shear thickening' behavior can be
changed by the 'shear-thinning' (i.e., the reduction of viscosity with increasing
rate of shear in an averaged steady flow). In particular, this fact can be checked
by performing an analysis similar to that expressed by equations (10) and (13),
but modified in accordance with the Bretherton theory (Komev [30]).
1164

I I I I I I I I I i i I

a)
,90
>-
l--
v/
0
o
**,/- OATA
50
m

>' 10"
Ld
THEORY,
V) rl# + Tleon
"I"
12.
I
(r

k-
Z
w
THEORY,J ' ' " ' " ' ' " -
< rl# ONLY ""-.

I
0.1 5
vn.ocrry (cm/,)

104 I i * m i i i m I I

b)
DATA,

F Xf<I

i--
THEORY,
0 TI. + Tlcon
10 3 q.. ,..

"" " DATA,


I--
-7' "-.... / Xf:l
i.d

THEORY, "-.
1"Is ONLY "-.

10 z 9 , , , , , , . . . . . . . .

0.01 0.1
GAS-PHASE VELOCITY (cm/,)
Figure 7 Apparent foam viscosity in glass bead packs as a function of gas-phase
velocity. Physical conditions: the capillary pressure Pc ~ 1900 dynes cm -~ , the
1165

ordinary gas-phase relative permeability k~ = 0.056. a) The controlled-bubble-size


regime, r b/rca p ~ 3.8, where r b is the effective radius of a foam bubble and Gapis an
equivalent capillary radius, b) The pack-generated bubble-size regime,
r~/r~,, ~ 0.92. The parameter x e = S~/S~ is the ratio of the moving-gas saturation
Se, estimated from the measured residence time of flowing gas, and the gas
saturation, Se. For the pack-generated bubble-size regime, xe = 1; r/, is the
Hirasaki-Lawson viscosity. (After Falls, et al [46]).

Therefore, one of the possible explanations of the experimentally observed


'shear thinning' behavior (Figure 6) lies in a nonlinearity of the friction law as it
might be expected from the Bretherton theory (Falls, et al [46]).
The above mentioned results are applicable only for the following range of
capillary numbers 10-4 < Ca < < 1 (Schwartz, et al [48]; Olbricht [49]) and for
'short' bubble trains. The bubble train is named 'short' when the length of the
sample is on the same order of magnitude as the correlation length
1
N~o, ~ 2~rK ~ ' K# < <1 of a train (Dautov, et al [32]). (The Falls-Musters-

Ratulowski experiments fall into this range). In the opposite case, the effect of
chain elasticity cannot be ignored and the theory must be modified.

4. MOTION OF BUBBLE TRAINS

In order to examine the foam flow mechanisms, consider a less-studied


alternative wave regime caused by oscillations of the gas pressure within the
bubbles (Kornev and Kurdyumov [50]; Kornev [30]). There are at least two areas
of applicability of the wave mechanism of foam transport: acoustic flows, for
which the high frequency modes are important, and transport of the bubble trains
itself. Illustrating these flow regimes, we shall first consider the specific features
of the propagation of the acoustic waves through a perfectly ordered one-
dimensional foam, i.e., through a bubble chain.
The standard starting point in any analysis of a wave picture, lies in a
specification of the corresponding dispersion relations (Kittel [51]). It is
instructive, therefore, to consider such a formula for our case. For the model
expressed by the linearized version of equations (8) and (4), the respective
relation has the following form (we neglect for a time the lamella friction)
1166

= 2 + 2~r2K/~ ' ~ = hK2,' (14)


Pliquid
where co is the frequency, k is the wave number, and h is the lameUa thickness.
This equation implies that there are two critical values for the frequency, below
and above which the acoustic wave will not propagate through the chain. The
lower critical boundary corresponds to the natural frequency of the individual
lamella. This frequency is reached when the wave length tends to infinity (or
k ~ 0), i.e., when the bubble train oscillates as a whole, and when all the
lamellae vibrate at the throats of pore channel m unison. Another critical
frequency selects the range of wave numbers admissible for the theory (when
the wave length becomes comparable with the distance between adjacent
lameUae, the theory fails to describe the triple substrate-gas-lamella interactions).
Thus, the region, transparent to propagation of the acoustic waves, is bounded
from above and below by well-defined values of the frequency. It should be
stressed that the above discussed theory does not confine itself to the pore
channels with small pore aspect ratios. Such a small-amplitude analysis is valid
for all the channels within which foam resides as perfectly ordered.
At a glance, because the screening effect is originated from the Laplacian
blocking action of the lamellae, we should conclude that any perturbation will be
suppressed if its frequency lies outside the admissible gap. It should be repeated
that the above presented analysis relates only to the waves of a small amplitude,
and it cannot be extended directly to a nonlinear case. Considering the stick-slip
motion of lamellae, we should focus on the family of wave-propagating solutions
of the respective nonlinear model, equations (8) and (4). This kind of dynamics
occurs when we initially excite just one lamella, while all the others are initially at
rest.
In its general features, the stick-slip motion of the lamellae resembles the
Frenkel-Kontorova mechanism for the flow of dislocations in crystals (Frenkel'
and Kontorova [35]). The excited lameUa initiates a wave propagating through
the bubble chain in a 'falling dominos' type of motion towards the other end of
the chain, which is then reflected, moving in the opposite direction towards the
initially excited lamella (Figure 8). In such a motion, practically no energy is lost
during the flow, because the surface energy of the lamellae is compensated
almost entirely by the elastic energy of the bubbles.
Again, focusing on the dynamics within a short time interval, we find a soliton
like that of the Frenkel-Kontorova kink (Frenkel' and Kontorova [35]; Lonngren
[52]; Dodd, et al [53]). A scenario of the generation of a soliton is as follows.
1167

Just after the external pressure drop has overcome the threshold G - 4Vr--~

(Komev [30]; Dautov, et al [32]), the domain wall, which separates the stagnation
zones ahead of and behind the front of the wave, depins and runs forward.
Therefore, in soliton dynamics, the competition between the inertial, elastic, and
pinning forces is important, and the viscous forces do not crucially change the
character of the soliton propagation during a short period of time. If a bubble train
moves for a long period of time, the inertia forces are negligibly small with
respect to the three major forces - the elastic, viscous, and capillary forces. These
forces struggle among themselves and drive the displacement waves. Notably, if
the external pressure drop is maintained at a long time interval, the number of the
domain walls changes. In fact, the number of the domain walls is proportional to
the number of the channel periods over which the front domain wall has skipped.
Therefore, the longer is the bubble train, the greater is the number of domain
walls, and, consequently, the greater is the bubble train resistance.

ta il lam e Ila
124 ;.-
122 ~- N=80; I<=1
120 ~- #=10 .3
1.18

1.18
L
1.14.
middle lamella

~ 1.12

1.10
1D8
1.00

1.04
head lamella
102
= 9 ~ I ~, I =, I = ~ ~ r ~ ' , I
1J00
3.g5 3.g8 3.97 3.g8 3.98 4.00

time

Figure 8. Wave mechanism of foam motion. The head and tail lamellae move m
unison, while the middle lamella drifts in antiphase. Numerical experiment based
on equations (8) and (4), e -- 0 (after Musin [56]).
1168

Illustrating a priority of the collective effects in the bubble train friction, we shall
use a somewhat simplified model of lamella interactions. For clarity, a linear
elastic spring will represent the elastic interactions of the lamellae. Moreover,
because each individual domain wall plays a key role in the chain friction, we
shall focus on the motion of a solitary wave. Then, in a continuous limit, equation
(10) is written as (Kornev [30])

c~ p = 0 2 p /~ sin p + q~. (15)


3t ds 2 K

Here q' is a dimensionless external force which represents the total action of the
external pressure gradient onto each lamella in the chain. In the vicinity of the
threshold, the profile of a domain wall can be adequately approximated by the
Frenkel-Kontorova kink as

p = 4 t a n - I exp[-(x- Xo)] +OL~]---~) , X = . (16)

Then, rewriting equation (15) in a wave fixed reference frame, the velocity of the
domain wall is found as the solvability condition to equation (15) (Langer [54];
Kornev [30])

- ooL-d-x)dx = q~ ? dd~xX ,
--o0
(17)

where x - ~ - ~ ( s - vt) and the function p in equation (17) is expressed by equation


(16). Substituting equation (16) into equation (17), we have
qJ

Thus, equation (18) may be treated as an equation of the motion of an individual


effective lamella along an active channel. The friction coefficient of such a
macro-lamella is affected by the binding and pinning energies of the chain. A
dependency like equation (18) has been recently reported by Braiman, et al [55]
in a different comext. Their and Musin's [56] numerical experimems indicate that
1169

the friction coefficient grows with the number of lalnellae in the chain (the
number of moving domain walls) and scales similarly to equation (18).

5. WEAK FOAMS. FLOW OF 'SOLUTIONS' OF BUBBLE CHAINS

As shown in the previous section, during its motion, a bubble train is subdivided
into a system of domain walls. The domain walls separate the regions within
which the lamellae are displaced only slightly so that the principal part of the
dimensionless displacements, p,, remains at an almost the same constant level.
The size of each wall is much greater than the pore size. Therefore, on a
macroscale level of a sample, the train may again be imagined as a one-
dimensional coarse-grained foam in which the domain walls serve as
macrolamellae. Such lamellae interact between themselves via an renormalized
elastic potential. The potential incorporates all the microlevel elastic and pinning
interactions and displays the collective properties of the native chain (Lonngren
and Scott [52]; Dodd, et al [53]) . It is important that the distance between
neighboring effective lamellae crucially depends upon the history of loading. For
instance, if a piston is applied to the tail end of the chain, and if such a load is
maintained during a long period of time which is sufficient to shift the head
lamella of the chain, then the number of domain walls will be a well-def'med
constant. This constant is a universal characteristic of the given bubble train, the
pore channel and the applied pressure drop. The resulting structure of the chain
on the macroscale level of the sample will be periodic, i.e., the domain walls will
form a new perfect lattice. But if it is granted that the load is continuously
redistributed over the train (the situation which is most likely to occur in a natural
porous medium), then the chain acquires, generally speaking, a nonregular
structure: in the coarse-grained foam, the cell size will alter from one cell to
another.
A viewpoint which designates the bubble chain as a macroobject enables one to
construct a hydrodynamic theory for a so-called weak foam. For the latter, there
is a very small fraction of trapped bubbles, so that the gas and a system of bubble
trains flow together through a porous medium. The gas mobility is reduced due to
an enhanced friction of the bubble trains. We encounter a similar situation in the
hydrodynamics of polymer solutions (Doi and Edwards [57]). The action of
bubble trains on a free gas, which is flowing through a porous medium, resembles
the action of polymer molecules on a flowing solvent. Because the bubble train
resistance is accumulated almost entirely within the domain walls, and, at the
same time, each domain wall is much greater than a pore size, the hydrodynamics
of a weak foam may be considered as the hydrodynamics of an ensemble of the
1170

bubble chains 'dissolved' into a Darcian fluid, provided that the free gas flow
obeys the Darcy law ( i.e., the relation between the pressure gradient and the gas
velocity is linear).
In a non-uniform flow, where each bubble train undergoes the action of an
extemal pressure gradient which is imposed by the carrying free gas, any train
will be stretched or contracted (because each macrolamella in the chain acquires
a velocity imposed by the kinematics conditions). But, due to an inherent
elasticity, the chain will tend to restore its equilibrium length. By restoring its
equilibrium state, the chain draws into motion the surrounding gas and,
consequently, other chains. Therefore, the chains which are stretched only slightly
will flow to the regions in which the chains are more extended. As a result, a net
restoring force appears. A schematic picture of the phenomenon is presented in
Figure 9, where a dumbbell plays a role of a foam filled macrobubble or a bubble
train.

/crv-rrr O

2-rcrZ- "-

Figure 9. The sketch elucidating the mechanism of creation of a restoring force in


a gradient flow.

The above discussed scenario of foam flow can be theoretically described by


making use of equation (18) as a basic formula of a microscopic theory. Then,
phenomenologically expressing qJ in the formula (18) through the pressure
gradient, the flux of a foam within the sample can be written as (Kornev and
Kurdyumov [50]; Kornev [30])

t
J = -cfl jds < u(s,t)u(s,t) > - V O , (19)
0
1171

here c is the concentration of the bubble trains, 1 is the length of the train, and p
is a phenomenological constant. In equation (19), the vector u(s) denotes a vector
which joins two adjacent macrolamellae, and arclength s is measured along the
train. The angle brackets denote an average over the orientations of the
macrobubbles which link the macrolamellae.
It is convenient to rewrite equation (19) in the following form

/,
J = J o - cp jd,sS. V P , (20)
0
Jo = - c - ~ - V P ,
1
S= < u(s,t)u(s,t)-;l >, I ij =6ij , S ' u = S i j u j , S : u = S i j u i u j ,

where tensor S is the tensor of the order parameter. One can see that the first
term on the right hand side of equation (20) may be included into the ordinary
Darcy law associated with the flow of flee gas. Thus, equation (20) is written in a
form in which it is convenient to treat the foam motion in a porous medium as a
flow of a "solution' of bubble chains (Doi and Edwards [57]).
For such a hypothetical system, the detailed derivation of the constitutive
equations has been presented by Kornev and Kurdyumov [50]. The system of
these equations contains a generalized Darcy law for gas flow in the presence of a
foam

v = - k f VP -L, L =J- Jo ' (21)

and a kinetic equation which describes an evolution of the "blocking force' L due
to the action of the velocity gradients. In a one-dimensional case, e.g., for a foam
motion in a porous tube, the model has the following form

Ov
(22)
Ox
DL DL L
~ - l - V . . . . (23)
Dt Dx r
3t:'
v =-k ~-L . (24)
f 3x

Here v is the gas velocity, v is a relaxation time, and k~ is an effective seepage


coefficient. The system of equations (22)-(24) may also be easily elucidated in
1172

the terms of the so-called lamella "break-and-reform' mechanism of foam motion


(Holm [58]). Namely, equation (24) expresses the balance of forces for the
moving foam. The viscous force on the left hand side of the equation is balanced
by the pressure gradient and a blocking force. The latter arises by vim~e of a
blockage of the gas channels by 'valve' lamellae. In the first approximation, the
blocking force is directly proportional to the number of 'valve' lamellae. We thus
can express the lamella density through the function L and, ignoring the
generation rate of lamellae, we may treat equation (23) as a kinetic equation for
the lamella population balance in a moving foam. The parameter r should thus be
treated as a lifetime of a 'valve' lamella, and it may be represented as

1 1 1
r rt rh

where r, is a 'thermodynamic' lifetime of a lamella, and r, is its hydrodynamic


traveling time. So, each 'valve' lamella blocks a gas path until it rips due to
inherent thermodynamic or hydrodynamic instability.
Within the framework of the above formulated bubble-train-flow-mechanism,
the relaxation time can be obtained by analyzing a random walk of a bubble train
as a whole. To describe this random walk, the following h y p o t h e s i s has been
used: once formed, the system of active channels cannot be destroyed. Then the
picture of a random walk resembles a reptation of a polymer molecule through the
obstacles (de Gennes [59]; Doi and Edwards [57]; Kornev and Kurdyumov [50]).
A bubble train can change its active channel only by moving through a network of
active channels in a worm-like manner. In fact, only its ends participate in
movement, but the rest of the macrolamellae are effectively trapped within the
existing (at the given moment) active channel.
The model in equations (22)-(24) reproduces the main features of the flow of a
foam through a porous medium (Figures 10-11). The one-dimensional flow is
governed by the parameters r and b = Lok I r ! H , where Lo is a boundary value
of the blocking force L, and H is the length of the sample. These parameters can
be extracted from the experimental data by using a piecewise-linear
approximation (Figure 10) as follows

AP
kf H = ( l + b ) v , v --->0, (25)
AP H
kf H =v+ b, v ~o (26)
1173

where AP is the applied pressure drop. The point at which the straight lines
intersect has the coordinates

vr/H =1, APrkf I H 2 = l + b ( 1 - e - 1 ) = l + b . (27)

W
5~

11 2/ 31 4

0 2 4 6 8 10 12 14 16
AP
Figure 10. The velocity as a function of the pressure drop. 1-b--1, 2 - b = 3,
3-b=5, 4-b=10.

The coordinates of the imersection poim can be used as the fitting parameters.
An additional parameter needed to specify the model is the breakthrough time
T* = t*r . At this instant, a from of the foam first reaches the exiting end of the

sample. The dimensionless parameter t* can be found from the following


transcendental equation
1174

1= f t b 1 - exp(-(1 + b)t*)
H 2 1 + b + 1 +-~" 1+ b " (28)

The system of equations (25)-(28) is closed so that the parameters b , r and ke


can be found in each individual experiment. Then, these parameters can be
plotted as functions of the degree of water saturation, properties of the foaming
agent, etc. Despite the fact that the model contains a small number of the physical
constants, the theory cannot be directly spread onto the two- or three-dimensional
flows (because the parameter b depends upon the boundary value of the order
parameter S). The similar problem occurs in the theory of polymer viscoelasticity
(Bird, et al [60]) and also remains unresolved.

10-

t*
u

1 " II 1 l b
0 2 4 6 8 1
AP
Figure 11. The breakthrough time as a function of the pressure drop. 1 - b = 1,
2-b=3, 3-b-5, 4-b=10.
1175

6. STRONG FOAMS

As follows from the previous sections, the main feature of foam rheology
concems its 'pseudoplasticity'. The foam pseudoplasticity is very sensitive to the
foam texture inside a porous medium and to a fraction of the pores blocked by
lamellae. The physical reasons of the blocking action of weak and strong foams
are different. For a weak foam, even in the absence of a start-up pressure
gradient, an enhanced friction of bubble chains will result in an apparent
pseudoplasticity of the gas-foam system. But for a strong foam, when the bubble
trains flow through a small fraction of the pore channels, the blocking effect is
originated from at least two mechanisms. The first is caused by trapped lamellae
which do not participate in the movement. Such a blockage leads to a
permeability that is orders of magnitude lower than that for the single phase flow.
The second contribution to a reduction of the gas flux in the presence of a strong
foam concerns an enhanced friction of the bubble trains. Thus, the total gas
mobility, i.e., the coefficient in the generalized Darcy law, is the product of
permeability divided by apparent viscosity. (If the liquid also participates in the
movement so that its saturation changes during the process and alters the fraction
of the trapped pores, then, instead of the absolute permeability, the relative
permeability to gas has to be considered). In addition to a nonlinearity of the gas
mobility, a start-up pressure gradient plays the key role in the treatment of the
hydrodynamic effects in strong foams. This is the pressure gradient which is
required to depin some bubble trains and to create a network of active channels
(Cottrell [61]; Read [62]; Hirth and Lothe [63]; Suzuki, et al [64], Komev [30],
Dautov, et al [32]).
Thus, the term plasticity, in its own physical meaning, must be attributed to
strong foams. Both, the bubble trains and the corresponding network of active
channels play the same role as the one prescribed to the dislocations and the
network of dislocations in solid state physics. In fact, studies have shown that the
trapped gas saturation of strong foams in porous media can be as high as 80%
(Rossen [14]). Therefore, the trapped foam forms an elastic field around the
network of active channels. In a short period of time, the trapped foam resides as
effectively motionless lamellae. However, the diffusion processes need to be
considered to characterize the long term behavior of a foam (Falls, et al [65];
Cohen, et al [66]). In particular, a mechanism like the Nabarro-Herring-Lifshitz
mechanism [67-69] of a diffusion-induced plasticity might play an important role
in the motion of trapped foam. Indeed, under a pressure gradient, the gas will
diffuse through the lamellae so that the gas pressure within the bubbles will
change with time. Because of a difference in the Laplacian and gas pressures, the
1176

trapped lamellae will creep as well. A sophisticated analysis of this effect is


highly desirable.
There is an important evidence of self-organized criticality in foam flows
through porous media. Namely, the typical manifold reduction of the gas mobility
speaks in favor of the fact that, during the flow, the network of the active
channels remains nearby in the same state, as it would be at the percolation
threshold (Rossen and Gauglitz [70]). Scaling estimates of the gas mobility
reduction by foams have been obtained by operating the ordinary percolation
theory (de Gennes [71]; Rossen and Gauglitz [70]; Entov and Musin [72]).
However, the critical behavior of the network of active channels has been
assumed ad-hoc. Though most researches believe that the transport phenomena in
strong foams obey the percolation laws, the physics of the hydrodynamic
processes remains unclear and puzzling.

7. FOAMS IN FIBER SYSTEMS

Foams have been found various applications in industrial technologies dealing


with fiber structures and fibrous materials, such as dyeing, printing, mercerizing,
and finishing of textile fabrics, paper coating, resin-impregnation of fibrous mats
and fabrics [73-78]. In these processes, the usage of foams instead of bulk
liquids, as a vehicle for delivering small amounts of liquid solutions to fiber
surfaces, leads to substantial energy savings because of small amounts of
residual solvent to be removed at the drying step. Another class of processes,
involving bubble generation in fiber structures, is fabricating fiber reinforced
composites. Herewith, bubble formation during the stage of liquid resin
impregnation causes a negative effect of non-uniform polymer distribution in the
products (Judd and Wright [79]), and the mechanisms of air entrapment and
bubble interactions within a fiber network have been studied aiming to diminish
this phenomenon at technological conditions (Mahale, et al [80, 81 ]).
Interactions of bubbles with fibers and fiber networks have certain specifics
compared with capillaries and pore networks in solid-wall materials. The major
difference is that in contract with granular materials, where the sizes of pores and
grains are commensurable, the typical diameter of fibers, which constitute a
skeleton of a pore structure, is commonly smaller than the typical diameter of
voids/pores between the fibers. In fibrous materials, it is not easy to identify
single pores, their shapes and dimensions. The definition of pore sizes and their
distribution in fibrous materials is a matter of convention. The most rational way
to introduce the pore dimensions in a real fiber system is to consider a model
system of "effective" pores, in which some characteristic processes would occur
1177

obeying the same peculiarities as in the fiber system under consideration. For
example, the effective pore sizes are estimated from the experiments with
capillary equilibrium of immissible fluids (commonly, wetting liquid and gas) and
with steady or quasi-steady forced flow of a non-wetting fluid (commonly, gas)
within and/or through a fibrous sample. An advanced technique of liquid
porosimetry has been developed by Miller and Tyomkin [82] for determining
pore volume distributions in fiber systems and other materials. The method is
based on the consecutive, quasi-equilibrium wetting fluid - gas displacement
under precisely controlled pressure at isothermal environment. Therewith, the
effective radius of a pore, where the liquid and gas phases coexist at given
pressure, is operationally related to the mean radius of curvature of the
equilibrium meniscus between the phases through the Laplace equation.
During the process of gas-liquid displacement in fibrous materials, bubble
formation occurs mostly due to a hydrodynamic instability of the wetting film
clinging to the fibers. This is the common mechanism of lamella formation in any
porous media [27]. However, the fiber structure causes some peculiarities. It is
likely that in case of most of fibrous materials, we deal with strong foams with
the bubbles commensurate to the pores. This conclusion follows from the
experiments of Gido, et al [83], who examined the flow of foams through fibrous
mats by characterizing the bubble sizes before and after injection. The authors
observed that the size of the output bubbles exiting the fiber system was
independent of the bubble size of the input bulk foam. For fibrous mats with
different pore structure, the output bubble sizes were found correlated with the
pore size distributions measured by the liquid porosimetry [82]. This result can be
interpreted assuming that the foam bubbles, residing within the fiber network
during the foam flow through this network, are grouped basically into two
configurations: a system of immobilized bubbles strongly pinned to fibers and
system of unpinned bubbles, which are formed in the bubble trains sliding along
the active channels confined by the immobilized bubbles. The bubbles are
immobilized when they are transpired by several fibers which intersect or are not
collinear. These bubbles are crucified at fiber crossing and/or stretched by
differently oriented fibers. The mobile bubbles are commensurate with the pore
constrictions in order to pass through them without essential deformation. The
movement of lamellae in the active channels within a fibrous structure should be
like the lamella stick-slip motion within porous solids described in detail above.
However, foam flow through a fiber mat is more difficult to formalize than the
flow in a solid-wall porous body. At present, no quantitative approach exists to
describe foam flow in fibrous materials. This problem is still awaiting its solution.
1178

While considering equilibrium distribution of bubbles and foam flow in fiber


networks, we have to account for a specific behavior of wetting films on fiber
surfaces. In particular, the lamellae (bubble films) should coexist with the wetting
films and the Plateau borders at the intersections of lamellae and/or lamellae and
wetting films. The equilibrium configurations of such a complex system are
determined by capillary and surface forces (the latter is expressed via an
additional so-called Derjaguin's disjoining pressure [84], [15], [16]). The fiber
surfaces are commonly convex. This means that the capillary pressure acting on
the liquid-gas interface of the wetting film covering the fiber is positive and tends
to squeeze liquid out of film into the regions of fiber crossings. This tendency
leads to a reduced mobility of wetting films in fiber systems compared with
ordinary capillary systems. These effects have been considered in literature as
related to the liquid spreading and drop residence on fibers first by Carroll [85,
86], who accounted for the capillary forces only, and then in great detail by
French researchers from the de Gennes group [87-91 ]. Brochard was the first to
emphasize a central role of long-range intennolectdar forces in residence,
stability, and spreading of films and drops on fiber surfaces [87, 88]. Di Meglio
[89] experimentally observed wetting films stabilized by the Van der Wa,~s
forces, and proved that mass transfer between the drops residing on a fiber occurs
through these films. Similar effects should be important in phenomena involving
the bubbles on fibers and in fiber networks, however their description is lacking
in literature. A theory of foams in fiber systems cannot be advanced without a
solution of a chain of particular problems: equilibrium shape of a bubble residing
on a fiber, transition zone between the bubble lamella and wetting film coating
the fiber, slippage of a bubble along a fiber, bubble crucifixion at a fiber
crossing, motion of a system of contacting bubbles transpired by a fiber ("bubbles
on a spit"), etc.

8. CONCLUSIONS

The motion of foams through porous media is a challenging problem in physico-


chemical hydrodynamics. The basic mechanisms of foam transport reviewed in
this article contain some, but certainly not all, of the relevant physics of foam
flow in porous media. Foam flow in porous media is a multifaceted process in
which, on one hand, foam texture strongly governs foam rheology, and on the
other hand, foam texture is in turn regulated by the porous medium through the
capillary pressure. We have analyzed the main features of this process on
examples of foam motion in model pore channels. The modem theories of the
foam lamella transport in pore channels of varying cross-section and the models
1179

of the "weak foam" flow are discussed in detail. Careful analyses of the flow on
the scale of individual pores or channels are useful in exposing effects of various
physical parameters on foam motion and in identifying flow-induced patterns. In
addition, the basic physical mechanisms of foam microhydrodynamics tmderlie a
variety of technological processes in oil recovery, groundwater/soil remediation,
textile manufacturing, etc.

REFERENCES

1. C.V. Boys, Soap Bubbles and the Forces Which Mould them. Soc.for
Promoting Christian Knowledge, E. and J.B.Yotmg, London,1890.Reprinted
in Doubleday Anchor Books, New York, 1959.
2. K.J. Mysels, K. Shinoda and S. Frankel, Soap Films, Studies of their Thinning
and a Bibliography, Pergamon Press, New York, 1959.
3. J.J. Bikerman, Foams, Springer-Verlag, New York, 1973.
4. I.B. Ivanov (ed.), Thin Liquid Films: Fundamentals and Applications, Marcel
Dekker, New York, 1988.
5. A.M. Kraynik, Ann.Rev.Fluid Mech., 20 (1988) 325.
6. A. Wilson (ed.), Foams: Physics, Chemistry and Structure, Springer-Vedag,
New York, 1989.
7. P.M. Kruglyakov and D.R. Exerowa, Foam and Foam Films, Khimia,
Moscow, 1990.
8. J. Stavans, Rep.Prog.Phys., 56 (1993) 733.
9. R.K. Prud'homme and S.A. Khan (eds.), Foams:Fundamentals and
Applications, Marcel Dekker, New York, 1995.
10. S.H. Raza, Soc.Petr.Eng.J., 10 (1970) 328.
11. S.S. Marsden, Foams in Porous Media - SUPRI TR-49, US DOE, 1986.
12. J.P. Heller and M.S. K u n t a m ~ l a , Ind.Eng.Chem.Res., 26 (1987) 318.
13. L.L. Schramm (ed.), Foams: Fundamentals and Applications in the Petroleum
Industry, Advances in Chemistry Series 242, 1994.
14. W.R. Rossen, in [9], p.413.
15. L.I. Kheifetz and A.V.Neimark, Multiphase processes in porous media.
Khimia, Moscow, 1982.
16. B.V. Derjaguin and N.V. Churaev and V.M. Muller, Surface Forces, Nauka,
Moscow, Nauka, 1985; Surface Forces, Consultants Bureau, New York,
1987.
17. A.V. Neimark and M. Vignes-Adler, Phys. Rev. E, 51 (1995) 788.
18. P.G. de Gennes, Rev.Mod.Phys., 57 (1985) 827.
1180

19. A.H. Falls, G.J. Hirasaki, T.W. Patzek, P.A. Gauglitz, D.D. Miller and
Y.Ratulowski, SPE Res. Eng., 3 (1988) 884.
20. A.N. Fried, The Foam-Drive Process for Increasing the Recovery of Oil,
Report US Bureau of Mines R.I.5866 (1961).
21. G.J. Hirasaki and J. Lawson, Soc.Petr.Eng.J., 25 (1985) 176.
22. C.W. Nutt and R.W. Burley, in [6], p. 105.
23. A.V. Bazilevsky, K.Komev and A. Rozl~ov, in Proceedings of the ASME
Symposium on Rheology & Fluid Mechanics of Nonlinear Materials, Atlanta,
November 17-22, 1996.
24. O.S. Owete and V.E. Brigham, SPE Res.Eng., 2 (1987) 315.
25. K.T.Chambers and C.J. Radke, in N.Morrow (ed.), Interfacial Phenomena in
Petroleum Recovery, Marcel Dekker, New York, 1990, p. 191.
26. R.A. Ettinger and C.J. Radke, SPE Res.Eng., 7 (1992) 83.
27. A.R. Kovscek and C.J. Radke, in [ 13], p. 113.
28. E. Rabinowicz, Friction and Wear of Materials, Wiley, New York., 1965.
29. F.P. Bowden and D.T. Tabor, Friction and Lubrication of Solids, Claredon
Press, Oxford, 1986.
30. K.G. Kornev, JETP, 80 (1995) 1049.
31. K.G. Kornev, V. Mourzenko, and R. Dautov, in E.P. Zhidkov (ed.),
Proceedings of the Conference on Computational Modelling and Computing
in Physics, JINR, Dubna, September 16-21, 1996.
32. R. Dautov, K. Kornev, and V. Mourzenko, To appear in Phys.Rev. E.
33. V.L. Pokrovsky and A.L. Talapov, Zh.Eksp.Teor.Fiz., 78 (1980) 269.
34. P. Bak, Rep.Prog.Phys., 45 (1982) 587.
35. I. Frenkel' and T.A.Kontorova, Zh.Eksp.Teor.Fiz., 8 (1938) 1340.
36. B.D. Josephson, Adv. in Phys., 14 (1965) 419.
37. I.O. Kulik, Zh.Eksp. Teor.Fiz., 51 (1966) 1952.
38. A. Seeger and P. Schiller, in W.P.Mason and R.N.Thurston (eds.), Physical
Acoustics, Academic, New York, 1966, Vol. IliA, p.361.
39. A. Barone and G. Paterno, Physics and Applications of the Josephson Effect,
Wiley, 1982.
40. A.J. Lichtenberg and M.A. Lieberman, Regular and Chaotic Dynamics, 2end
ed., Springer-Verlag, New York, 1992.
41. L.D. Landau and E.M. Lifshitz, Mechanics, Nauka, Moscow, 1965.
42. G.M. Zaslavsky and R.Z. Sagdeev, An Introduction to Nonlinear Physics:
From Pendulum to Turbulence and Chaos, Nauka, Moscow, 1988.
43. L.G. Aslamasov and A.I. Larkin, Zh.Eksp.Teor.Fiz.Pis'ma, 9 (1969) 150.
44. H.A. Barnes, J.F. HuRon, and K. Waiters, An Introduction to Rheology.
Second ed., Elsevier, 1993.
1181

45. F.P. Bretherton, J.Fluid Mech., 10 (1961) 166.


46. A.H. Falls, J.J. Musters, and J. Ratulowski, SPE Res.Eng., 4 (1989) 55.
47. J.E. Hanssen and M. Dalland, in [ 13], 319.
48. L.W. Schwartz, H.M. Princen, and A.D. Kiss, J.Fluid Mech., 172 (1986) 259
49. W.L. Olbricht, Ann.Rev.Fluid Mech., 28 (1996) 187.
50. K.G. Komev and V.N. Kurdyumov, JETP, 79 (1994) 252.
51. C. Kittel, Introduction to Solid State Physics, Wiley, New York, 1956.
52. K. Lonngren and A. Scott, Solitons in Action, Academic Press. N.Y., 1978.
53. R.K. Dodd, J.C. Eilbeck, J.D. Gibbon, and H.C. Morris, Solitons and
Nonlinear Wave Equations, Academic Press, London, 1984.
54. J.S. Langer, in G. Grinstein and G. Mazenko (eds.), Directions in Condensed
Matter Physics, World Scientific, Singapore, 1986, p. 165.
55. Y. Braiman, F. Family, and H.G.E. Hentschel, Phys.Rev.E., 53 (1996)
R3005.
56. R. Musin, Ph.D. Thesis, Moscow, 1997.
57. M. Doi and S.F. Edwards, The Theory of Polymer Dynamics. Claredon Press,
Oxford, 1986.
58. L.W. Holm, Soc.Petr.Eng.J.,.8 (1968) 359.
59. P.G. de Gennes, J.Chem.Phys., 55 (1971) 572.
60. R.B. Bird, R.C. Armstrong, O. Hassager, and C.F. Curtis, Dynamics of
Polymeric Liquids, Wiley, New York, 1977, Vols. 1, 2.
61. A.H. Cottrell, Dislocations and Plastic Flow in Crystals. Oxford University
Press, London, 1953
62. W.T. Read, Dislocations in Crystals, McGraw-Hill, New York, 1953.
63. J.P. Hirth. and J. Lothe, Theory of Dislocations, Wiley, New York, 1968.
64. T. Suzuki, H. Yoshinaga. and S. Takeuchi, Dynamics of Dislocations and
Plasticity. Mir, Moscow, 1989.
65. A.H. Falls, J.B. Lawson, and G.J. Hirasaki, JPT, Jan (1988) 95.
66. D. Cohen, T.W. Patzek, and C.J. Radke, J.Colloid Interface Sci., 179 (1996)
357.
67. R.N. Nabarro. Report of the Conference of the Strength of Solids,
Phys.Soc.London, London, 1948, p. 75.
68. C.J. Heri~g, J.Appl.Phys., 21 (1950) 5.
69. I.M. Lifshitz, Zh.Eksp.Teor.Fiz., 44 (1963) 1349; The Collected Works of
II'ya Lifshitz, Nauka, Moscow,1987. Vol.1.
70. W.R. Rossen and P.A. Gauglitz, AIChE J., 36 (1990) 1176.
71. P. G. de Gennes, Revue De L'lnstitut Francais Du Petrole, 47 (1992) 249.
72. V.M. Entov and R.M. Musin, Preprint IPM RAN, Moscow, No 560 (1996).
1182

73. T.F. Cooke and D.E. Hirt, in Foams, ed. R.K. Prud'homme and S.A. Khan,
Marcel Dekker, NY, 1995, p.339.
74. C.C. Namboodfi and M.W. Duke, Textile Res. J., 49 (1979) 156.
75. Gregorian, R.C., Text. Chem. Color., 19 (1987) 13.
76. D.E. Hirt, R.K. Prud'homme, L. Rebenfeld, AIChE J., 34 (1988) 326.
77. D.E.Hirt, R.K. Prud'homme, L. Rebenfeld, Textile Research J., 61 (1991) 47.
78. E.L. Wright, Textile Research J., 51(1981) 251.
79. N.C. Judd and W.W. Wright, SAMPE, 14 (1978) 10.
80. A.D. Mahale, R.K. Prud'homme, and L. Rebenfeld, Polymer Eng. & Sci., 32
(1992)319.
81. A.D. Mahale, R.K. Prud'homme, and L. Rebenfeld, 4 (1993) 199.
82. B. Miller and I. Tyomkin, J. Colloid & Interface Sci., 162 (1994) 163.
83. S.P. Gido, D.E. Hirt, S.M. Montgomery, R.K. Prud'homme, and L.
Rebenfeld, J. Dispersion Sci. & Tech., 10 (1989) 785.
84. B.V. Derjagum, Kolloid Zeits, 69 (1934) 155.
85. B.J. Carroll, J. Colloid & Interface Sci., 57 (1976) 488.
86. B.J. Carroll, Langmuir, 2 (1986) 248.
87. F. Brochard, J. Chem. Phys., 84 (1986)4664.
88. F. Brochard-Wyart, C.R. Acad. Sc. Paris, Serie II, 303 (1986) 1077.
89. J.-M. di Meglio, C.R. Acad. Sc. Paris, Serie II, 303 (1986)437.
90. F. Brochard-Wyart, J.-M. di Meglio, and D. Quere, C.R. Acad. Sc. Paris,
Serie II, 304 (1987) 553.
91. D. Quere, J.-M. di Meglio, and F. Brochard-Wyart, Revue Phys. Appl., 23
(1988) 1023.
1183

FLOW OF NON-NEWTONIAN FLUIDS IN POROUS MEDIA

Shapour Vossoughi

University of Kansas
Department of Chemical and Petroleum Engineering
Lawrence, KS., USA

1. INTRODUCTION

Non-Newtonian fluid flow through porous media has become increasingly


important in a wide range of disciplines and industrial segments. This is the result
of availability of a wide variety of polymers that have interesting fluid flow
properties, and there is a growing demand for their industrial use. Catalytic
polymerization process, the injection of polymer and surfactant solutions into
petroleum reservoirs to enhance oil recovery, food processing, and fluid flow
through riving tissues are examples of the vitality of understanding the non-
Newtonian fluid flow through porous structures.
This section deals with the different aspects of non-Newtonian fluid flow through
porous media and will bring together the different treatments of the subject matter
commonly practiced in different disciplines. It will cover the complexity of both
non-Newtonian fluid flow behavior and flow through porous media. Anomalous
behavior of non-Newtonian fluid flow through porous media could be due to the
fluid, the nature of porous media, or the interaction of fluid and porous media.
Therefore, flow study should be carefully designed to distinguish between these
effects or at least to acquire knowledge of which aspect of the non-Newtonian fluid
flow is being studied.
In this chapter, the nature of non-Newtonian fluids that are commonly employed
will be studied, followed by a look at the nature of the idealized porous media and
the geometrical complexity of true porous media. Fluid and porous media
interaction, such as adsorption, mechanical entrapment, and inaccessible pore
volumes, will have direct effect on the flow and will be analyzed and quantified.
Microscopic and macroscopic view of the flow will be studied next; and, finally, the
predictive models presently available for the study of the non-Newtonian fluid flow
through porous media will be investigated. This will cover models based on
hydraulic radius concept, friction factor/Reynolds number relationship, and
empirical methods.
1184

2. NATURE OF FLUIDS

Non-Newtonian fluid flow through porous media is not limited to the polymer
solutions. There are a wide variety of non-Newtonian fluids that could be of
interest. Following are samples that have already established their association with
porous media.

2.1 Polymer Melts and Polymer Solutions


Polymers come in a variety of forms with respect to their average molecular size
and the geometrical shape of their molecules. This leads to a wide variety of
rheological behavior termed as "non-Newtonian". To illustrate the variety of
viscoelastic responses, Ferry [1 ] sampled polymer from seven different groups; four
of them were uncross-linked polymers and the other three were cross-linked.
Among the uncross-linked ones, he picked up amorphous polymers of low and high
molecular weight, amorphous polymers of high molecular weight with long side
groups, and amorphous polymers of high molecular weight below its glass transition
temperature. Among the cross-linked polymers, he selected a lightly cross-linked
amorphous polymer, a dilute cross-linked gel, and a highly crystalline polymer. He
showed that each of these seven structural types has a characteristic viscoelastic
behavior. He then concluded that there is a strong correlation between the
viscoelastic behavior and the molecular structure of the polymer.
A macromolecule chain in a solution is capable of assuming a variety of
configurations by rotating around its chemical bonds. This makes the polymer
solutions behave differently from their parent polymer gels. The polymer solutions
might behave significantly different depending on their level of polymer
concentrations, and interaction between polymer molecules increases as polymer
concentration increases. The effect of the polymer molecules entanglement on the
solution viscosity may become highly significant in the case of a concentrated
polymer solution. The polymer solution viscosity becomes an increasingly nonlinear
function of the polymer concentration for the concentrated polymer solutions.
Injection of polymer solution into oil reservoirs to enhance oil recovery has
become a common practice. Oil recovery during the primary stage is due to the
depletion of the initial energy stored in the reservoir. Over seventy percent of the
oil is still trapped in the rock pores after the primary recovery is depleted. During
the secondary recovery stage, a fluid, such as water or gas, is injected into the
reservoir to sweep the oil out and push the oil bank toward the production well.
Waterflooding is a common secondary oil recovery technique [2]. Water, because
of its low viscosity, tends to finger through the oil zone and creates early water
breakthrough, but when a small amount of polymer is added to the water to increase
its viscosity it creates a more stable front. Synthetic polymer, such as partially
hydrolyzed polyacryalmide, and biopolymer, such as Xanthan, are the two types of
1185

polymers most commonly used for this purpose. Understanding the flow of aqueous
polymer solution through porous media is essential for producing a reliable model
for reservoir simulation.

2.2 Boger Fluids


Constant viscosity fluids that are highly viscous and highly elastic were first
introduced by Boger [3], hence, they are called Boger fluids. The first report of the
fluids which exhibited elastic properties but remained Newtonian in viscous
behavior was made by Giesekus [4]. The Boger fluid is prepared by dissolving a
small amount of polymer in a highly viscous solvent. These fluids can be divided
into two general groups; aqueous solutions consisting of a small amount of
polyacrylamide dissolved in corn syrup [5,6]; and an organic-based Boger fluid
consisting of a small amount of polyisobutylene dissolved in a mixture of kerosene
oil and polybutene [7-9], or high molecular weight polystyrene dissolved in a
solvent composed of low molecular weight polystyrene in dioctylphthalate [9]. The
two classes of Boger fluids seem to have some differences in their rheological
behavior. For example, Kemielewski, et al. [10], observed significant differences
in the drag ratio measured for the two classes of Boger fluids over the same
Weissenberg number range. The general reported characteristics of these fluids are;
high viscosity, approximately non-shear thinning; high relaxation time; and
optically clear. The Boger fluids are ideal model fluids for studying viscoelastic
behavior in the absence of shear thinning.

2.3 Micro and Macro Emulsions


Micro and macro emulsions are mixtures of two immiscible liquids in the form of
small droplets of one phase into the other. The size of the droplets in the
microemulsion solution is much smaller than those in the macroemulsion.
Macroemulsions are turbid and thermodynamically unstable. The two phases will
eventually separate into the original two immiscible liquids.On the other hand,
microemulsions are translucent and thermodynamically stable. These fluids have
been known for many years, and a wealth of literature is available on their
properties and on their production techniques [ 11-13].
Micro and macro emulsions and their flow through porous media are frequently
encountered in the oil industry. The produced crude oil is often in the form of
emulsion with water, and the emulsion is broken to separate the oil from the water
before it is shipped. Therefore, the flow of macroemulsion through porous media
is an important aspect of the fluid flow near the production well. On the other hand,
injection of microemulsion into the oil reservoir has been practiced to enhance oil
recovery [ 14]. Microemulsions, sometimes called micellar solutions, are formulated
for a specific crude-oil/reservoir-brine system to achieve ultra low interfacial
tension of less than 10.3 dynes/cm. These microemulsions consist of hydrocarbons,
1186

water, surfactant, cosurfactant (such as alcohol or another surfactant), and


electrolytes [ 15]. In many cases, polymer is also added to increase the viscosity of
the micellar solution, since higher viscosity is desired to achieve a stable front and
minimize viscous fingering effect.
Micellar solutions are generally injected in various slug sizes to economize the
chemical flood process. The micellar slug is then chased by a polymer solution to
maintain its mobility control. Loss of surfactant due to the adsorption to the rock
surface and dispersion at the front and at the back of the slug are the main reasons
for the eventual breakdown of the slug effectiveness to mobilize the residual oil.
Microemulsion can also be generated in situ as observed in alkaline flooding of oil
reservoirs. It has been established [16-18] that some components of crude oil, such
as organic acids, asphaltenes, and resins, react with alkaline solutions and form
micellar-type structures.
Rheological properties of an emulsion depend strongly on its composition. Its
viscosity could change an order of magnitude in a narrow range of its concentration
change. This is believed to be caused by the structural change of the microemulsion.
The presence of electrolytes, such as salt, enhances the non-Newtonian behavior of
the emulsions, and the effect of polymer addition on the viscosity of the
microemulsions can be quite significant [19]. This enhanced viscosity is shear
sensitive, and the viscosity recovery after the removal of the high shear is extremely
slow. In general, suspensions and emulsions do not exhibit the same level of
viscoelastic behavior as polymer solutions and polymer melts. However, emulsions
of gel-like structure may exhibit marked viscoelastic behavior [13].

2.4 Suspensions
Suspensions, sometimes called slurries, differ from emulsions in that one of the
two phases is solid. The dispersed phase, which is a solid, is finely ground and
mixed into a liquid. If the dispersion lasts, the mixture is called a suspension.
Surfactants are usually added to make the suspensions more stable by preventing
agglomeration of the solid particles. Suspensions at high volume concentrations are
affected by many factors, such as hydrodynamic interaction between particles,
doublet and high-order agglomerate formations, ultimately mechanical interference,
and surface chemical effects between the particles as packed bed concentrations are
approached. There is a critical concentration, after which, the viscosity increases
sharply [20,21]. The abrupt increase in the viscosity levels may be caused by the
strong inter-particle forces between the solid particles.
Vossoughi and AI-Husaini [22] studied rheological behavior of the coal, oil, water
slurries, and the effect of polymer as an additive. All the systems they studied were
pseudoplastic and showed shear thinning behavior. Slurries made of finer coal
particles were more viscous than those made of coarser grains. They studied the
dynamic properties of coal slurries with and without polymer, using a small
1187

amplitude oscillatory test. In general, for all slurries tested, the dynamic viscosity
was observed to be a decreasing function with frequency, and the modulus of
rigidity was found to be an increasing function with frequency. Their study was
limited to the frequency range of 0.24 to 6.0 cps.
Flow of suspensions and slurries through porous media involves many challenges
unique to this class of non-Newtonian fluids, and they need to be met. For example,
in the filtration of suspensions or drilling mud infiltration near the wellbore in the
petroleum drilling industry.

2.5 Gels
Placement of gel in petroleum reservoirs to improve oil recovery has become an
accepted practice. The technology is known as permeability modification or profile
modification. Water, or any other fluids that are injected into the reservoir to
displace oil, tends to pass through the more permeable zones leaving behind a
significant amount of oil in the reservoir. Gelling solutions are injected following
the waterflood to plug the already swept zones. Resumption of waterflood, forces
the water to find a new path which leads to additional oil recovery. Gelling
solutions are typically an aqueous polymer solution and some kind of heavy metal
ions, such as chromium or aluminum, as cross linkers. Gels are highly non-
Newtonian and their rheological properties are unique for each gel system [23].
A somewhat different process for profile modification, known as combination
process, consists of the sequential injection of polymer solution and aluminum
citrate solution [24]. The permeability reduction is believed to be due to the
formation of layers of polymer/aluminum ion structure onto the wall surface of the
pores. The adsorbed polymer molecules from the first polymer treatment acts as a
base for the buildup of the structure. To enhance creation of the base and to increase
adsorption of the polymer molecules to the rock, the rock was first treated by
cationic polymer before the injection of the first anionic polymer cycle [25].
In-depth Permeability modification for the oil reservoirs with high permeability
variation has also been achieved by injecting colloidal dispersion gels into the
reservoir [26-28]. This consists of polymer with aluminum citrate crosslinker
injected as a homogeneous solution.
Above examples clearly reveal the complexity of the non-Newtonian fluid flow
through porous media associated with the gel treatment. The fluid flow behavior is
not simply a function of the rheological properties of the fluids involved, but the
interaction between the porous media and the fluids plays an important role.

2.6 Foam
Foams are a dispersion of a gas phase into a liquid phase, are unstable, and break
easily. However, addition of surfactants can increase their stability and prolong the
life of the foams almost indefinitely. Foams are injected into the petroleum
1188

reservoirs for the same reasons as gels. The profile modification for the case of
foam injection is aimed toward the displacement processes where gas is being used
as the displacing phase such as carbon dioxide flood, steam flood, or any other gas
injection schemes [29,30]. Flow of foams through porous media has been studied
in the literature [31,32]. The apparent viscosity of the foams flowing through
sandpacks was measured [33] and was found to be significantly higher than the
viscosities of the constituent fluids. This is one major factor for its effectiveness in
profile modification. In some studies the permeability reduction for foam flow was
found to be an order of magnitude larger than the permeability reduction predicted
just based on the gas/brine mixture viscosity [34].
Foam rheology is crucial in designing an effective permeability modification for
a given reservoir. Foams are non-Newtonian pseudoplastic fluid. Viscosity
measurement in a capillary viscometer revealed that the data fit the Ostwald and de
Waele (power law) relationship [35]. It was noticed that the bubbles do not behave
as rigid particles but flow and slip at the same time. Therefore, a fixed-slip velocity
could not be determined from the data.

3. N A T U R E OF POROUS MEDIA

Any solid body containing space to hold a fluid can be considered a porous
medium. This can be as simple as a pipe or as complex as riving tissue. Only the
interconnected pores, which could contribute to the flow, is of interest to the study
of the fluid flow through porous media, but not all the dead pores should be ignored.
Those dead pores, which have connection to the main flow path, could act as sink
or source for the flow. Experimental investigation of the fluid flow through porous
media is frequently carried out on idealized porous medium to avoid the complexity
of the true nature of porous medium.

3.1 Idealized Porous Media


Idealization, or simplification, of porous media is aimed toward studying a
particular aspect of fluid flow through porous media. For example, capillary tube
models are ideal to study viscous behavior of the fluid flow through porous media,
and the Hele-Shaw model is ideal for studying interface instability based on
perturbation theory. One should be careful in selection of idealized physical model
and its capability to reflect the physical phenomenon of interest. A bundle of
capillary tubes will not be able to reflect the viscoelastic nature of a fluid and,
similarly, in studying two-phase flow through porous media in the absence of
capillary effect, one should pick up a physical model with large pores to allow
neglecting capillary forces. Following are a few examples of idealized porous media
studied in the literature.
1189

3.1.1 Hele-Shaw Model


One hundred years ago, Hele-Shaw [36] observed that streamlines in an inviscid
flow can be visualized by making the gap of the two parallel plates small enough
so that a sheet of water as thin as the boundary layer could only flow through. He
added localized color to visualize streamlines. His remarkable observation put him,
as an experimentalist, ahead of the mathematicians to experimentally generate
streamlines for the geometries that fluid flow equations had not been solved yet. To
show his remarkable observation, two of his photographs generated for the flow of
water through a sudden enlargement are scanned and presented here. Figure 1 is
when the gap between the two plates is large, i.e. a thick sheet of water is flowing

Figure 2. Sudden enlargement (thin sheet) [36]

through the gap, and Figure 2 is when the gap is very small. This clearly shows that
by reducing the gap, a truly two-dimensional inviscid flow can be produced.
A Hele-Shaw model can be simply constructed by placing two parallel plates very
close together. Incompressible fluid flow through porous media and Hele-Shaw
model becomes analogous [37] and the average velocity components will be
identical if permeability in the Hele-Shaw model is defined as,
1190

h2
(i)
12

where, k is permeability and h is the spacing between the two plates of the Hele-
Shaw model.
Hele-Shaw cell has been used to study interface instability related to the viscous
fingering in porous media [38-45]. For example, Wiggert and Maxworthy [45]
injected air into a Hele-Shaw cell saturated with silicon oil. The unstable immiscible
interface resulted in a series of viscous fingering patterns that were photographed
and digitized. Their Hele-Shaw cell consisted of two glass plates separated by a gap
of 0.21 cm.
Viscous fingering is an interface instability which occurs when the displacing
phase is less viscous than the displaced phase. The creation of viscous fingers and
their growth have been visualized in Hele-Shaw models as well as sand packs.
Figure 3 is an example of the existence of the fingers even in laboratory systems
where porous media are much more uniform and much more homogeneous than the
actual reservoirs. Figure 3 is reproduction of one of the photographs taken by van
Meurs [46] in his transparent three-dimensional glass beads model where oil was
displaced by water.

Figure 3. Viscous fingering [46]

Addition of a small amount of polymer to the displacing water has been shown to
have significant effect on the stability of the front [47] and, consequently, produces
much higher recovery efficiency.

3.1.2 Porous Media Made out of Glass Rods


A two-dimensional flow that more closely mimics the actual reservoirs can be
physically modeled by assembling glass rods parallel to each other and having a
flow perpendicular to the axis of the rods. Vossoughi [48] used such a model to
1191

visualize the flow path in porous media and to study the non-Newtonian flow
distribution in a porous bed. He used a regularly spaced matrix of 6 mm diameter
glass rods in a triangular arrangement. Figure 4 is a schematic top view of the bed
geometry used in this study.

d ~ X

@
0 T0 C) Flow

Figure 4. Schematic of glass-rods bed geometry [49]

The cylindrical rods were positioned in a rectangular box of 5.2 by 6.5 cm inside
cross section. The box contained 40 rows of cylinders, each row containing 9 or 10
cylinders in width. The dimension of the smallest opening, d, varied randomly
between about 0.045 and 0.079 cm. Therefore, the bed is considered homogeneous
on a macroscopic scale, but on small or microscopic scale there are considerable
variations in pore diameter. A similar model with cylinders of shorter length and
larger diameter was used by Kyle and Perrine [50].

3.1.3 Porous Media Made out of Glass Beads


Visualization of three-dimensional flow in porous media can be achieved by
physical models made out of glass beads. Vossoughi [47] used a packed glass beads
model in his study on viscous fingering in immiscible displacement in porous
media. The model was made up of two transparent plates of 18 by 24 inches. The
plates were spaced V2 inch apart, and two sizes of beads were studied. The larger
bead size was 0.47 cm and the smaller was 0.15 cm in diameter.
Use of glass-bead packs to study fluid flow through porous media is a common
practice among investigators. Arman [51] used glass-bead pack in his study of
relative permeability curves. Naar, et al. [52] used a laboratory five-spot model,
made up of glass beads, to study the areal sweep efficiency of waterflooding.
Rapoport, et al. [53] performed waterflooding in oil-wet glass-bead packs to
1192

establish that a laboratory-developed relationship between linear and five-spot


flooding behavior holds regardless of the wettability or of the porous medium.
In general, flow of Newtonian and non-Newtonian fluids through granular and
non-granular packed beds is of considerable interest in many disciplines such as
filtration, chemical processes involving flow through catalyst bed, and oil recovery
operations. Packed beds are mainly used to study flow parameters in terms of
pressure drop and flow rate. Visualization of flow distribution in displacement
processes using non-glass beads pack can be achieved by X-ray shadowgraph
technique developed in early 1950's [54]. In this technique, an X-ray absorber is
added to either displaced or displacing phase. The phase containing the X-ray
absorbent casts a shadow on a piece of film upon exposure to an X-ray beam. The
density of the shadow is proportional to the saturation of that phase.

3.2 Complexity of True Porous Media


The porous media presented above are over-simplified compared to the pore
geometry of the actual rocks. Even the non-granular packed beds do not reflect the
many complex features associated with naturally occurring rocks. It is realized that
the naturally occurring rocks are composed of a variety of particle sizes with
various angularity, particle size distribution, and particle arrangement. The pore
space configuration is obviously different from that obtained by packing uniform
spheres in a bed. Furthermore, part of the pore space is naturally filled by clay and
cementing material. Cementing materials may block part of the pores and reduces,
or, in some cases, eliminates interconnectivity.
Figure 5 is composed of photographs of impregnated rocks revealing the complex
pore geometry for the two cases of fine and coarse intergranular sandstone [55].
The complex pore geometry is arising from many factors in the geological
environment of the deposit. The complexity in pore configuration becomes even
more pronounced for the case of carbonate rocks. Fractures and vugs are frequently
encountered in carbonate rocks as shown in Figure 6. In this figure, four typical
carbonate reservoir rocks are depicted with a) vugular porosity, b) vugular with pin-
point porosity, c) fractures, and d) with conglomerate [56].

4. FLUID/POROUS MEDIA INTERACTION

In addition to the complexity of the pore geometry, rock/fluid interaction becomes


more pronounced in the case of naturally occurring rocks. Surface wettability could
make a significant difference in flow behavior of wetting phase versus non-wetting
phase. Rock/fluid interaction plays a more important role in injecting of polymer
solution into a naturally occumng rock. In the case of naturally occurring rocks, the
role of adsorption and entanglement of polymer molecules, pore blockage by
polymer molecules, and inaccessible pore volumes to the polymer molecules
1193

become significantly more important and will contribute much larger share to the
non-Newtonian flow behavior through porous media.

4.1 Adsorption and Mechanical Entrapment


Adsorption of polymer molecules onto the rock surface plays a significant role in
the fluid flow behavior of polymer solution through porous media. Adsorption is not
always desirable and the loss of polymer molecules causes dilution and the
apparent viscosity of the polymer solution decreases with time. This could
eventually lead to the breakdown of its effectiveness as a mobility control additive.
Adsorption of the macromolecules will also reduce the pore size available for the
fluid to flow. The level of adsorption differs from one type of polymer to the other.
Polyacrylamide-type polymers, commonly used in oil displacement processes,
adsorbs strongly on mineral surfaces. The level of adsorption can be reduced by
partially hydrolyzing the polyacrylamide with a base such as sodium hydroxide,
potassium hydroxide, or sodium carbonate.
The amount of polymer adsorbed per unit surface area increases with polymer
concentration and with its molecular weight [57]. Rowland and Eirich [58] found
that the thickness of the adsorbed polymer (for uncharged polymer) is in the order
of the average diameter of the free polymer molecule coils in solution, i. e.
approximately proportional to M ~/2, where M is the average polymer molecular
weight. Michael and Morelos [59] studied adsorption of partially hydrolyzed
polyacrylamide and sodium polymethyl methacrylate on kaolinite and observed
higher adsorption with lower pH of the polymer solution. They concluded that the
mechanism of adsorption was hydrogen bonding. Similar observation is also
reported by Schmidt and Eirich [60].
There is a great deal of evidence in the literature that shows adsorption of the
polymer molecules is affected by the presence of electrolytes. Polymer adsorption
increases with increasing concentration of electrolytes [61-64]. Smith [62] reported
an adsorption of 200 gg of a partially hydrolyzed polyacrylamide per square meter
of surface area of silica powder when the polymer solvent was a 2% NaC1 solution.
Mungan [63] measured an adsorption of 400 gg/rn ~of silica powder for Pusher 700
in 2% NaC1 solution.
Mechanical entrapment of the polymer molecules is equally important in fluid
flow behavior of macromolecule solutions through porous media. Entrapment
mainly occurs at the pore throat, where the diameter becomes same or smaller than
the diameter of the approaching molecule. Adsorption of the polymer molecules
will definitely enhance entrapment by further reducing the pore diameter, especially
near the pore throat.
Dominguez [65] performed an experimental investigation of polymer retention in
the absence of polymer adsorption by performing his experiments in porous media
made of TEFLON | powder. Cores were prepared by compressing the powder. The
1194

Figure 5. Cast of pore space of typical reservoir rock. (a) Fine intergranular
sandstone; (b) coarse intergranular sandstone. [55]

Teflon powder cores have also been used by others. For example, Mungan [66] and
Lefebre du Prey [67] used Teflon cores for their study of wettability effects and
Sarem [68] used them to study polymer retention. Domingues [65] measured
polymer retention for his Teflon core within the range of 10 to 21 gg/g of Teflon
powder. Other reported values in the literature [64, 69-72] for polymer retention in
1195

Figure 6. Typical carbonate reservoir rocks (From Core Laboratories, Inc.) [56]

porous media are from 6 gg/g to as high as 160 gg/g of sand grain.
Pye [73] and Sandiford [74], in their experimental investigation of the flow of
polyacrylamide solution through porous media, observed a reduction in water
mobility of 5 to 20 times more than would be expected from the solution viscosity
alone. The additional resistance to fluid flow is mainly attributed to the adsorption
1196

and mechanical entrapment of the macromolecules. This effect is characterized by


introducing a resistance factor, F, defined as,

Fr=z <2)
p

where, ~, is the mobility and the indices w and p stand for water and polymer,
respectively. Mobility is defined as,

k
(3)

where, k is permeability and kt is viscosity.

The permeability reduction due to the adsorption and/or mechanical entrapment


is partially irreversible. This is clearly demonstrated by Burcik [75] in his flow
experiment of partially hydrolyzed polyacrylamide through slices of Berea
sandstone. He observed a reduction in brine mobility after the injection of the
polymer and the reduction persisted even after injecting 100 pore volumes of brine.
This effect, subsequently called residual resistance factor, F is defined as the ratio
of the initial brine mobility to the mobility of brine after the porous medium has
been exposed to the polymer solution.

~w (4)

, wp
where, )~wis the mobility of brine before injection of polymer and Lw~is the mobility
of brine after the bed was exposed to polymer injection and all the mobile polymer
molecules have been displaced by subsequent brine injection. Residual resistance
factor in the range of 1.08 to 15 has been reported in literature [76-79]. A value of
residual resistance factor of 15 means that the porous medium permeability has
been permanently reduced by a factor of 15.

4.2 Inaccessible Pore Volume


Inaccessible pore volume is a phenomenon associated with the flow of polymeric
fluids through porous media. This occurs when the size of the polymer molecules
in the solution becomes the same order of magnitude as the pore sizes. This causes
some of the pores which are smaller than the polymer molecules to be excluded
from the polymer flow. This phenomenon is well established and well recognized
1197

in the oil displacement processes where polymer solution is used as the displacing
phase. Unfortunately, it is not well documented in other disciplines when polymeric
fluid flow through porous media is discussed.
Dawson and Lantz [80] studied inaccessible pore volume in a consolidated
sandstone core by injecting polymer solution into the core. Two types of polymer
were studied. One was polyacrylamide-type polymer (Pusher 700, The Dow
Chemical Co.) and the other was a polysaccharide (XC biopolymer, Xanco, Div. Of
Kelco Co.). The cores were initially saturated with brine, then polymer solution was
injected continuously to satisfy the polymer retention. At complete equilibrium, i.e.
injected polymer concentration being the same as the effluent polymer
concentration, a slug of different polymer solution was injected. Injection of the
initial solution was resumed and continued till the end of the experiment. They
monitored the effluent polymer and the effluent salt concentrations and produced
concentration profiles as given in Figure 7. It is evident from Figure 7 that the salt
peak is arriving at 1 PV fluid injected as expected. However, the polymer peak is
arriving at 0.76 PV of fluid injected. The early arrival of polymer clearly indicates
that not all the pores are available for polymer to flow. This corresponds to an
inaccessible pore volume of 24%. Inaccessible pore volumes have been reported in
literature to range from 1 to 30% depending on the nature of the polymer and the
porous media involved [81,82].

Polymer Retention on Rock Satisfied ---~I I


Pulse Size
0.5 _ Before Injection of Pulse
~o
,o-o'o,~,
t~0.4 -
ii d
#
dO. 3 -
.6" Polyacrylamide \ ,
. ,o
~,

t- 0.2 -

#
Jo
#
: j $

~0.1 p'o
~.g,.., o,
~ L 0.0
0.6 0.7 0.8 0.9 1.0 1.1 1.2
Fluid Injection, P V

Figure 7. Early arrival of polymer peak due to inaccessible pore volume [80]

Adsorption and entrapment of polymer molecules in porous media, on the other


hand, cause the effluent polymer concentration to lag the solvent flow. The
combined effect, therefore, could totally or partially mask one effect by the other.
1198

That is, even if the polymer concentration profile appears at the same time as salt
concentration, this does not necessarily indicate that there is no inaccessible pore
volume and/or no adsorption and entrapment. It is possible that the dilution due to
adsorption and entrapment has been compensated and, therefore, concealed by the
inaccessible pore volume effect.
Inaccessible pore volume is an important, unavoidable phenomenon associated
with polymeric fluid flow through porous media and should not be ignored. Any
realistic fluid flow model for such a system should take into account all the
important fluid/rock interaction phenomena such as adsorption, entrapment, and
inaccessible pore volume.

5. MICROSCOPIC VIEW

Study of fluid flow through porous media at the microscopic level provides
information at the pore level. This information can be used to generate a predictive
model for the macroscopic behavior of the flow system. This obviously requires
significant simplification of the flow geometry. The idealized porous media
presented earlier can be taken advantage of for this purpose.

5.1 Photomicrography Technique


Photomicrography is a powerful technique to visualize flow path at the pore level.
Chatenever and Calhoun [83] are the pioneer in microscopic studies of the dynamic
fluid behavior in porous media. They used an idealized porous media consisted of
a layer of spheres sandwiched between two transparent parallel plates.
Simultaneous flow of oil and water was observed and photographed to reveal the
flow channels created by each phase. Later, Amoco Production Company [84]
(formerly Stanolind Oil and Gas Co.) studied fluid distribution in cylindrical sand
packs. The wetting phase was simulated by Wood's metal and the non-wetting
phase by colored plastic. At various saturations the Wood's metal and the colored
plastic were solidified in place in the core. The face of the core was magnified and
photographed at each time that it was shaved off. This produced a graphic three-
dimensional view of the wetting and non-wetting phase distribution when the
photographs were projected at motion picture speed.
Vossoughi [48,49] applied streak photography technique to visualize the flow path
and to measure velocity profile at the pore throat. In this technique, the velocity
profile can be established by measuring the length of the streaks of known duration.
He used an idealized porous media consisted of glass rods as described in Section
3.1.2. Flow visualization was achieved by adding approximately 2 cm 3 of aluminum
particles to 106 crn~ of solution. The aluminum particles had an average diameter of
20 microns. Light was allowed to enter from a narrow slit at the side while pictures
were taken from the top of the bed through a 12X microscope focused in the plane
1199

of fight. The fight was interrupted at known time intervals with a high-speed spoked
wheel. Therefore, velocity of fluid particles can be calculated by simply measuring
the length of an individual streak and dividing this measurement by the known time
interval. The schematic of the photomicrography assembly and the plan view of the
optical system used in this study are presented in Figures 8 and 9, respectively. The
photographs were projected onto a screen to achieve a magnification of
approximately 100X. The actual magnification was determined by projecting a
photograph of a precision steel rule.

@m
a era 9j - I~ Camera Stand

Phototubre ------.-....~~}

Microscope

9 U LI " Light Slit

J Flow Direction
! '" S! ,d

-] r
Microscope Stand Glass Rods Box

Figure 8. Schematic of photomicrography assembly [48]

Fluid flow distribution with pore opening was studied for Newtonian as well as
Non-Newtonian fluid. Pure glycerol of 13.6 poise viscosity was used as a
Newtonian fluid. The non-Newtonian fluid was a polymer solution of 2% Separan
AP 273 (a partially hydrolyzed polyacrylamide supplied by Dow Chemical
Company). Relaxation time of the polymer solution, as estimated from normal
stress measurements, was approximately 0.1 sec. Viscosities of the two fluids,
measured with a Weissenberg Rheogoniometer, are presented in Figure 10. The
polymer solution used in this study was highly pseudoplastic with power-law
indices as given on the figure. Figures 11 and 12 are typical photographs taken at
three different flow rates for the Newtonian and the non-Newtonian fluid,
respectively.

5.2 Flow Distribution with Pore Size


As mentioned earlier, Vossoughi [48] generated velocity profiles at the pore
openings from the photographs produced by streak photography technique. He
1200

Fluid Outle d
- q
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
L Fluid Inlet
,..__. ~.~

0 0 0 0

Glass Rods Box


Chopper Disk

Variable Speed
Transmission
Light Beam and Motor

Adjustable Slit

Converging Lens
Source

Figure 9. Schematic plan view of optical system [48]

100 IIIIII I I I III

r
Newtonian
.,..B

O
13_
:i.
~O Z:l
~

0
o
Polymer
>
(-. ]

m=7.5
< Power Law {
n = 0.43
0.1
1 10 100 1000
Shear Rate, 7, secl

Figure 10. Viscometric behavior of the experimental fluids [49]


1201

Figure 12. Typical photographs for non-Newtonian fluid flow [48]


1202

plotted the projected streak length, as measured at local magnification of


approximately 100, versus the position between the two cylinders. However, the
scale was chosen so that the profiles present the dimensionless interstitial velocity
versus position. He studied local velocity profiles for four different flow rates of
0.685, 1.717, 3.42, and 6.84 cm/s,
3 and he observed the relative distribution of the
flow among the pores being independent of the flow rates. Figures 13 and 14
present velocity profiles at one pore opening generated for Newtonian and
polymeric fluid flow, respectively.

,o I . i :o

~ Ai=0.360cm ~ ~ i =0.354cm
o
d = 0.0526 cm NCL = 0.0889 cm

.~~E to..- o~176176176 Ai = 0.382 cm ,w.......~...~....t


iS0
NCL = 0.17 cm
zx Q = 0.685 cm3/s o Q = 3.42 cm3/s
n Q=1.717 cm3/s 9 Q = 6.82 cm3/s

Figure 13. Experimental velocity profiles for Newtonian fluid [48]

In these plots, data points of all the four flow rates (whenever streaks are detected
at that particular pore opening) are included. The three profiles in each figure are
for three locations of the same pore. The profile designated by "d" is for the pore
throat (i.e. minimum opening between the two cylinders), and the one designated
by "NCL" (stands for Not Center Line) is for locations away from the pore throat.
The small variation in "d" values in the two figures is probably caused from the
small movement of the cylinders from one run to another. The fines in Figures 13
and 14 are eye-fitted curves through the data points. The graphical integration of the
1203

o Q = 3.42 cm3/s
"O5 9 Q = 6.82 cm3/s
r-
zx Q = 0.685 cm3/s
E
i5 ~'Ai = 0.236 cm 1 [] Q=1.717cm3/s
0 d = 0.0496cm

f
~_ ~,1: " A~ = 0.257 cm
t f
%1,, ~ ~,,~ A~ = 0.226 c
1
0 0
n NCL = 0.089 cm n NCL = 0.1016 cm

Figure 14. Experimental velocity profiles for non-Newtonian fluid [48]

areas under the curves are shown by A~. It should be noted that the area under the
curve, which is proportional to the flow going through that pore, should be the same
for all the profiles generated for a given pore regardless of the position of the profile
with respect to the pore throat. The sum of the individual flow rates in every pores
of a single row should equal the total flow rate. This provides an independent check
on the consistency of the data. In general, this was within +5 % of the total flow rate
determined by the pump calibration.
The total number of velocity profiles at various positions and flow rates studied
in this work [48] was in excess of 400 and included approximately 5000 individual
data points. The range of variables covered in this study is given in Table 1. The
range of shear rate, given in Table 1, is within the power-law region of the
viscometric behavior of the polymer solution presented in Figure 10. Therefore, the
power-law parameters estimated from Figure 10 provide a reasonable
approximation to the viscous properties of the polymer solution during the
photographic experiments.

Table 1. Range of variables for velocity measurements

Flow Rate Reynold' s Number Wall Shear


(c~/s) Newtonian Polymer Rate, (s -~)
.,,
Minimum 0.68 7.5 x 105 3.2 x 1 0 -4 3.3
Maximum 6.8 7.5 x 104 1.4 x 10.2 33.4

To predict flow behavior based on the viscometric properties of the fluids


1204

involved, Vossoughi [48] made further assumption that the flow geometry at the
pore level could be adequately simulated by flow through parallel plates. This
assumption was justified based on two observations: 1) Most of the pressure loss
will occur at the narrow openings between cylinders in a given row. 2) The narrow
openings between cylinders are small compared to the diameter of the cylinders
(0.05 cm compared to 0.6 cm). Thus, as shown by Figure 15, fluid flow at pore
level was approximated by flow between parallel flat plates with spacing d (the
narrow openings between the cylinders), but undetermined length in the flow
direction.

Figure 15. Schematic for parallel-plate analysis [49]

The equation of motion is then easily solved for creeping flow in the absence of
any elastic effects. Flow of power-law fluid in the z-direction between the parallel
plates produces [48,49]"

1 l+n n+l
u-( n ;~L d 2x (5)
n + l )(m )n(2 ) n [1-(--~-) n ]

or,

u -
[1 _(9_)
2x n+l
o] (6)
Umax

where, C is inserted as a tortuousity factor. Equation (6) reduces to Newtonian fluid


for n = 1. Equation (5) can be integrated to produce the volumetric flow rate per
unit depth of the pore.
1205

d12

q-2 Iudx (7)


o

For power-law model, u is substituted from Equation 5 and then integrated,

= 2n ( kp ) ,1- (_~_.)
d l+2n
,
(8)
q P 1 + 2n m CAL

Equation (8) reduces to Newtonian fluid by setting n = 1 and m - It,

2 Ap d)3
qN -- ~ (#CA L)(-2 (9 )

Equations (8) and (9) predict the volumetric flow rate being a function of pore
opening to the power (l+2n)/n and 3 for power-law and Newtonian fluid
respectively. For the power-law exponent of n = 0.43, as given in Figure 10,
Equation (8) predicts a dependency of the volumetric flow rate on pore opening to
the power greater than 4. Figures 16 and 17 compare the predicted and the
measured flow distribution with pore opening for the Newtonian and polymeric
fluid, respectively. Experimental data in Figure 17 indicate that the flow is
distributed among the pores more uniformly than what is predicted from power-law
model. It is even less than the dependence on pore size for the Newtonian fluid.

6. MACROSCOPIC VIEW

Non-Newtonian fluid flow through porous media can also be studied from
macroscopic point of view. In this approach attention is shifted from the pore level
to the global behavior of the fluid flow. Flow parameters, such as pressure drop and
recovery efficiency, are measured to characterize the flow process. Single phase and
two-phase flow of non-Newtonian fluid through porous media will be considered
here.

6.1 Single-Phase Flow


Single-phase flow of non-Newtonian fluids through porous media became a topic
of interest in research with the work published by Christopher and Middleman [85],
Sadowski [86], and Sadowski and Bird [87] in 1965. Interest has grown and a large
number of publications has appeared in the literature since then. The research work
has mainly been focused to devise a model to predict the resistance to flow of non-
Newtonian fluid through porous media. A majority of the papers deal with
1206

0.4 i i I i

0.3
Theory

0
q
5 o2

0 0

O
8
OD" I I I I
0.1
1 2 3
d/d min

Figure 16. Flow distribution with pore opening for Newtonian fluid [49]

0.4 I I I I i

_ Theory d4
0.3

q
Q 0.2

0.1 I I I
1 2 3
did min

Figure 17. Flow distribution with pore opening for polymeric fluid [49]
1207

viscometric behavior of non-Newtonian fluids and partially or totally ignoring


viscoelastic properties of the fluid. Some authors, such as Jennings, et al. [88],
Smith [62], and Harvey [89], believed that the effect of elasticity of the low
polymer concentrations of polymer solutions in porous media is insignificant.
Sadowski and Bird [87] reasoned that no significant elastic effect would be
observed provided the fluid relaxation time is small compared to the transit time
through a pore length.
It is well established that the effect of elasticity could be significant if the
geometry of the media is such that the transit time of fluid flow through a
contraction or expansion in a tortuous channel is comparable with the relaxation
time of the fluid. This was initially shown experimentally by Marshall and Metzner
[90] by flowing polymeric solutions through a sintered bronze disk. They observed
upward deviation from Newtonian line in friction factor plotted versus Reynolds
number. They correlated their data with a dimensionless number called Deborah
number, NDou, and provided a limiting value for the Deborah number above which
the additional resistance due to elasticity is expected to be felt.
Deborah number is defined as the ratio of two time scales. One time scale is the
characteristic time of the fluid, such as relaxation time, and the other is the
characteristic time of the flow, such as the time required for the fluid to pass
through one pore length.

Of (io)
NDe b -- Op r

in which Of is a relaxation time of the fluid and 0pr is a measure of the rate of
elongational deformation in the flow. Intuitively, elastic effects are expected to be
felt when these two time scales become same order of magnitude. Vossoughi and
Seyer [49] defined Deborah number as following:

relaxation time of fluid


NDeb = Inverse of velocity gradient in flow direction (ii)

caused by sudden constriction

They estimated the Deborah number by considering the elongation rate of a fluid
element moving from point A in Figure 4 to a point at the minimum opening [91 ].
Thus
0u z v/~0~-0
(12)
1208

where, v is superficial velocity defined as volumetric flow rate divided by cross


sectional area of bed, ~ is the area porosity at minimum spacing of a row, and Az
is the length defined by Figure 4. Therefore,

~sA z
Op r "-- (13)
%;

Substituting into Equation (10)

Ofv
NDeb = (14 )
CsaZ
Equation (14) allows estimation of the Deborah number from a knowledge of the
bed geometry and the fluid relaxation time. The choice of Deborah number and
even the choice of characteristic times are rather arbitrary. Some investigators have
used Ellis number instead, to correlate their pressure data [87]. There is no rigorous
treatment in the literature for additional pressure loss in porous media caused by
fluid elasticity. Kemblowski, et al. [92] provided a good review of the subject and
produced a table for the limiting values of Deborah number published in the
literature. The values, above which additional pressure drop due to elasticity has
been observed, range all the way from 0.05 to 3 with various definitions of Deborah
number.
Vossoughi and Seyer [49,91] presented data of friction factor plotted versus
Reynolds number for the Newtonian and polymeric fluid through their idealized
porous media consisted of glass rods. The upward deviation of the polymeric data
from the Newtonian data is clearly demonstrated in Figure 18. The data were also
presented in the form of friction factor multiplied by Reynolds number plotted as
a function of Deborah number. This is presented in Figure 19. In this figure data
from Marshal and Metzner [90] and the theoretical curve derived by Wissler [93]
are also included.
In a different study, Vossoughi [47] observed similar behavior for the flow of
polymer solution through glass beads pack.. The friction factor- Reynolds number
data of this study are reproduced in Figure 20. Two beds, one packed with large
glass beads and the other with small glass beads, were employed in this study. The
characteristics of the beds were presented earlier in Section 3.1.3. Data presented
in Figure 20 follow the linear dependency of friction factor with Reynolds number
for the Newtonian fluid. However, upward deviation from the straight line for
higher Reynolds number is evident for the polymeric data.
1209

10 5 i i

10 4

10 3

f 10 2

101 -
[] Polymer
10 0 - 0 Newtonian
f=l/N Re
10-1 1 I t 1 I
10-5 10-4 10-3 10-2 10-1 10 0 101
NRe
Figure 18. Friction factor - Reynolds number plot for glass rods bed [49]

As before, deviation from linearity can be also demonstrated by plotting the product
of friction factor by Reynolds number versus Deborah number. This is presented in
Figure 21. As observed from Figure 21, different polymer concentrations tend to
show independent trends of upward deviation. The polymer solutions of lower
polymer concentrations reveal earlier upward trend. This upward deviation is not
caused by adsorption, or plugging effects. Because, if this was the case, the more
concentrated solutions would have appeared to show higher upward deviation at an
even smaller Deborah number.
100 i i i liitil i i i itiiil i i i iiiii1 i i i~
9 Separan 273 (48)
- El
- zx
Carbopol (90)
Polyisobutylene (90) -I 1
10 _ O ET 597 (90)

- f ' N a e = 1 + 90N~)EB
f-NRe 1

1.0 zx ~ -
m
- [:n:IED n

f- NRe = 1 + 1 0 N b E B-
0.1 i i IIIIIII I I IIIIIII t t t titttl i i it
1-3 1-2 1-1 10
NDeb

Figure 19. Effect of Deborah number for glass rods bed [91 ]
1210

r 107 I I ! I i
00
o
t-
0
10 6
~

9 10 5
E
a
~

~ 10 4
0 Large Small
Beads Beads
o 10 a 9 O Tap Water
9 g~, 100 cs Dow Coming
t--
9 i-I 12500cs Dow Coming
.o 102 9 V 0.05% Separan AP 273
=m.
o (~ 0.1% Separan AP 273
t,...

u_ 1011 I I I
I =
10-7 10-6 10-5 10-4
10-3 10-2
Reynolds Number, NRe, Dimensionless

Figure 20. Friction factor- Reynolds number plot for glass-bead packs [47]

6.2 T w o - P h a s e F l o w
Two-phase flow through porous media is of significant interest in oil industry for
displacement of oil through the reservoir. Injection of non-Newtonian fluids, such
as polymer and micellar solutions, to displace oil is presently in practice.
Displacement process is an unsteady-state process because of the continuous
changes of the phase saturation with time. Displacement performance can be
predicted either by numerical solution of the partial differential equations involved,
or by the Buckley-Leverett model [94]. The Buckley-Leverett model, also called
frontal advance equation, is an old and systematic approach which can be solved
easily with graphical technique. The derivation of Buckley-Leverett model can be
achieved either by material balance, as discussed in detail by Willhite [95], or by
the method of characteristics applied by Sheldon, et al. [96], Scheidegger [97], and
Vossoughi [47]. The advantage of latter approach is its built-in sharp displacing
phase saturation gradient which must be entered arbitrarily in the other approach.
The method of characteristics is a powerful tool to deal with a system of linear
partial differential equations commonly encountered in the propagation of the plane
waves. Roughly speaking, a characteristic path is a propagation path along which
a physical disturbance, or entity, is propagated. The continuity equation for a one-
dimensional two-phase flow through porous media is given by,

(15)
1211

50 I I I I I i i t I i u
Large Beads Small Beads
or)
O 0.05% Sep. AP-273 9 0.05% Sep. AP-273
t.- o 0.1% Sep. AP-273 am 0.1% Sep. AP-273
.9 9 0.2% Sep. AP-273
. . . . . Newtonian Fluid Behavior
c'10
11)
.
E
m
_ O
CI
0 mm
$
- l l , i n
x
1 _o .,....o...
oo 9 or'&"
9 l_c~..li~. 4.~.,. tr--~'r-
.~...~
0.5 ,, i l i I i i I l i I i I
5x10 -4 10 -3 10-2 10-1 1 10 15
Deborah Number, NDe b, Dimensionless

Figure 21. Effect of Deborah number for glass-bead packs [47]

Oqo=-oAOSo (16)
Ox &

where, q is volumetric flow rate, S is volume fraction of the respective phases, t is


time, x is position, and subscripts w and o stand generally for the displacing and
displaced phases or simply water and oil respectively. Fractional flow rates are
defined as,

qw (17)

L qo
----I (18)

where, f is fractional flow rate and q is total volumetric flow rate. By substituting
Equation (17) into (15), or Equation (18) into (16), the following partial differential
equation is obtained.

~+~~(s)~-0
~?S q ~?S
(19)

where,
1212

df
f ' (S) : - -
dS (20)

The subscripts are omitted, because similar equations are produced for displacing
and displaced phases. Since phase saturation is a function of position and time, total
derivative of S becomes,

as as
. s - --d +x dt (21)

Equation (22) is the matrix representation of Equations (19) and (21).

1
#A
(22)

dt & ~ as

Along the characteristic curves following determinants are identical to zero.

q f'(S)
~A
(23)
=0
dt dx

1 0
=0 (24)

dt dS
1213

q f ' (S)
oa
=0 (25)
dS dx

From Equation (23),

d x : q f , (S) (26)
dt ~A

and, from Equation 24,

dS = 0 (27)

That is,

S = Const. (2 8 )

which also satisfies Equation (25). Therefore, S is constant along the characteristic
path, the equation of which is given by Equation (26). Equation (26) is identical to
the equation derived by Buckley and Leverett [94] using the material balance
concept. In derivation of Equation 26, no assumption was made concerning
rheological behavior of the solutions involved. Therefore, by proper selection of
fractional flow function, Buckley and Leverett model is applicable for Newtonian
as well as non-Newtonian fluids.
Equation 26 implies that at constant volumetric flow rate, q, a given saturation
plane, S, moves at constant speed, dx/dt. Therefore, a plot of x, the position of each
saturation plane, versus qt/~A must yield a set of straight lines with slopes equal to
the first derivative of the fractional flow function evaluated at that specific
saturation. Vossoughi [47,98] performed linear displacement tests in a transparent
rectangular bed packed with either 0.15-cm or 0.47-cm diameter glass beads. Clear
silicone liquids, trade named Dow Coming 200 Fluid with kinematic viscosities of
100, 1000, and 12500 cs, were used as oil phases. Water, various concentrations of
glycerol solutions, and several concentrations of aqueous polyacrylamide-type
polymer solutions were used as displacing phases. He measured the position of the
zero-water-saturation plane, which corresponds to the tip of the longest finger, from
the moving pictures generated for each displacement test. The position of this plane,
Xo, was plotted against qt/#A for Newtonian and non-Newtonian displacing phase.
Figure 22 is an example of the linear advance of the zero-saturation plane of a
1214

polymeric flood. Similar linear behavior was obtained for Newtonian as well as
non-Newtonian displacing fluids.
To explore the possible effect of polymer solution elasticity on the displacement
process, Vossoughi [47] undertook the following approach. Two polymers, one
more elastic than the other, were chosen. These polymers were Separan AP-273 and
Pusher 500 from Dow Chemical. The former one had a molecular weight in the
range of 10 million and the latter in the range of 2-3 million. The concentrations of
the two polymer solutions were adjusted to obtain a similar viscosity level at shear
rates encountered in the displacement tests. The relaxation time defined by Bueche
theory was chosen as a criterion to compare the elasticity effect of the two polymer
solutions. Assuming a molecular weight ratio of Separan to Pusher of
approximately 3, this produced a relaxation time ratio of 7.9 [47]. That is the
relaxation time estimated for 0.1% Separan AP-273 solution is approximately eight
times that of the 0.2% Pusher 500 solution.
Displacement behavior, such as advancing rate of the zero saturation plane,
saturation profile, and breakthrough recovery efficiency, are compared in Figures
23 and 24 and Table 2, respectively. Data indicate similar behavior with small
differences attributed to the small variations in viscosity levels. Therefore, the effect
of polymer solution elasticity on the displacement process was concluded to be
insignificant for the two polymers studied in this work.

Table 2. Comparison of the Separan and Pusher displacement recoveries

Flow rate, Recove_ry efficiency, % of recoverable oil


Cm~/s 0.1 ~ Separan 0.2% Pusher
0.16 28 31
0.64 18 17

7. METHODS OF PREDICTION

In spite of various attempts recorded in the literature, presently, there is no model


available to predict the behavior of the non-Newtonian fluid flow through porous
media. Most of the attempts have been focused on pressure drop prediction with
tittle progress. It is because of the complex nature of the porous media coupled with
complex rheological behavior of the non-Newtonian fluids. To this, the fluid-rock
interaction that was discussed earlier in this chapter, should be also added and kept
in mind.
In general, theoretical treatments of the non-Newtonian fluid flow through porous
media that are commonly practiced fall into three categories" 1) Those that are
1215

I I I I I I
45

40

,[3
35

30

IE
o 25 o,,
x
2O

15 q,, Y
O Small Beads
10 O,~ D Large Beads
,e
,e
I I I I I I
0 5 10 15 20 25 30 35
qt/~A, cm

Figure 22. Linear frontal advance of polymer displacement tests

basically an extension of Newtonian fluid flow through porous media; such as


hydraulic radius concept, Darcy's law adaptation, and friction factor- Reynolds
number relationship; 2) simple converging - diverging geometry coupled with some
kind of nonlinear constitutive equation; and, 3) empirical methods.

7.1 Hydraulic Radius Model


This is the most common treatment of the non-Newtonian fluid flow through
porous media, realizing that it only considers the viscous behavior of the fluid and
is insensitive to the fluid elastic component. The method has been exhaustively
covered in the literature and Kemblowski, et al. [92] have provided a good review
of the topic. Here, the technique will be only applied to the idealized porous media
composed of glass rods as presented in Section 3.1.3. The treatment here is parallel
to those of earlier works [85,100] but with parallel plates approximating the walls
of the capillary pores.
As depicted in Figure 15, fluid flow through parallel glass rods was approximated
1216

~ m m I jf/t

//
/,

40 //
s
// /
#
35 O

//
// o
30
l
Z
E 25
O #
X

20
--O 0.2% Pusher 500
q = 0.159 cm3/sec
15
- - .A - - 0.2% Pusher 500
q = 0.635 cm3/sec
10
--0-- 0.1% Sep. AP-273
q = 0.159 cm3/sec

- - - ~ - - - 0.1% Sep. AP-273


q = 0.635 cm3/sec
0 m
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Q t / ~A, cm

Figure 23. Frontal advance rates comparison [98]

by Vossoughi [48,49] as a rectilinear flow through parallel plates separated by a


distance equal to the pore throat. His approximation was justified based on the
observations that most of the pressure drop occurs at the pore throat and the
diameter of the glass rods were an order of magnitude larger than the pore throat
(0.6 cm compared to 0.05 cm). The choice of coordinate system and the direction
of flow for this derivation are defined in Figure 25.
1217

100 I I I I I I I I I

~90
~8o I
9
o. 1% Separan AP-273
0.2% Pusher- 500
m 70
ff
o
60
9 9
~ 50
o
0
4-, 40 9 9
m 30 9 9

.-. 20
9 I 9 9 9 o9 ooo
NlO
i I i vt i 9 t v
10 15 20 25 30 35 40 45 5"0
Position in the Bed, x, cm

Figure 24. Saturation Profiles comparison [98]

Y
|

Figure 25. Rectilinear flow through parallel plates


1218

The volumetric flow is obtained by integration of the linear velocity in the z-


direction as given by Equation 29.

lR

qi - 2 f f uzdydx (29)
00

where, q~ is volumetric flow rate; u z is linear velocity in z-direction, 1 is the length


of the channel in transverse direction, and R is half of the gap between the plates.
Since flow is assumed rectilinear, u, is a function of x only,

qi
(30)
o

Integrating Equation (30) by parts,

R R
Iu -u xl -IX .z (31)

o 0

Assuming no slip at the wall,

UzXI~--0 (32)

From Equations (31) and (32), Equation (30) can be written as,

R
q__Li= - I xduz (33)
2l o

or,

R
(34)
2-ii _ - X ( ~ x ) d x
o

The equation of motion for the laminar flow of Figure (25) produces,
1219

P~-PL) x AP
-- ' X

v-( M~ M~ (35)

where, z represents xz-component of the stress tensor, z=. From Equation (35) the
wall shear stress, %, is obtained.

AP
Vw =~R
AL (36)

Dividing Equation (35) by Equation (36) and solving for x yields,

R
X----T
(37)
Tw

or,

R
dx - ~ d'c (38)
Tw

By substitution of Equations (37) and (38) into Equation (34), Equation (39) is
produced,

qi _ R2 ~ (dUz )d. c
~-- 2 T (39)
21 "t:w o dx

or,

qi (40)
--Um -- ,./72 "~" T
21R w 0

where, u mis average velocity and dUz/dXis shear rate and will be designated by y;
therefore,

um = T2 Ty d~" (41)
w o

For the assumed steady flow, z is a function of 7 only. For Newtonian fluids,
1220

v --~ty (42)

and for a fluid approximated by the power-law model,

T - mlTl" (43)

where, ITI is the absolute value of 7, m and n are power-law indices. Let us define
F as an average nominal shear rate,

1-' -- um (44)
R

Equation (41) in terms of F becomes,

F 'r 2 r'}' dr (45)


w 0

To modify Equation (45) for the multi-cylinder bed, it is sufficient to introduce v/~)~
in place of u m,

Um ~ ~ (46)
~8

Where, v is superficial velocity, defined as volumetric flow rate divided by the


cross sectional area of the bed, and ~ is the surface porosity. Surface porosity is
calculated as,

~d

where, D is the diameter of the cylinder, and d is the minimum opening between
two adjacent cylinders of the same row. Therefore, Equations (44) and (45) become,

V 1 %
r
- =-~lv3 7dv (48)
G T, o
1221

where, subscript q~indicates that the equation has been modified for the porous bed.
Similarly, equation (36) can be modified for the porous bed as follows,

AP
-~R
T0- E (49)

where, L is replaced by L' to include tormousity of the flow path, i. e.,

L'=CL (50)

C is geometric constant which accounts for tormousity of the path of the fluid
particles. The magnitude of C could be approximately estimated as,

C = (zR)N
L (51)

where, ~R is half the circumference of the cylinder and N the number of rows. The
more accurate value of C must be determined experimentally by fitting the pressure
data to the model. Replacing L' in Equation (49) from Equation (50) gives,

AP
--R (52)
zo CL

Equation (48) can be integrated using Equation (42) for Newtonian fluid,

_TO
3p (53)

Solving for g,

T0
P - ~ (54)
3F 0

The expression found for g is defined as Darcy's viscosity. Substituting for z, and
F, from Equations (48) and (52),
1222

q~R 2 Ap
r 3C Lv (55)

From Darcy's law,

l.l-k Ap
Lv (56)

An expression for the permeability of the bed is obtained by comparing Equations


(55) and (56).

k - ~0~R2
(57)
3C

As stated earlier, C must be evaluated experimentally.

For power-law model, Equation (43) is solved for 7,

y -- _(L) 1/n
m (58)

where, the minus sign is chosen because of the choice of coordinates. Equation (58)
is substituted into Equation (48) and then is integrated,

vo (59)
l+2n m

or,

,.co - m[l + 2n Fo] "


(60)
n

Further, % and F, can be written in terms of k and C using Equations (48, 52, 57).

An
,r ~, - . , / ~ / COs -~ ~61~
L

Fo - v / a/3Ck~s (62)
1223

Equations (61) and (62) are substituted into Equation (60) and the resulting equation
is solved for v n,

(3k / CCs )1,2 (3Ckr mp


Vn -- (63)
m(l+2n)n L
n

From modified Darcy's law [85] for the flow of power-law model fluids through
porous media, Equation (64) is obtained,

(64)

where, H is the viscosity level parameter. An expression for H can be obtained by


comparing Equation (64) with Equation (63).

H - m 1 + 2n). )(1-.)/2
7( ,, (3CkO (65)

It is evident from Equation (65) that H = m = g when n = 1 for a Newtonian fluid.

Using Ergun's definition of friction factor [1 O1], one obtains,

2dO ~2Ap
f= pv2L (66)

where, f is friction factor and p is the density of the fluid. The Reynold's number,
NRo, is arbitrarily defined so that,

f - 1 / NRe (67)

Therefore, from Equations (57, 64, 66) and knowing that R = d/2, one could obtain
an expression for Reynold's number, NRo.

NR = Pdv2-~ (68)
24C0 H
1224

For a Newtonian fluid, n = 1, and therefore,

NR ~ _ pflv
24C0,r (69)

The experimental data of friction factors versus Reynold's numbers are compared
with the above derivation, i.e. equations (66,68,69) in Figure 18. A value of 0.682
was used for C in generating the data from pressure drop measurements. This was
determined experimentally using measured values of surface porosity and
permeability in equation (57). The inertia effect is felt around Reynold's number
equal to unity. The upward deviation due to elastic properties of the polymer
solution is observed long before the inertia effect is felt.
It should be realized that the opening between parallel plates are taken equal to the
minimum opening between the two neighboring cylinders of the same row. This
gives the highest possible value for a characteristic shear rate; therefore, smallest
value for viscosity. In fact, in the bed, there is a range of values of viscosity which
occurs and, therefore, the bed would have a somewhat higher value of average
viscosity. In terms of f - NRo, the low value of viscosity causes higher NRo which
makes the data points of polymeric solutions systematically shift to the fight of
Newtonian in Figure 18.

7.2 Converging-Diverging Channel


The analysis presented above reflects only the viscous behavior of the fluid and
does not take into account the elastic effect of the solution. Wissler [93] considered
flow of a viscoelastic fluid through a converging-diverging channel and performed
a third-order perturbation analysis of the flow. The system considered was a
converging-diverging flow between plane walls as shown in Figure (26).

111111. .ILLL I

Figure 26. Converging-diverging plane walls [93]


1225

Analysis was performed in a cylindrical coordinate system and conditions along


the z-axis were considered to be uniform. The r and 0 components of the equation
of motion are as following,

p ( ~12r O~ r ]d00qVr ]202


+ Ur ) --
r 30 r

(70)
__ _ _ ]_ G~T rr -Jr 1 ~ q~rO ( T rr -- T O0) +Pgr
cgr 3r r 30

p(~VO ~0 120 ~ 0 VrVO


St + Ur --Sr r 30 r )-

1 c?p t- 1 o~ (r2T 1 3 ~ oo (71)


- - ; 3-0 - - ~ -'~F rO) -] r cOO t- Pgo

The equation of continuity becomes,

1 cO ( F V r ) ' k1- -Ov


- ~o : O (72)
r3r r 30

The volumetric flow rate is obtained by integrating the velocity component in r


direction.

q - 2Io~rVr(r,O)dO (73)

A constitutive equation is needed to describe the rheological behavior of the fluid


involved. Wissler [93] utilized a nonlinear Maxwell model as following,

c)z ~j 3 r ~j 3 v~_ 3U 2
- ~~'ki +-~(~'~l%)6ij] - 2rleij (74)
~J + ~ [ at + vk c) x k 3 x~ rkj 3 x~

where, )~ is the fluid characteristic time, and 1] is a constant. He then applied


perturbation analysis by constructing solutions of the following forms,
1226

ot~

Vi -- E ~n vi(n)
n=0

oo

p - E ~-p(.) (75)
n=0

oo

~ ij -- E ~ n z ij (n)
n=0

where, n is the order of approximation. The lowest n corresponds to the Newtonian


fluid which is well characterized. He arrived at solutions for extra stress, extra
force, vorticity, and stream functions for each order of approximation. He further
evaluated the net force acting on the fluid at the surface of the channel and
intuitively extrapolated his solution to the pressure drop measurements in porous
media as following,

(76)
APviscoelasti c -- Apviscous [1 + A( - )2 ]
tJp

where, q/q~is the mean interstitial velocity, Dp is the average particle diameter, and
A is a constant. Equation (76) can be equally presented in terms of friction factor
- Reynolds number as given in equation (77).

fNR~-[I+A(r
Zq):
p ] (77)

From the definition of Deborah number given earlier by equation (10),

fNRe -- 1 + A N 2 Deb (7 ~ )

Experimental data presented earlier in Figure 19 seem to follow the trend of the
above prediction.
Other investigators [102-104] have also observed significant increase in flow
resistance for polymer solutions flowing through porous media. It is believed that
elongational flow regime is dominant in porous media. This is because of the rapid
changes of cross-sectional area of the pore space in the direction of the flow. James
and McLaren [103] attributes the increased flow resistance to the increase in
1227

elongational viscosity of the solution. This is directly related to the coil-stretch


transition of the polymer molecules. Coil-stretch transition occurs when the strain
rates to which the macromolecules are subjected exceed a critical value [105]. This
concept was successfully applied by Vorwerk and Brunn [106] to predict the
increased flow resistance for a random bed of spheres using three adjustable
parameters. Others [107,108] believe that the formation of transient networks of
polymer molecules is responsible for the increased flow resistance.

7.3 Empirical Methods


Empirical methods are basically application of Darcy's law with an apparent
viscosity, or mobility, to be correlated with the properties of fluid and porous
media. An example of this approach is the work done by Hejri [76]. He studied
flow of a biopolymer (Xanthan gum) through porous media made out of sandpacks
and tried to correlate the observed polymer mobility with the predicted values.
Darcy's law can be extended to the non-Newtonian fluid by introducing apparent
viscosity for the viscosity term,

V --
kho
w m
(79)
qL

where, k is permeability, v is superficial velocity, and 1] is the apparent viscosity to


be determined experimentally. In the shear-thinning region, power-law model can
be applied to describe the viscous behavior of the polymer solutions,

- m y (n-l) ( 80 )

Shear rate, 3' appearing in equation (80) is the average shear rate within the porous
media. There are different models to predict the average shear rate within a porous
media. In general they are a function of the superficial velocity and of an unknown
function of fluid and porous media properties, i. e.,

y = f(k,r (81)

Substituting equations (80,81) into equation (79) yields,

~, Ap (82)
v"- P L

where,
1228

k (83)
A*p - m [ f (k'(~'n)ln- 1

is the polymer mobility constant to be determined experimentally.


It is evident from equation (82) that a plot of pressure gradient, Ap/L, versus
superficial velocity, v, in log-log scale should produce a straight line with slope of
n and y-intercept equal to the inverse of L'p. Figure (27) is a typical plot generated
by Hejri [76] for the flow of 3000 ppm Xantan gum through a sand pack of 2896
md permeability.
The flow rate studied covered a wide range of Darcy velocity from 1.59 x 10.4
cm/s to 1.41 x 10.2 (0.45 to 40 ft/d). Data presented in this plot were generated by
changing the flow rate from low to high and vice versa; therefore, there was no
hysteresis in the pressure drop data. Similar plots were generated for polymer
concentrations of 1000, 1500, and 2000 ppm and sandpacks of different mesh sizes
with permeability to brine ranged from 894 md to 17,394 md. The permeabilities
to brine after the sandpacks were exposed to the polymer solution (residual
permeability, kwp) ranged from 525 md to 15260 md. Polymer solutions were
prepared with 30,000 ppm KC1 and 1500 ppm formaldehyde as biocide. The
polymer in this study was supplied by Pfizer Incorporated as a broth and is
commercially referred to as Flocon 4800.
The values of the flow behavior index, n, calculated from the core experiments
were somewhat different from the power-law exponent generated from viscometric
data for the same polymer solution. Polymer solutions were less shear thinning in
the sandpack as compared to the rheometer. A linear relationship was obtained
between the two flow behavior indices.
The experimental values of polymer mobility constants generated above were
correlated against predicted values based on four different models from literature.
These models are based on the capillary bundle model with variation in the
definition of the average shear rate in porous media. They produce different
functions of f(k, ~), n) appearing in equation (81). These functions can be extracted
from the models as following:

1) Teeuw and Hasselink [109] model for a power-law fluid,

3n+l 1
f (k,d?,n) = (84)
n ( 8 k ~ ) 1/2
1229

10 2 _ m
I I I llllll I I I llllll I I I lll___
m

~_. _

"ID
ml01
(.9 -
L_

kwp = 2,596 md _
Conc. = 3,000 ppm
t_

n 9 High - to- Low Rate


13 Low- to- High Rate -

lO0 I I i llllll i I i llllll I I I IIIII


lO-1 10 0 101 10 2
D a r c y V e l o c i t y , ft/d

Figure 27. Pressure gradient versus Darcy velocity in a sandpack [76]

2) Willhite and Uhl [77] model for a power-law fluid,

3n + 1)n_1 (85)
f ( k , O , n ) - ( 4n (0.5kq}) 1/2

3) Jennings, et al. [88] model for a Newtonian fluid,

1
f(k,r (86)
(0.5kr a/2

4) Modified Blake-Kozeny model [85],

9n+3 1
[12( )-n]l-n
f(k,r n (87)
( 1 5 0 k q } ) 1/2

In general, the predicted values of the polymer mobility based on the above
models were poor compared to the experimental values. Hejri [76] correlated his
experimentally measured polymer mobility constants with those predicted by the
above models. His empirical correlations are as following,
1230

~p - 1.076~;~ 92 (88)

/~p -- 1.964A~ (89)

Z p -- 2.685Z~ 884 (90)

/~p -- 1.638X~934z (91)

where, the numeral subscripts on ~ refers to the corresponding capillary bundle


model presented above.
Hejri [76] further derived an expression to predict the average shear rate in a
porous medium from the porous medium properties (1% and ~), flow behavior index
in porous media, n, and power-law parameters (n and m) for the polymer solutions
derived from steady shear measurements. In this derivation, it is assumed that the
apparent viscosity (viscosity in porous media) is equal to the bulk viscosity
(viscosity in rheometer) at a given value of shear rate. Equations (79,82) are re-
written for the porous media by replacing k with kp and n by no respectively.

kwe Ap (92 )
V -- ~ m
71 L

Vn c _A;Ap s
(93)

where, kw~is the residual permeability, and no is the flow behavior index in porous
media. From equations (92,93), apparent viscosity, 1] can be derived as,

kwp
-- ,~p*vl-nc (9 4)

Assuming the apparent viscosity in a porous medium, rl, is the same as the bulk
viscosity, g, represented by power-law model given by equation (95), an expression
for the average shear rate in porous media can be derived as presented by equation
(96).
1231

] . / - m y "~-1 (9 5 )

m~p 1 1-n~
9' - v X-.v (96)

where, n~ is the flow behavior index, and m, consistency index determined from
viscometric measurements, ~p* and n are determined from porous media
experiments as presented in Figure (27).

DEDICATION

This chapter is dedicated to Dr. F.A. Seyer, the author's former M.Sc. and Ph.D.
advisor.

REFERENCES

1. J.D. Ferry, Viscoelastic Properties of Polymers, John Wiley & Sons Inc.,New
York and London, 1961.
2. G.P. Willhite, Waterflooding, SPE Textbook Series Vol. 3, Soc. Pet. Eng.,
1986.
3. D.V. Boger, J. Non-Newtonian Fluid Mech., 3 (1977/1978) 87.
4. V.H. Giesekus, Rheol. Acta, 6 (1967) 339.
5. D.V. Boger and R. Binnington, Trans. Soc. Rheol. 21 (1977) 515.
6. D.V. Boger and H. Nguyen, Polym. Eng. Sci., 18 (1978) 1037.
7. G. Prilutski, R.K. Gupta, T. Sridhar, and M.E. Ryan, J. Non-Newtonian Fluid
Mech., 12 (1983) 233.
8. R.J. Binnington and D.V. Boger, Polym. Eng. Sci. 26 (1986) 133.
9. J.J. Magda and R.J. Larson, J. Non-Newtonian Fluid Mechanics, 30 (1988) 1.
10. C. Kemielewski, K.L. Nichols, and K. Jayaraman, J. Non-Newtonian Fluid
Mech., 35 (1990) 37.
11. D.O. Shah, editor: "Macro and Microemulsions," ACS symposium series 272,
American Chemical Society, Washington, D.C., 1985.
12. H. Bennett, J.L. Bishop, Jr., and M.F. Wulfinghoff, Practical Emulsions, Vol.
1" Materials and Equipment, Chemical Publishing Company, Inc., New York,
1968.
13. K.J. Lissant, editor: "Emulsions and Emulsion Technology, Part I, II, and III,"
Surfactant Science Series, Vol. 6, Marcel Dekker, Inc., New York, 1974.
14. M. Baviere, editor" "Basic Concepts in Enhanced Oil Recoverey Processes,"
Elsevier Applied Science, London and New York, 1991.
15. F.H. Poetmann, Microemulsion Flooding, Secondary and Tertiary Oil
1232

Recoverey Processes, Interstate Oil Compact Commission, Oklahoma City


(1974) 67.
16. H.Y. Jennings, Jr., Soc. Pet. Eng. J., 15 (3) (1975) 197.
17. W.K. Seifert, and W.G. Howells, Analyt. Chem., 41 (1969) 554.
18. P.A. Farmanian, N. Davis, J.T. Kwan, R.M. Weinbrandt, and T.F. Yen, ACS
Symp. Series, 91, American Chemical Society, Washington D.C. (1979) 103.
19. G.A. Pope, et al., Soc. Pet. Eng. J. (Dec. 1982) 816.
20. C. Castillo, and M.C. Williams, Chem. Eng. Commun., 3 (1979) 529.
21. D. Bouchez, A. Faure, G. Scherer, L.A. Tranie, and G. Antonini, COM: The
French Program, preparation, stabilization, and handling of COM, IV
International symp. on Coal Slurry Combustion, Orlando, Fla., 1982.
22. S. Vossoughi and O.S. A1-Husaini, Rheological Characterization of the
Coal/Oil/Water Slurries and the Effect of Polymer, 19u' International Technical
Conference on Coal Utilization & Fuel Systems, March 21-24, 1994,
Clearwater, Florida, USA.
23. R. Karrat and S. Vossoughi, Rheological Behavior of the Gel Systems used in
Enhanced Oil Recovery, Theoretical and Applied Rheology, Volume 1, edited
by: P. Moldenaers and R. Keunings, Elsevier Science Publishers (1992) 478.
24. R.B. Needham, C.B. Threlkeld, and J.W. Gall, SPE 4747 presented at the Third
Annual Symposium on Improved Oil Recovery, Tulsa, OK, April 22-24, 1974.
25. J.C. Mack, and M.L. Durall, SPE 12929 presented at the 1984 Rocky Mountain
Regional Meeting of SPE- AIME, Casper, Wyoming, May 21-23, 1984.
26. J.C. Mack, and J.E. Srrfith, SPE/DOE 27780, presented at the SPE/DOE Ninth
Symposium on Improved Oil Recovery, Tulsa, OK, April 17-20, 1994.
27. J.E. Smith, SPE 28989 presented at the SPE International Symposium on
Oilfield Chemistry, San Antonio, TX, February 14-17, 1995.
28. J.E. Smith, J.C. Mack, and A.B. Nicol, SPE/DOE 35352 presented at the 1996
SPE/DOE Tenth Symposium on Improved Oil Recovery, Tulsa, OK, 21-24
April, 1996.
29. G.J. Hirasaki, J. Pet. Tech., 41(5) (1989) 449.
30. T.W. Patzek and M.T. Koinis, J. Pet. Tech. 42 (1990) 496.
31. O.S. Owete, L.M. Castanier, and W.E. Brigham, Proc. AIChE Annual Meeting,
Las Vegas, Nevada, September 1982.
32. G.J. Hirasaki and J.B. Lawson, SPE 12129 presented at 58~ Annual Tech. Conf.
And Exhibition of Soc. Pet. Eng., San Francisco, 5-8 October 1983.
33. R.J. Treinen, W.E. Brigham, and L.M. Castanier, SUPRI TR 48, US
Department of Energy, Report DOE/SF/11564-13 (DE 86000260), 1985
34. J.T. Patton, et al., "Enhanced Oil Recovery by CO2 Foam Flooding," Second
Annual Report for U.S. DOE under contract NO. DE- AC21-78MC03259 (Nov.
1980).
35. J.T. Patton, S.T. Holbrook, and W. Hsu, Soc. Pet. Eng. J., 23(3) (June 1983)
1233

456.
36. H.S.S. Hele-Shaw, Nature 58 (1898) 34.
37. R.A. Greenkorn, Flow Phenomena in Porous Media, Marcel Dekker, Inc., New
York, 1983.
38. M.A. Nobles and H.B. Janzen, Trans. AIME, 213 (1958) 356.
39. R.L. Chuake, P. van Muers, and C. van der Poul, Soc. Pet. Eng. J., 188 (1959);
and Trans. AIME, 216.
40. J.L. Mahaffey, W.M. Rutherford, and C.W. Matthews, Soc. Pet. Eng. J. (March
1996) 73; and Trans AIME, 237.
41. R.A. Greenkorn, J.E. Matar, and R.C. Smith, AIChE J., 13 (1967) 273.
42. M.R. Todd, and W.J. Longstaff, J. Pet. Tech. (July 1972) 874.
43. S.D. Gupta, J.E. Varnon, and R.A. Greenkorn, Water Resour. Res. 9(4) (1973)
1039.
44. B. Habermann, Trans. AIME, 219 (1960) 264.
45. J. Wiggert, and T. Maxworthy, Physical Review E, 47(3) (March 1993) 1931.
46. P. van Meurs, Trans. AIME, 210 (1957) 295.
47. S. Vossoughi, Viscous Fingering in Immiscible Displacement, PhD
Dissertation, U. of Alberta, Edmonton, Alta., Canada (1976).
48. S. Vossoughi, Non-Newtonian Flow in a Porous Bed, M.Sc. Thesis, U. of
Alberta, Edmonton, Alta., Canada (1973).
49. S. Vossoughi and F.A. Seyer, J. of Can. Pet. Technol. 16(3) (1977) 110.
50. C.R. Kyle and R.L. Perrine, Can. J. Ch. E., 49 (Feb. 1971) 19.
51. I.H. Arman, Relative Permeability Studies, MS Thesis, U. of Oklahoma,,
Norman, OK (1952).
52. J. Naar, R.J. Wygal, and J.H. Henderson, Soc. Pet. Eng. J. (March, 1962) 13.
53. L.A. Rapoport, C.W. Carpenter, and W.J. Leas, Trans. AIME, 213 (1958) 113.
54. R.L. Slabod, and B.H. Caudle, Trans. AIME, 195 (1952) 265.
55. W.F. Nuss, and R.L. Whiting, Bull. Am. Assoc. Petrol. Geologists (November
1947) 2044.
56. J.W. Amyx, D.M. Bass, JR., and R.L., Whiting, Petroleum Reservoir
Engineering Physical Properties, McGraw-Hill Book Company, New York,
1960.
57. A. Silberberg, J. Physical Chemistry, 66 (1962) 1884.
58. F.W. Rowland, and F.R. Eirich, J. Polymer Science, Part A-1, 4 (1966) 2401.
59. A.S. Michael, and O. Morelos, Ind. Eng. Chem., 47(9).
60. W. Schmidt, and F.R. Eirich, J. Phys. Chem., 66 (1962).
61. W.P. Shyluk, J. Polymer Sci., part A-2, 6 (1968) 2009.
62. F.W. Smith, J. Pet. Tech. (Feb. 1970) 148.
63. N. Mungan, Revue de L'Institut Francais du Petrole, 24(2) (1969) 232.
64. M.J. Szabo, SPE 4668 presented at SPE 48 ~ Fall Meeting, Las Vegas, Nevada
(Sept. 1973).
1234

65. J.G. Dominguez D., Polymer Retention and Flow Characteristics of Polymer
Solutions in Porous Media," M.Sc. Thesis, U. of Kansas, Lawrence, KS (1974).
66. N. Mungan, Soc. Pet. Eng. J. (Sept. 1966) 247.
67. E.J. Lefebre du Prey, Soc. Pet. Eng. J., 13(1) (Feb. 1973).
68. A.M. Sarem, SPE 3002 presented at the 45" Annual Fall Meeting of the Society
of Petroleum Engineers of AIME, Houston, TX (Oct. 1970).
69. A.C. Uzoigwe, F.C. Scanlon, and R.L. Jewett, SPE 4024 presented at SPE 47 ~
Annual Fall Meeting, San Antonio, TX (Oct. ! 972).
70. R. Dawson and R.B. Lantz, Soc. Pet. Eng. J. (Oct. 1972) 448.
71. G.H. Hirasaki, and G.A. Pope, SPE 4026 presented at SPE Fall Meeting, San
Antonio, TX (Oct. 1972).
72. N. Mungan, J. Can. Pet. Tech., 8(2) (1969) 45.
73. D.J. Pye, Trans. AIME (1965) 911.
74. R.B. Sandiford, J. Pet. Tech. (August 1964) 917.
75. E.J. Burcik, Producers Monthly, 29(6) (June 1965).
76. S. Hejri, An Experimental Investigation into the Flow and Rheological Behavior
of Xanthan Solutions and Xanthan/Cr(III) Gel System in Porous Media, PhD
Dissertation, U. of Kansas, Lawrence, KS (1989).
77. G.P. Willhite and J.T. Uhl, Correlation of the Flow of Flocon 4800 Biopolymer
with Polymer Concentration and Rock Properties in Berea Sandstone, Water-
Soluble Polymers for Petroleum Recovery, G.A. Stahl and D.N. Schultz
(editors), Plenum Press, New York (1988).
78. F.D. Martin, et al., SPE 11786 presented at the 1983 Soc. Pet. Eng. International
Symposium on Oilfield and Geothermal Chemistry, Denver (June 1983).
79. W.B. Gogarty, Soc. Pet. Eng. J. (June 1967) 161; and Trans. AIME, 240.
80. R. Dawson and R.B. Lantz, Soc. Pet. Eng. J. (Oct. 1972) 448; and Trans. AIME
253.
81. B. Shah, An Experimental Study of Inaccessible Pore Volume as a Function of
Polymer Concentration during Flow through Porous Media, M.Sc. Thesis, U.
of Kansas, Lawrence, KS (1978).
82. D.S. Hughes, et al., Soc. Pet. Eng. Res. Eng. J. (Feb. 1990) 33.
83. A. Chatenever and J.C. Calhoun, Jr., Trans. AIME, 195 (1952) 149.
84. Fluid Distribution in Porous Systems - A Preview of the Motion Picture,
Stanolind Oil and Gas Co. (1952); subsequently reprinted by Pan American
Petroleum Corp. and Amoco Production Co..
85. R.H. Christopher and S. Middleman, Ind. Eng. Chem. Fund., 4 (1965) 422.
86. T.J. Sadowski, Trans. Soc. Rheol., 9 (1965) 251.
87. T.J. Sadowski and B. Bird, Trans. Soc. Rheol., 9 (1965) 243.
88. R.R. Jennings, J.H. Rogers, and T.J. West, J. Pet. Technol. (March 1971) 391.
89. A.H. Harvey, An Investigation of the Flow of Polymer Solutions through
Porous Media, PhD Dissertation, U. of Oklahoma, Norman, OK (1967).
1235

90. R.J. Marshal and A.B. Metzner, Ind. Eng. Chem. Fund., 6 (1967) 393.
91. S. Vossoughi and F.A. Seyer, Can. J. Chem. Eng., 52 (Oct. 1974) 666.
92. Z. Kemblowski, M. Dziubinski, and J. Sek, Flow of Non-Newtonian Fluids
through Granular Media, from: "Transport Phenomena in Polymeric Systems,"
R. A. Mashelkar, A.S. Mujumdar, and R. Kamal (editors), Ellis Horwood Series
in Physical Chemistry, John Wiley & Sons, New York (1989) 117.
93. E.H. Wissler, Ind. Eng. Chem. Fund., 10(3) (1971) 411.
94. S.E. Buckley and M.C. Leverett, Trans. AIME, 146 (1942) 107.
95. G.P. Willhite, Waterflooding, SPE Textbook Series Volume 3, Society of
Petroleum Engineers, Richardson, TX, 1986.
96. J.W. Sheldon, B. Zondek, W.T. Cardwell, Jr., Trans. AIME, 216 (1959) 290.
97. A.E. Scheidegger, The Physics of Flow through Porous Media, University of
Toronto Press, 3~dedition, Chapter 9 (1974).
98. S. Vossoughi and F.A. Seyer, Ind. Eng. Chem. Fundam., 23(1) (1984) 64.
99. S. Middleman, The Flow of High Polymers, Continuum and Molecular
Rheology, Interscience Publishers, John Wiley & Sons, Inc., New York (1968)
148.
100. J.G. Savins, Ind. Eng. Chem., 16(10) (October 1969) 18.
101. S. Ergun, Chem. Eng. Prog., 48(2) (1952) 89.
102. D.L. Dauben and D.E. Menzie, J. Pet. Tech. (August 1967) 1065.
103. D.F. James and D.R. Mc Laren, J. Fluid Mech., 70 (1975) 733.
104. W.M. Kulicke and R. Haas, Ind. Eng. Chem. Fundam., 23 (1984) 308.
105. P.G. de Gennes, J. Chem. Phys., 60 (1974) 5030.
106. J. Vorwerk and P.O. Brunn, J. Non-Newtonian Fluid Mech., 41 (1991) 119.
107. J.A. Odell, A.J. Muller, and A. Keller, Polymer, 29 (1991) 119.
108. S. Rodriguez, C. Romero, M.L. Sargenti, A.J. Muller, and A.E. Saez, J. Non-
Newtonian Fluid Mech., 49 (1993) 63.
109. D. Teeuw and F.T. Hesselink, SPE 8982 presented at the 5th International
Symposium on Oilfield and Geothermal Chemistry, Stanford, California (May
1980).
1237

FLUID DYNAMICS OF FINE SUSPENSION FLOW

Y. A. Buyevich

Center for Risk Studies and Safety, University of California Santa Barbara,
6740 Cortona Dr., Santa Barbara, 6.4 93117

1. INTRODUCTION

Flows of colloids and suspensions of fme particles are important for a wide
variety of multifarious applications encountered in nature and in various fields of
modem industry. First of all, engineers and researchers are commonly interested
in global hydraulic characteristics of such flows. However, more subtle
peculiarities featured by such flows may be of great practical concern as well.
For instance, it is imperative to obtain a certain information on the kinetics of
flow stratification which would allow to evaluate the rate of particle deposition
on channel walls, the intensity of phase separation in vertical flows, etc. The
information of such a kind should also be of crucial significance in modifying
many widespread industrial processes with an ultimate purpose in view to
improve quality and performance characteristics of those processes.
However, theoretical study of suspension flow usually meets with formidable
difficulties of principal nature, even in cases where only comparatively sketchy
hydraulic characteristics pertaining to a smooth laminar flow as a whole are
under question. These difficulties, which are briefly indicated and discussed
below, are in the first place due to the obvious lack of a sufficiently reliable
hydrodynamic model even for laminar suspension flow. Such a model is at
present not available, despite considerable efforts undertaken during the last
decades in the field of both suspension hydrodynamics and rheology.
Correspondingly, the main intended objective of this Chapter consists in putting
forward some relatively simple considerations that hopefully will prove
themselves both beneficial and advantageous in the matter of developing the
needed fluid dynamic model for laminar flow of concentrated suspensions. A
1238

tangible tentative approximate version of such a model is presented and


discussed in this Chapter as well.
When dealing with colloid and fine suspension flows, there is a strong
temptation to consider the dispersed medium under question as a fictitious
homogeneous fluid characterized by some effective properties of its own, such as
density and viscosity, in which case any flow is presumed to be governed by a
single set of mass and momentum conservation equations. The effective
properties are then implicitly regarded as quantities dependent on the fictitious
fluid composition, and they may be identified on either theoretical or empirical
grounds, or both.
However, such a simple picture fails to adequately portray the flow in
numerous cases where the fictitious fluid composition happens to be unsteady or
non-uniform, so that its effective properties cannot be looked upon as invariable,
but are instead certain, and not always known, functions of time and coordinates.
First of all, this occurs in different flow processes that by their very nature imply
separation of the suspension phases, or of different particulate species of the
dispersed phase, such as sedimentation, crossflow microfiltration, and field-flow
fi'actionation. However, this often occurs also in colloid and suspension flow in
cases where separation is out of the question, and flow stratification arises as a
consequence of uncontrollable natural factors. In such cases, the fictitious fluid
properties can be again regarded as functions of the local mean suspension
concentration. To find the concentration field, and thus to determine the said
properties as fimctions of coordinates and time, an analysis based on the concept
of the only fictitious fluid is clearly insufficient. An alternative continuum
approach must instead be employed for this purpose, within the framework of
which the suspension phases are viewed as separate interacting co-existing
media with their own properties. This means that at least two sets of averaged
field conservation equations have to be used to describe the suspension flow, one
for each of the phases. Unknown variables of these equations are the mean
particle and ambient fluid velocities, the mean interstitial pressure, and the mean
suspension concentration.
A serious stumbling block commonly confronted when attempting to find out
the suspension concentration distribution on the basis of this continutun approach
is caused by the fact that stratification of suspension flows is governed not only
by regular forces experienced by suspended particles, but also by particle
diffusion due to different physical reasons. Hence it follows that the conservation
equations that do not include the impact of diffusional effects can hardly be
anticipated to be helpful in the matter of determining the wanted concentration
distribution. This expectation comes to be true, if random particle and fluid
fluctuations that give rise to diffusional phenomena and that contribute to
1239

stresses acting in flow of the suspension phases are completely left out of
account.
As has been discussed in [1 ], the intrinsic inconsistency of conventional fluid
dynamic schemes of suspension flow which do not allow for stress contributions
due to the fluctuations consists in the fact that their governing equation do not
permit the concentration field to be found alongside the fields of mean phase
velocities and fluid pressure. The latter fields can be obtained merely on
condition that the former one is given beforehand. However, the mere fact of the
concentration field having been prescribed in advance decreases by unity the
number of unknown variables in the governing equations, so that one of these
equations cannot actually be satisfied and, in fact, has to be discarded.
The physical reason of this deficiency is usually due to the occurrence in a
flow of a transverse force experienced by suspended particles in the direction
normal to the flow streamlines. This force causes the particles to migrate in the
same direction and, inasmuch as the corresponding particle flux is not
compensated for by any diffusion flux, it must eventually give rise to the flow
stratification. The simplest example is provided by flows of not neutrally buoyant
suspensions in inclined channels. In such flows, the lateral component of gravity
as corrected for buoyancy brings about the initiation of both a region which is
filled with close-packed particles and a region which is entirely devoid of
particles. Under steady conditions, the whole flow domain must consist of only
such limiting regions, without any region containing suspended particles in
between [ 1].
In vertical flows, the similar particle migration is stipulated by the inertial
lateral lift force acting on particles that rotate in a sheared flow [2]. When being
not compensated by diffusion, this migration virtually results in the accumulation
of particles in either a central or a peripheral flow region, depending on whether
the particles are heavier or lighter than the suspending fluid.
Our primary purpose consists in demonstrating in the present Chapter what
has to be actually done in order to develop a consistent workable scheme aimed
at tackling stratified flows of colloids and freely dispersed suspensions. The term
"finely dispersed" insinuates that collisions of suspended particles are totally
irrelevant in the interparticle exchange by momentum and energy which is
therefore assumed to be carried out via the velocity and pressure fields of the
intervening fluid. The preference is given to simplicity and clarity of
presentation, rather than to strictness of our development which seems
unnecessary at its present stage . For this reason, we shall address only the
simplest possible type of dispersions, that is, a suspension of identical solid
spheres in an incompressible Newtonian fluid which exhibit steric and
hydrodynamic, but not any molecular mterparticle interactions that are usually
1240

specific to colloidal systems. The spheres are assumed to be free of embedded


dipoles of any physical origin. Besides, when needed, we shall prefer sometimes
to use certain semi-empirical, albeit intuitively perceptible, relationships and
considerations stemming from other cognate fields of science, rather than to stick
to rigorous derivation of the wanted relationships.
The Chapter is organized as follows. As a beginning, we shall briefly review
averaged conservation equations and relevant constitutive rheological equations
for the phases of a suspension at neglect of random particle and fluid
fluctuations. After that, we shall explain the influence of thermal particle
fluctuations on the rheology of Brownian suspensions with the help of a specific
"thermodynamic" constituent of the interphase interaction force. We shall also
consider an alternative way to describe this influence by means of specifying
appropriate contributions to the effective stress tensors that affect flow of the
suspension phases.
In the next place, we shall address hydrodynamically induced particle and
fluid fluctuations that may obtain as a result of two possible physical
mechanisms. The first mechanism bears upon random displacements of particles
caused by relative motion of neighboring particle layers in shear flow. The
second mechanism is due to the relative fluid flow working at random
fluctuations of the suspension concentration, and thus originating peculiar
"pseudoturbulent" fluctuations. The impact of both shear-induced and
pseudoturbulent fluctuations on suspension flow is again described by
introducing pertinent contributions to the effective stress tensors.
Because a comprehensive presentation of these topics takes a considerable
space, it is impossible to address application of the theory developed to various
particular flow problems, which have thus to be considered in the future.

2. FIELD CONSERVATION EQUATIONS

We start with formulating averaged field equations of mass and momentum


conservation for the interpenetrating co-existing continua that model the
continuous and dispersed phases of a suspension with no fluctuations. Such
equations are to be obtained by using some smoothing procedure, such as those
implying time [3] or volume [4] averaging of local conservation equations that
are presumed to be valid within the phase materials. Unfortunately, however,
these averaging procedures result in seemingly incongnlent forms of the field
equations, as exemplified in [5]. For this reason, we prefer to use the technique
of averaging over the ensemble of physically possible configurations of the
assemblage of suspended spheres as developed in [6] and as recently discussed
in [7]. This technique leads to a set of field equations for mass conservation
1241

3t ' 3t
+v.(ow)-0 (1)

and for momentum conservation

cpf ~t9+ v . V / v = V . c r - f - G p f V O (2)

C9
CPp(~
~ + w . V I w = f - qkppV ff~ (3)

where v and w are the mean fluid and particle velocities, ~b and e = 1 - ~b are
the particle concentration by volume and the void fraction, respectively, and (I)
is the potential of an external body force field. Vector f describes the average
mterphase interaction force per unit volume, and ~ is the effective stress tensor.
What happens to be of principally significance is that effective stresses
appear only in the equation (2) of continuous phase momentum conservation,
whereas averaged stresses that might be expected to affect mean flow of the
dispersed phase identically tum to zero [6,7]. Note, however, that this general
conclusion is valid only for disperse systems in flow of which the impact of
fluctuations on the phase stresses is negligible, so that it may be ignored.

3. CLOSURE OF FIELD EQUATIONS

According to the ensemble averaging technique, both interphase interaction


force and effective stresses that influence the continuous phase flow are
expressible in terms of integrals over the surface of a chosen "test" suspended
sphere. The corresponding mtegrands involve stresses wlfich are averaged over
particle configurations that are conditioned by the test sphere center having been
positioned at a prescribed point. These conditioned averages can in principle be
determined by solving a problem for mean suspension flow around the test
sphere. Averaged equations that govern this flow contain integrals of stresses
obtained by averaging over configurations that have centers of two spheres fixed
in space. As explained in [6,7], when continuing such a process, we arrive an a
practically infinite chain of interconnected equations, and of hydrodynamic
problems, for flow around different nmnbers of fixed spheres, and a familiar task
arises of cutting off this chain.
An approximate method of resolving this task implies viewing the test sphere
as one immersed in a fictitious medium whose properties vary with the distance
from the test sphere center, but coincide with those of the suspension as a whole
1242

as this distance goes to infinity. Within the framework of a certain


approximation, the exact type of this variation is shown in [6, 7] to be dictated
by the pair distribution function for suspended spheres. As a result, depending on
a model used to describe this function, we obtain different rheological models for
the suspension.
In practice, the pair distribution function is influenced by suspension flow,
and so cannot be regarded as an equilibrium property. This fact was well
recognized by Batchelor who succeeded in calculating relative viscosity of a
dilute suspension for flow of pure elongation [8], and also for simple shear flow
in the limit of strong Brownian motion of the spheres, when the suspension
approximately behaves like a Newtonian medium [9]. In the generalized case of
an arbitrary flow, averaged stresses are not Newtonian, and they depend on the
type of flow, due to hydrodynamic interactions that affect space distribution of
suspended particles.
If the suspension is not dilute in the sense that not only binary hydrodynamic
interactions are essential, one might presume simultaneous interactions of many
particles to work similarly to strong Brownian motion in rendering the structure
of a flowing suspension statistically isotropic, and in thereby making the
suspension behave as a Newtonian fluid. It is such a point of view that is actually
adopted when describing the pair distribution function with no allowance made
for effects that flow is likely to exert on the suspension statistics. In what
follows, we are going to neglect the effects on the short-range order in a flowing
suspension as produced both by the flow itself and by particle random
fluctuations. Correspondingly, we shall use some representations for the pair
distribution function resulting from the equilibrium statistical mechanics of dense
systems of hard spheres.
Having represented the conditional averaged stresses that act at the test
particle surface in terms of the unconditional ones, we must fitaher proceed to
expressing the latter stresses, as well as the interphase interaction force that
appear in equations (2) and (3), through unknown variables of the field equations
and their derivatives. This has been suggested in [6,7] to achieve with the help of
the self-consistent field theory. According to such an idea, the averaged stress
tensor divergence and the interphase interaction force vector are first presented
as linear combinations of relevant vector quantities that are likely to completely
determine the situation at any point within the flow domain. This enables us to
close the governing equations, and ultimately, to solve the flow problem around
the test sphere and to calculate the stresses at the test sphere surface. After that,
the original integrals representing the stress divergence and the interphase
interaction force can be calculated to yield expressions for the said quantities as
functions of tmknown coefficients involved in the aforementioned linear
1243

combinations. The last coefficients have to be found afterwards from self-


consistency conditions which require these expressions to be identical to the
original linear combinations. This leaves us with a set of algebraic equations in
which the wanted coefficients serve as unknown variables. Solving these
algebraic equations determine all constitutive rheological equations for fine
suspensions, and so makes for final closure of the suspension fluid dynamic
theory.

3.1 Effective Medium Model


The simplest possible rheological model comes about as a result of the
supposition that the averages conditioned by positioning the test sphere center at
a certain point are indistinguishable from the unconditional averages. The
fictitious medium the test sphere is assumed to be immersed into is then uniform,
and its properties are precisely the same as those of the suspension as a whole.
This model corresponds to entirely ignoring the non-overlapping property of hard
suspended spheres, which are thus permitted to overlap. It is sensible to expect
this model to be approximately valid for moderately concentrated suspensions in
which the overlapping of the spheres is unlikely to produce a significant effect.
Similar models according to which discrete particles are supposed to be inserted
into an effective homogeneous medium were repeatedly formulated during the
last decades on purely empirical grounds. As a matter of fact, such models were
developed not only in connection with suspension rheology, but also while
treating effective properties of various dispersions and composite materials, such
as thermal conductivity, electric and magnetic permeability, moduli of elasticity,
etc. Unsteady suspension flows have been recently considered in detail within
the framework of a simple model of this kind in [10], and here we are going to
exhibit main conclusions of that paper.
According to [10], the effective stress tensor displays relaxation phenomena
the type of which depends on a sign of a newly introduced quantity joe that has
the density dimensionality

'I 4 5
Pe = -~ P - -~ Pp + -~ M (3p - qkpp , p = ~pf + r M =~
1
(4)
1-2.50

The stress tensor is expressible as follows:

pe a2
cr=-pi+ , o'~=21ufM 1+ T~ +w" ev , (5)
- F,S M
1244

where p is the mean interstitial pressure, I is the unit tensor, a is the sphere
radius, ktf is the ambient fluid viscosity, and e~ is the strain rate tensor
corresponding to the continuous phase mean velocity field. Quantity M plays the
role of relative suspension viscosity attained in steady flow, whereas Tu is a
specific relaxation time. If p~ is positive, as is the case for colloids and for the
majority of suspensions encountered in practice, an adjunct relaxation
relationship can be formulated as [10]

--
cgt
+ w- ev - ev,, - ev
' ev's 21ufM
= ~ (6)

Hence it follows that the strain rate tensor relaxes to its value which would
establish itself in steady flow characterized by a given actual tensor of mean
viscous stresses. In the opposite case of negative p~, which can be realized for
gas-solid mixtures, we have another relaxation relationship, instead of that in
equation (6),

According to this relationship, the mean viscous stress tensor now relaxes to
its steady value corresponding to a given strain rate tensor.
These relaxation phenomena come about as a natural consequence of the
frequency dispersion effect for the suspension viscosity. This effect means that
the effective suspension viscosity, as it manifests itself in unsteady
monochromatic suspension flow, happens to be dependent on flow frequency.
The frequency dispersion may be quite insignificant as far as the averaged
stresses are concerned. However, as is demonstrated below, it proves to be
rather important in the matter of influencing an inertial part of the total interphase
interaction force that is inherent in dense suspension flow.
The interphase interaction force attributed to a unit volume of the mixture is a
sum of a few contributions having different physical meaning

f = f d +f/~ + fF + fi +fb (8)

Expressions for the viscous drag force, the hereditary Basset force, and the
Faxen force due to flow non-uniformity are of exactly the same form as the
corresponding force constituents for a single sphere in an unbounded fluid [ 11 ],
save for the fluid viscosity and density being substituted by the effective
suspension viscosity and density. They are expressible as
1245

f_9r uzM
a 2
(v_w) (9)

fB -
9~2( .tllfMll/2i(~ VII
p Jra2 ) + w-
dr'
(v - w)t - t' 4 t - t'
(10)
--00

f F -- 3~4 luf M Av (l])

The inertial force equals to

fi - /(
p - - ~ pp + 3~bPe ~ - + w v-w,
Supposedly, this force describes not only the effect of fluid virtual mass
acceleration and the effect of accelerated motion of the ambient fluid itself, but
also the influence on the interphase interaction force of the frequency dispersion
of effective suspension viscosity.
An equation for the effective buoyancy force contains two terms

fb-qkp l + r b ~-~-+w. V.+ ~-~-+w. w


(13)
2 o~pp pfa 2
rb = -~ e p f + r pp u f M

The first term within the curly brackets describes a contribution to buoyancy
caused by the suspension being under action of an external body force field that
is characterized by potential O. This contribution is sensitive to changes in this
potential with time and (or) in space. This may be important for a centrifugal
force field that influences flow of suspensions rotating with either varying or
constant angular velocity. However, this contribution does not evince relaxation
if the specific acceleration of the external field is invariable, as is the case with
gravity. The second term in equation (13) represents a buoyancy force
contribution stipulated by an additional effective body force field that makes its
appearance owing to particle inertia. It is worth noting that an additional term
describing a relaxation effect associated with this buoyancy contribution makes
its appearance [10]. This term can be shown, however, to be of a higher order of
1246

magnitude that other terms retained in equations (9) - (13), and for this reason, it
must be omitted from equation (13).
At v = 0 and in the dilute limit ~ ~ 0, force (8) can be easily demonstrated
to reduce to the force that is experienced by a single Stokesian sphere moving in
an unbounded fluid and that was previously evaluated in [11 ].
To get a deeper insight into the nature of the inertial phenomena that affect
interphase interaction, we also reproduce an equation of motion for a single
suspended sphere that can be straightforwardly deduced from equations (3), and
(8) - (12). This equation can be written down in the following form

dw d ( w - v) dv
mp ~ = - m a +m- +R (14)
dt dt dt

Here, operator d/dt is understood as that of full time differentiation along the
averaged particle trajectory (or alternatively, along the dispersed flow
streamlines). Vector R stands for a sum of all forces experienced by the sphere
except for those proportional to the full time derivatives of the continuous and
dispersed phase velocities (that is, except for the inertial force (12) and the
second contribution to the buoyancy force (13)). The following effective masses
are introduced:

m = c r n f + q~mp , ma = k m f ,
1/
k =-~ e + (15)
pf

where mp and my are the sphere mass and the mass of the fluid replaced by the
sphere, respectively. In the dilute limit, m~ tends to m/2, and m tends to mf, so
that equation (14) with coefficients (15) reduces to the well-known equation for
unsteady motion of a body in an unsteadily flowing unbounded fluid [11, 12].
The first term on the right-hand side of equation (14) describes a combined
effect of the virtual suspension mass acceleration and of the frequency dispersion
phenomenon. Coefficient k can be conventionally interpreted as an effective
added mass coefficient. It depends not only on suspension concentration, but
also on the ratio of particle and fluid densities. This is due to the fact that the
suspension cannot apparently be regarded as a truly homogeneous effective
medium that surrounds an accelerating suspended particle when evaluating the
effective suspension mass that is carried along with this particle.
The second term in the right of equation (14) is due to the inertial force that
makes its appearance in the suspension momentum conservation equation when
this equation is formulated in a coordinate system that moves with velocity v, as
1247

is explained in [12]. This force must not be dependent on the suspension being a
two-phase system. Accordingly, coefficient m that appears in this term
represents the mass of the suspension that is replaced by a sphere, as if the
suspension were a homogeneous one-phase medium.
The determination of the added mass coefficient was a predominant subject of
many papers, beginning with that by Zuber [13], representative examples of
which have to be found in [14-18]. Nonetheless, conclusions inferred in these
and other papers on the subject with respect to the added mass coefficient can
hardly be unambiguously compared to our formula for k as listed in equation
(15). This is due to the fact that practically all such conclusions commonly refer
to a system of particles (most usually bubbles) immersed into a potential or
almost potential flow, where the frequency dispersion of viscosity and the
ensuing stress relaxation is of no consequence. At the same time, the contribution
to k appears to be quite significant for fine particles in viscous flow, and it
comes about as a result of the stress relaxation and which is not taken into
accotmt in the majority of available papers. (Admittedly, this contribution
correspond to the term in equation (15) that is proportional to effective density
p~ involved in the definition of relaxation time T, in accordance with equation
(5) .) At any rate, the added mass coefficient as detennined by equation (15)
increases with ~ considerably faster than similar coefficients calculated in [13-
18]. This point deserves attention in the future work.
It is significant that it is the mean suspension density, but not that of the pure
ambient fluid, that is involved in equation (13) expressing the effective buoyancy
force. As has been discussed in [10], this inference brings to an end a recent
rather hot debate on which density must be used in the mentioned context while
treating fluidized beds and other suspension flows (examples of different
controversial arguments used in this debate can be found in [19, 20]).
For future reference, we write down an expression for the sum of inertial and
buoyancy forces for a suspension in the gravity field. We have

fJ +fb = r +W. v+kq~pf -69t


- + w-V / (v - w) - Cpg (16)

g being the gravity acceleration.


In conclusion to this subsection, we enumerate main assumptions made when
deriving the above constitutive equations. First of all, as has already been
indicated, we 1) have utterly ignored the impact of possible random fluctuations
of both suspended spheres and ambient fluid, and 2) have addressed only
suspensions of low or moderate concentration in flow of which the non-
overlapping property of hard spheres may be justifiably overlooked.
1248

Besides, certain other simplifications have been made in the calculation,


which can be summarized as follows.
1. The effective stress tensor was proven in [6] to contain a contribution that
depends on the angular velocity of sphere rotation in shear flow. Such a
contribution has been ignored, which is permissible to do if the strong inequality
holds true [7]

(Pp/Pf )(a2y/vf ) <<1 (17)

y being a characteristic mean shear rate.


2. Only moderately unsteady flows have been considered [10], meaning that
another strong inequality is valid that is akin to that in equation (17),

(18)

co being a characteristic flow frequency. In practice, inequality (18) is


commonly not at all restrictive, except for flows generated by sound and shock
waves. A generalization of the above constitutive relations to such wavy high-
frequency flows has been recently contemplated in [21 ].
3. At last, the length scale over which averaged quantities, and in particular
the suspension concentration, vary significantly has been assumed to be much
larger than radius a of suspended spheres. Moreover, a possible contribution to
the mterphase interaction force proportional to the suspension concentration
gradient has been completely left out of account, despite the known fact that, in
the generalized case, such a contribution must undoubtedly be included in the set
of the relevant vectors describing the local situation near the test sphere [22].

3.2 Other Models


Within the framework of the effective medium model presented, relative
suspension viscosity M is defined in equation (4). This expression for M was
more than once derived by different theoretical means, and in particular in [23].
It diverges as the volume particle concentration approaches 0.4.
Correspondingly, this causes other suspension properties to diverge as well. This
divergence is due to the fact that we have allowed the suspended hard spheres to
overlap, so that the local volume concentration of particles near the test sphere
is not influenced at all by preassignmg this sphere position (curve 5 in Figure 1).
This can hardly be looked upon as a reasonable approximation for highly
concentrated suspensions. Thus, to expand our model into a region of higher
concentrations, we have to consider more sophisticated representations for the
1249

pair distribution function of suspended spheres which prohibit sphere


overlapping. Some such representations have been discussed in [6], and also
used in [24] to numerically calculate the relative viscosity of suspensions and
emulsions as it manifests itself in steady shear flows.
First of all, the mere fact that suspended spherical particles do not overlap
results in the appearance of a "forbidden" sphere of radius 2a concentric with
any given particle into which centers of other particles cannot penetrate. Then, it
seems quite natural to approximate the pair distribution function as a step
function. This function identically equals the distribution function for one particle
(which coincides with the particle number concentration) everywhere outside the
forbidden sphere, and it turns to zero inside this sphere. (Such an approximation
can be proven to correspond to neglect of steric interactions in groups of
particles containing more than two particles.) This pair distribution function
corresponds to a local particle volume concentration in the vicinity of any
suspended sphere that monotonously increases with the distance r from the
sphere center (curve 3 in Figure 1). The local volume concentration equals zero
at the sphere surface, and it reaches r as the said distance becomes three times
as large as particle radius a. Correspondingly, the effective viscosity and density
of the fictitious medium the test sphere is assumed to be immersed into smoothly
change from their values specific to the pure ambient fluid to the effective
viscosity and mean density of the suspension as a whole [6].
Now, let us approximate the local volume concentration near the test particle
as a step fimction which is equal to zero within a spherical layer a < r < z a and
r at r > za, 1 < Z < 3. Then we automatically arrive at a model according to
which the test particle is separated from a homogeneous fictitious medium by a
layer having a thickness (Z-1) a that is filled with pure fluid. The fictitious
homogeneous medium is again characterized by the same properties as the
suspension, whereas properties inside the concentric layer are those of the
ambient fluid. A corresponding step-wise local concentration function is shown
at Z = 2 by curve 4 in Figure 1. Models of such a kind were also repeatedly used
in numerous works on effective properties, such as viscosity, elastic moduli,
effective thermal conductivity and diffusivity, of suspensions, emulsions and
composite materials.
Let us consider the next logical step in approximating the local concentration
variation in the vicinity of the test sphere of a dense suspension. This variation
detenmnes also the properties of the fictitious medium this sphere is assumed to
be immersed into [6]. This step may involve using different expressions for the
pair distribution function that have been developed in the statistical physics from
various approximate models of an assemblage of identical hard spheres, such as
that by Kirkwood, different versions of Percus and Yevick model, hyperchained
1250

model, etc. The local particle volume concentration distributions that stem from
two of such models at different values of mean suspension concentration are
plotted in Figure 1 as well.

1 5

0.5 4

0 !

1 2 3 4 r/a
Figure 1. Local particle volume concentration near test sphere surface related to
mean concentration as a function of relative distance from this sphere center; 1,
1" - model by Percus and Yevick at # = 0.4712 and 0.2612, respectively; 2, 2" -
model by Kirkwood at # = 0.484 and 0.2314; 3 - smoothed distribution
corresponding to neglect of interactions of more than two spheres; 4 -
approximation of the last distribution by a stepwise function; 5 - spheres are
permitted to overlap.

Figure 2 illustrates the concentration dependence of the relative suspension


viscosity in conformity with the models used to approximate the pair distribution
function in Figure 1. The curves presented in Figure 2 have been calculated by
Yendler [24] who also compared them with experimental data on steady-flow
suspension viscosity as available by the late seventies. According to his
conclusions, function M as defined in equation (4) gives satisfactory results up to
~ 0.2, the smoothed distribution is good enough up to # z 0.3, and the curves
that correspond to either Kirkwood or Percus - Yevick model may be used
through # z 0.4 - 0.5. Similar conclusions have been driven in [24] with respect
to the relative viscosity of emulsions, the viscosity ratio of the phase materials
being arbitrary.
Supposedly, the model of suspension viscosity in steady flows must become
somewhat dubious at very high concentrations. This is largely due to two
reasons. First, occasional direct contacts of suspended spheres, or at least the
formation of very thin lubrication films of intervening fluid separating pairs of
1251

spheres that come closer together, are hardly avoidable in highly concentrated
suspensions, even if there are no interparticle collisions in the generally accepted
sense. Secondly, the inner structure of a flowing highly concentrated suspension
must be especially sensitive to various structure-ordering processes, up to those
of the formation of crystalline patterns. This means that the actual pair
distribution function as it establishes itself in dense suspension flow must be
essentially different fi'om its approximations resulting from equilibrium statistical
models, and it must depend on flow parameters. In particular, one can expect the
occurrence of shear-thinning even in flow of a suspension composed of identical
hard spheres without molecular colloid interaction, if suspension concentration is
large enough.

M ....., , ,/;
6

4 5

I __ !

0 0.2 0.4

Figure 2. Relative suspension viscosity as a function of particle concentration by


volume according to different approximations for the pair distribution fimction
that are illustrated in Figure 1; notation is the same as in Figure 1.

Both the aforementioned points cannot in principle be addressed within the


framework of the model being developed because we have from the very
beginning completely neglected 1) forces that may arise as neighboring particles
come to a contact, and also 2) the flow influence on the pair distribution
function. This necessitates making use of some other independent, either
theoretical or empirical, representation for the relative viscosity of highly
concentrated suspension that does not emerge from our theory.

3.3 Provisional Approach


Unfortunately, constitutive equations for unsteady suspension flow have been
considered so far only for moderately concentrated suspensions where the non-
1252

overlapping property of suspended particles may be ignored [10]. Since the


relative suspension viscosity is determined by the formula given in equation (4)
which is approximately valid merely at ~b< 0.2 and which diverges at ~ ~ 0.4,
this leaves us with a severe problem how to generalize these equations to
suspensions of higher concentrations. To do this, we notice that all the
constitutive equations as listed above appear to be fully determinate, granted that
the only function, that of relative suspension steady-flow viscosity, is known.
Correspondingly, we may hope to get a supposedly reasonable evaluation of
these equations in unsteady flow of dense suspensions on purely semi-empirical
grounds. Namely, we may retain the concept of an effective fictitious medium
surrounding any suspended particle, but on condition that another suitable
formula for the relative suspension viscosity is used. A simple example is
provided by the function suggested in [10]

M = (1- ~b)-5/2 (19)

which can be proven to be sufficiently adequate up to # = 0.3. At any rate, this


function satisfies the Enstemian limit M oc 1 + 2.5 # at # << 1, and it behaves
in a correct manner as # increases.
By way of example, we shall illustrate the dependence of the added mass
coefficient on both suspension concentration and density ratio of the phase
materials for a suspension whose relative viscosity is determined by equation
(19). In this case, we obtain from equations (4), (15) and (19)

l+ , (20)
2 2 ,of

For suspensions composed of particles that are lighter than the suspending
fluid, and in particular, for suspensions of gaseous bubbles in a liquid, the added
mass coefficient is illustrated by curves in Figure 3, a. This coefficient proves to
be a rapidly increasing function of suspension concentration, and this function's
recline slightly increases as the density ratio grows. For reasons that have been
pointed out in Subsection 3.1, this recline is considerably larger than those in
other existing models that treat particles in a potential flow. To show this, the
added mass coefficient as follows from the Zuber's model [13]

k=l 1+2~ (21)


21-~b
1253

as well as the expansions of this coefficient valid for dilute suspensions

k = _1(1 + 2.76~b), 1 (1 + 3.32~) (22)


2

that were obtained in [15, 16], respectively, are also shown in Figure 3, a.

12 a. . . . . . 6 b

- , , , , , , , ,

0 0.2 0.4 q5 0 0.2 0.4 ~b


Figure 3. Dependence of added mass coefficient k = ma/mf on ~b for
suspensions whose particles are lighter than the ambient fluid (curves 1 on a
correspond to tc = pp/pf = 0 , 0.5 and 1, in ascending order), and similar
dependence of k ' = mdmp for suspensions with heavier particles (curves on b
correspond to K= 1, 2, 5, 10 and 0% in descending order); curve 3 presents the
formula by Zuber [13], and curves 4 and 5 give the added mass coefficient for
dilute suspensions according to [15] and [ 16], respectively.

It is worth noting in this connection that conclusions following from known


works on the virtual fluid mass of spheres suspended in a potential flow are
sensitive to the shape of the final sphere velocity distribution attained as a result
of the action of the impulsive forces that are assumed to generate the sphere
acceleration [17,18]. In contrast to this, equation (20) does not altogether involve
any assumption of this kind. It holds true in the general case of an arbitrary
moderately tmsteady flow that is supposed to satisfy inequalities (17) and (18).
If the density of suspended particles is larger than that of the ambient fluid, it
seems more judicious to relate the suspension additional mass to that of a sphere,
but not to the mass of fluid replaced by the sphere. Dependence of k ' = mdmp
on ~b at values of 1r that are larger than unity is illustrated in Figure 3, b.
1254

A couple of important effects that seem to be specific to gas-particle


mixtures, as well as to suspensions of very heavy particles in light liquids, are
worth mentioning. First, coefficient k" tends to decrease in a dilute or
moderately dense suspension when the suspension concentration increases.
Secondly, for suspensions of particles in a gas, this coefficient may even become
negative in a range of small and moderate concentrations. This means that the
acceleration of a single particle in such a suspension is not at all hampered by
any virtual mass of the suspension that is carried along by the particle, but rather
the surrounding suspension motion facilitates the particle acceleration.
Dependence of k" on # at large ~c that illustrates these effects is demonstrated
by the curves plotted in Figure 4.

0.1

b r 0.2
Figure 4. Variation of the virtual suspension mass for moderately concentrated
suspensions of heavy particles; the curves correspond, in descending order, to ~r
= 10, 30, 100 and 1000.

This conclusion is by no means connected with our having used approximate


equation (19) for the relative suspension viscosity, within the scope of the semi-
empirical provisional approach as proposed in this subsection. It also holds true
if the expression for M is used that results from the theory in [10], without
having invoked any additional empirical assumption and that is listed in equation
(4). It seems rather tempting to specially address this issue in more detail in
future work.
No matter how the relative suspension viscosity has been defined, the
interphase interaction force as expressed by equation (8) proves in the
generalized case to be insuiticient to ensure the dispersed phase momentmn
conservation equation (3) to be satisfied for any monotonous suspension
concentration distribution. For instance, let us consider the lateral component of
this equation in an arbitrary uni-directional steady suspension flow. It is not
1255

difficult to see from equations (9) - (13) that merely the buoyancy contribution to
the lateral component of f differs from zero in such a flow, so that the wanted
equation component reduces to a very simple form

where gt stands for the lateral component of the acceleration associated with an
external field of body forces that act on the suspension. It is easy to see that
equation (23) can never be fulfilled except for ~ - 0, if the suspension is not
neutrally buoyant, and if gt differs from zero.
Hence a conclusion seemingly follows that the steady uni-directional
suspension flow under consideration is impossible in the sense that the flow
domain must always separate into two regions. As has been already mentioned in
Introduction, one of these regions should be filled with pure fluid, where ~ = 0,
and equation (23) transforms into an identity. The other region should be
occupied with close-packed particles, where equation (23) is irrelevant because
certain direct interparticle forces arise which must be explicitly included in the
dispersed momentum conservation equation. Obviously, this conclusion is wrong
since it contradicts numerous experimental observations, if for no other reason.
In practice, the regular force that appears on the left-hand side of equation
(23) causes a particle convective flux in the direction of this force. Such a flux
modifies any initial spatial distribution of suspended particles. However, it not
necessarily leads to complete flow stratification with the formation of a region
totally devoid of particles, and of another region containing close-packed
particles. Indeed, the particles experience random fluctuations of different
physical origins. Particle displacements due to the fluctuations must result in the
occurrence of a specific particle migration, and a net particle flux stipulated by
migration must oppose the convective flux. Supposedly, the combined action of
these two fluxes prevents, or at least hampers, the flow domain separation into
the regions mentioned, and so it can result in establishing a steady suspension
concentration distribution. It is such a steady distribution that has actually to be
dealt with when considering steady flows.
If the particles undergo successive displacements that can be regarded as
mutually independent, the migration flux may be appropriately described as a
diffusion flux down a concentration gradient. It is rather tempting in such a case
to replace the dispersed phase momentum conservation equation by a convective
diffusion equation for suspended particles, as has been done more than once in
various papers on suspension flow. However, as will follow from the subsequent
analysis, this can be formally done only for one-dimensional suspension flows, in
which the suspension concentration depends on a single coordinate, and where it
1256

is only the corresponding lateral momentum conservation equation component


that has to be substituted by the diffusion equation. It is quite clear that a scalar
diffusion equation can never be substituted for the vector momentum
conservation equation in an arbitrary flow. This means that the impact of the
fluctuations on suspension flow must be treated in terms of additional forces or
stresses that should be incorporated in the momentum conservation equations,
but not in terms of fluxes. This point will be discussed in more detail later on.
In what follows, we shall consider 1) particle fluctuations caused by different
physical mechanisms, and 2) the origination of additional normal stresses in
suspension flow due to the fluctuations. For evident reasons, we start with
studying the influence on suspension rheology of thermal Brownian fluctuations,
with a mind to afterwards generalize some results obtained from such a study in
order to describe effects caused by hydrodynamically induced fluctuations.

4. ASSESSMENT OF THERMAL FLUCTUATIONS

There are great many papers that deal with the problem of Brownian motion
in a system of many interacting suspended particles, as well as with the
elucidation of the influence that this motion is likely to cause on the suspension
rheological properties. Recent examples are provided by the beautiful papers by
Brady [25, 26] who has successfully addressed the statistics of a crowded
assemblage of Brownian spheres when placed into a thermal bath characterized
by temperature kT in energy units, and also the Brownian motion effect on the
rheological behavior of concentrated suspensions. However, within the
framework of the present treatment, we are surely unable to reproduce
conclusions of the type of those drawn in [25, 26]. This is due to the fact that we
have given up analyzing the influence of both suspension flow and particle
fluctuations on the pair distribution function, if for no other reason. Thus, here
we confine ourselves to a more modest aim of obtaining simpler conclusions
pertaining only to normal stresses caused by the thermal fluctuations. Within the
scope of our treatment, the effective suspension viscosity is regarded as a given
fimction of suspension concentration, no matter whether it is influenced or not by
particle Brownian motion, or either by other physical factors that happen to
originate a short order in the suspension flow.
As an alternative to ingenious methods used in [25, 26] and in other papers of
the same type, we are going to resort to the method that is based on introducing a
"thermodynamic" constituent of the interphase interaction force and that was
thoroughly substantiated by Batchelor [27]. When applied to suspensions of
interacting particles, this method represents a certain generalization of the
classical thermodynamic force method devised by Einstein for random migration
1257

of a single particle in an unbounded fluid due to the particle interaction with fluid
molecules.
As has been stated in [28], the influence of Brownian motion on suspension
statistics and rheology consists in smoothing out probability density function for
suspended particles, in contrast to a tendency of suspension flow to make these
functions non-uniform in their respective phase spaces. This influence is partially
revealed through mutually interdependent thermodynamic forces experienced by
different particles. They affect mean stresses in suspension flow in two ways.
First, they cause an immediate contribution to these stresses. Secondly, they
change the probability density functions, and in particular, the pair distribution
function, thus implicitly affecting the stresses (and the effective suspension
viscosity).
Before proceeding further, we wish once again to outline some limitations
inherent in the present paper. For reasons given above, below we concentrate
only upon the stress contribution due to the first mechanism, and we make use of
the analysis earlier forwarded in [27] for dilute suspensions and generalized to
concentrated suspensions in [1]. However, it is worth noting that the failure to
allow for the changes in probability density functions results in a slight difference
of the resultant relative suspension viscosity from its representations found on
the basis of more sophisticated schemes, even for suspensions of low and
moderate concentrations. Thus, the coefficient at q~2 in the Taylor expansion of
function M as defined in equation (4) equals 25/4 = 6.25, whereas the same
coefficient was found to be equal to 6.2 for shear suspension flow dominated by
intensive Brownian motion [9], and to 7.6 for flow of pure elongation without
Brownian motion [8].

4.1 Thermodynamic Force


Random migration of a particle in a suspension may be considered as a
diffusion-like process if two conditions are satisfied. First, the migration time
scale must be much larger than the duration of independent displacement steps
made by the particle in succession, and this scale must, in its turn, be large as
compared to a viscous time az/vf that characterizes propagation of local
disturbances caused by the particle over distances of the order of particle radius
a [28]. Secondly, the configuration of neighboring particles surrounding the
particle under question must remain approximately invariable during time that it
takes for this last particle to accomplish a large number of successive
displacement steps, and for the particle diffusivity to attain its asymptotic value
[27]. The former condition is needed because the particle motion is non-
Brownian over times smaller than the viscous time, due to the transient
development of the flow field near the particle. The latter condition is needed to
1258

approximately regard the fluctuating force, that is exerted on the particle by the
surrounding fluid and that depends on positions of all the neighboring particles,
as a stationary random process.
To begin with, we shall briefly reproduce the reasoning employed in [27]. If a
suspended particle is at equilibrium when an external force field is applied, the
Boltzman distribution must be valid, and the particle Gibbs free energy must be
spatially uniform. This means that an external force acting on the particle must be
statistically compensated by the thermodynamic force Ft which is equal to the
particle chemical potential gradient taken with a negative sign. The
thermodynamic forces produce a particle convective flux that is precisely equal
to the actual diffusion flux. Furthermore, the motion of the particle due to the
external forces and to its diffusion caused by thermal fluctuations can be
justifiably presumed to result in additive particle displacements. For this reason,
the particle diffusion flux must be determined by exactly the same
thermodynamic forces even under non-equilibrium conditions where there are no
external force [27].
By its very definition, the thermodynamic force in an isothermal flow can be
expressed in a general form as

(oPPl Vrt -10PPl Vp (24)


F, = -k---8-;, J p,r :P J n,T
where p P is the particle chemical potential. In accordance with equation (24),
the thermodynamic force consists of two parts, the second one representing the
buoyancy force. Since the buoyancy force has been already accounted of in
equations (13) and (16), there is no need to retain it here. Hence, we are left with

(25)
On p,T p,T

As shown in [27], a similar force

V~ = _n_n_Ft (26)
F/=- On JP, f ~, Oqk ) p,T no

acts also upon each molecule of the ambient fluid, where flf and no are the
chemical potential and the number concentration of the molecules. It follows
1259

from equation (26) that the thermodynamic forces act simultaneously on both
phases of the suspension, and moreover, that their sum is zero for any suspension
volume, that is,

nF t + noF [ - 0 (27)

Equation (27) proves that the thermodynamic force per unit volume, ft = nFt,
has to be regarded as an additional constituent of the interphase interaction force.
In fact, equation (27) is a manifestation of the Gibbs - Duhem relation written out
for colloid particles dissolved in a fluid.
The chemical potential of hard spheres undergoing the Brownian motion can
be calculated with the help of the standard methods of statistical mechanics. The
calculation is trivial for very dilute suspensions of non-interacting spheres where
~tp coincides with the chemical potential of an ideal gas, in which case

~tp = bt,p + kT(ln 0 -~b) (28)

the first term on the left standing for the inner chemical potential of a sphere.
If a suspension is dilute, but the mterparticle interaction is taken into account,
the application of the well-known group expansion technique seems to be
natural. The latter leads to an expression for the chemical potential in the form of
a series in powers of q~ that resembles the standard virial expansion~ Each term
of this series is associated with the collective interaction between the particles
being members of various possible groups composed of the fixed numbers of
particles. Allowance for merely pair interactions conforms to retaining also the
second virial coefficient of the expansion. This gives rise to the term of 8q~kT in
the series for ktp, which has been previously calculated in [27].
The group expansion technique is unlikely to be of much use, however, when
applied to crowded assemblages of suspended spheres, since this technique does
not in fact give a tangible opporttmity to advance in a region of large
concentrations. In such a case, the calculation is handicapped by the fact that no
reliable rigorous statistical physical theory of dense gases and liquids has been
elaborated to date, so that an approximate model has to be used. It seems
sensible to make use of one of versions of the Pereus- Yevich theory of liquids
because it makes it possible to obtain analytical results for assemblages
composed of hard spheres. Following [1 ], we employ the version of this theory
worked out by Carnahan and Starling [29], which gives the following equation of
state for a dense gas of identical rigid spheres
1260

PV = NkTG(qk), G(qk) = 1 + + ~2 _ ~3 (29)


3

where P is the osmotic pressure, and V is the volume occupied by the system
of N rigid spheres.
Roughly speaking, equation (29) brings to light corrections to the classical
equation of state for an ideal gas that are due to sterie interactions between
spheres having fufite volume of their own. It should be pointed out once again,
however, that this equation fails to account for the formation of a crystalline
phase in dense sphere assemblages. This failure calls in question applicability of
this equation of state to suspensions of high concentrations that are close enough
to a hypothetical concentration associated with the close-packed state.
Equation (29) serves as a sound basis while evaluating the particle chemical
potential and calculating the thermodynamic force for spheres suspended in a
fluid that plays the role of a solvent. The difference of the wanted chemical
potential from that of the same spheres forming a dilute system can be quite
straightforwardly written down as follows:

A,u p = - k T dln QN (30)


ON

where QN is the configurational integral of the spheres divided by V N.


Using equation (29) and an alternative representation of the equation of state
in terms of the derivative of the configurational integral over volume yields

O I n ( V N QN)__~___P _ N G(~)
(31)
cTV kT V

Solving this equation results in [1]

lnQN = - N d ar
0
r (32)
=- fG(~)- 1 G(O)-I ( c7 _/
ON ,I -~ d ~k- N ~ p , T ,N o
p,T,No o

and this determines the chemical potential difference as defined by equation (30).
1261

Note that the differentiation is performed in equation (32) under the condition
that not only p and T, but also the total number No of solvent molecules are
maintained constant. By allowing for the expression of the osmotic pressure
function as shown in equation (29) and for evident relations

aN (O-N-/ = ~(1-~) (33)


r cr N + cro N o ' p,T, No N

where o- and o0 are the specific volumes of one particle and one molecule,
respectively, we derive from equation (32)

cTlnQN) _ 8 - 50
ON - - (---i~ 0 (34)
p,r, No _~)2

Hence and from equations (28) and (30) follows the resultant formula

8 - 5______L~
/u p - 1.1,p + k T F ( q k ) , F(~b) - l n 0 - r + (1 - ~)e ~b (35)

which finally determines the thermodynamic force as defined by equation (25).


The third term on the right-hand side of equation (35) is due to the steric
interaction of hard spheres having a finite volume.
Determination of the thermodynamic constituent of the mterphase interaction
force requires thermodynamic force Ft be multiplied by the particle number
concentration that is expressible, in its turn, as ~b divided by o-. Thus, this
constituent can now be formulated in the following form:

r, = , 1~, _- --0 l~, = - kZ 0 a ( ( r v 0 (36)


cr cr d~b

Equation (36) brings to completion the calculation of the thermodynamic


mterphase interaction force constituent acting in a non-uniform Brownian
suspension. The convective flux caused by this force constituent must identically
equal the diffusive flux down a suspension concentration gradient, and so it can
be substituted for this diffusive flux. To do this, the thermodynamic force
constituent has to be incorporated into the total interphase interaction force
alongside the other parts of this force listed in equation (8).
1262

4.2 Normal Stresses Due to Thermal Fluctuations


Next, we shall consider an alternative representation of the thermodynamic
force that act on all particles in a unit suspension volume. This representation
proves to very useful in our subsequent analysis. To start with, we reduce
equation (36) to another form by means of introducing a new function of
suspension concentration, L(~, that satisfies the following differential equation:

d = dF(~b)
a---g- (37)

where F(r is the function defined in equation (35). Integrating equation (37) at
an evident initial condition yields

-
2 In(l-O) +3 1-20 +
1 15-8~-~ 2
- (38)
r 1-0 2 (1-~) 2

Using equation (38) enables us to reformulate equation (36) as

-f t = kT d
cr d e [r L(r V~b- - V H , H - kTo.r L(~) (39)

It is clear that the scalar quantity H may be identified with the effective
isotropic normal stress that is caused by the particle thermal fluctuations and that
manifest itself in mean flow of the dispersed phase. Physically, it describes a
momentmn flux density produced by the particle thermal fluctuations, and by its
very definition, such a density must be interpreted as a stress.
Instead of adding the thermodynamic constituent to interphase interaction
force (8) to describe the impact of Brownian motion on suspension flow, we can
alternatively include the isotropic particulate stress H into momentum
conservation equation (3) of the dispersed phase. Simultaneously, we must
include a component of - FI to mean interstitial pressure p involved in
accordance with equation (5) in the defirfition of the effective stresses that appear
in momentum conservation equation (2) for the continuous phase. This amounts
to introducing the following stress constituents that are caused by Brownian
motion and that must be added to the stresses acting in a suspension flow without
fluctuations:

c~f=HI, c~=-HI (40)


1263

/2

0 0.2 0.4 O
Figure 5. Functions of suspension concentration that determine the osmotic
pressure of a pseudo-gas of rigid spheres and the normal stress that affects
dispersed phase flow of a suspension of the same spheres without collisions and
that is due to thermal fluctuations.

It is significant that function L(r is different from the osmotic pressure


function, G(~, introduced in equation (29), as follows from the curves plotted in
Figure 5. Correspondingly, normal particulate stress H that arises in flow of a
suspension of collisionless spheres differs from the osmotic pressure for a
pseudo-gas of the same hard spheres which are presumed to exchange
momenttma and energy by direct collisions. Apparently, this difference has to be
attributed to the distinction between the spheres that form a pseudo-gas and the
spheres that are suspended in and interact through a fluid. As evidenced by our
calculation, this distinction proves to be significant. The presence of the fluid that
plays the role of a solvent considerably affects statistical properties of the
spheres that resembles molecules of a solute, rather than molecules of a dense
gas. It is hardly surprising therefore that the thermodynamic forces as calculated
for a pseudo-gas and those for a solution of hard spheres happen to be different
as well.
Since kT identically equals the mean kinetic energy associated with one
degree of freedom of a single sphere, the expression of H in equation (39) can
be rewritten in terms of the variance of any i-th component of the particle
velocity.

(41)

Here w' is understood as a particle fluctuation velocity that comes about as a


result of isotropie thermal fluctuations and that actually represents the difference
1264

of the actual random velocity of a particle from the averaged velocity of the
dispersed phase.

4.3 Some Generalizations


Equations (40) and (41) can in principle be used on semi-empirical grounds to
describe normal stresses originated by particle fluctuations of another physical
origin. With this purpose in view, they are be reformulated in the form

8I = fi, 8p =-fi, fi- ppr (42)

where an asterisk denotes the operation of diadic multiplication. In the case of


isotropic Brownian motion, (w',w') -= (w~2) I, and the formulae in equation (42)
reduce to those in equations (40) and (41).
Equation (42) results actually from a quite plausible hypothesis that the stress
tensor originated by particle fluctuations must be proportional to the only tensor
available that characterizes the fluctuation intensity and that is quadratic in
fluctuation velocity components. Moreover, there is altogether no reasons to
expect that the scalar coefficient as defined in equation (42) is sensitive to the
cause that drives particles to fluctuate, provided that the particles 1) behave
themselves as statistically independent entities, and 2) do not come in direct
contacts with one another in consequence of their fluctuating motion. Thus, we
can supposedly retain the same value of this coefficient as that fotmd when
investigating thermal fluctuations. However speculative, the hypothesis under
discussion seems to be credible enough to offer a real opportunity of evaluating
normal stresses that are originated by hydrodynamically induced fluctuations of
statistically independent collisionless particles.
In conclusion to this section, we are going to propose, also on semi-empirical
grounds, a generalization of the above results to flows of very dense suspensions
whose concentration approaches that of close packing.
Formally, equation (40) can be rewritten in terms of a so-called Enskog factor
Z that is extensively used in the statistical physics

I-l=nmp[1+ 4~Z(~)] (w/2), Z(~)- L(~)- 1 (43)


4~
Roughly speaking, the physical meaning of this equation is as follows.
Frequency of interactions increases in a crowded particulate assemblage, and
momentum transport inside hard spheres is carried with a practically infinite rate.
The Enskog factor describes the influence that these effects are anticipated to
produce on averaged momentum transfer in a concentrated suspension.
1265

In flow of highly concentrated suspensions, particles are likely to come in


frequent contacts one with another, and the implicit mechanism of interparticle
exchange through intervening fluid may be expected to give way to a direct
exchange mechanism carried out by particle contacts. It is natural to expect in
such a case that function L(r must gradually tend to osmotic pressure function
G(~) defmed in equation (29) as the suspension concentration increases. An
equation for the corresponding Enskog factor then reads

Z(~b) = G(~b)- 1 _ 1-0.5_______~ (44)


4r (1-~) 3

Making a precise choice between these extremes requires a careful


examination of both particle contacts and long-range hydrodynamic interactions
in groups of many particles, and so presents a formidable task. As far as the
authors know, this problem has been rigorously addresses only in connection
with the dynamics of dilute gas-solid mixtures by Koch [30]. Fortunately enough,
the difference between functions G(~) and L(~) as evidenced by Figure 5 is
not too large, especially so for dilute and moderately concentrated suspensions.
For this reason, there is no urgent practical need to discriminate between these
functions.
As has already been mentioned, the model by Carnahan and Starling [29] can
hardly be applied to the description of the momentum transport augmentation in
flow of a highly concentrated suspensions whose concentration approaches that
of close-packed state. Therefore, in order to get a rough idea about flow of such
a suspension, we can make use of the expression for 1-I listed in equation (43),
in which the Enskog factor is empirically approximated from a suitable model of
very dense gases. In particular, we can employ the approximate "geometrical"
model of dense gases as developed by Enskog [31]. The Enskog factor is then
expressed as

4~ , GE(~)= 1 - (45)

where ~ is understood as a maximal particle concentration by vohune that is


attributed to a hypothetical close-packed state. The exact value of this
concentration is unknown, and it has to be regarded as an empirical constant.
When formulated in terms of the Enskog factor, the expression for the normal
stress tensor as advanced in equation (42) reads
1266

fI - nmp [1 + 4r w')= pp r [1 + 4r (w',w') (46)

Depending on concentration of a given suspension flow, different


relationships may be used for the Enskog factor, and in particular, those listed in
equations (43) - (45). In what follows, we shall check the equations mentioned
by applying them to the determination of particle distribution in a simple steady
Couette flow of a suspension of identical neutrally buoyant spherical particles
between concentric rotating cylinders.
If suspension concentration is sufficiently high for contacts between
suspended particles be essential, equations (44) or (45) has to be used for the
Enskog factor, instead of equation (43). In such a case, as before, equation (46)
describes the normal stresses that act in flow of the dispersed phase. However,
there is altogether no proof in this case that fluctuations cause the contributions
to the total stresses affecting the continuous phase flow that must precisely equal
the corresponding stresses as occur in the dispersed phase flow, when being
taken with a negative sign. Therefore, we are obliged to confine ourselves to
calculating only the dispersed phase normal stresses, whereas the said
contributions to the continuous phase stresses remain undetermined.
It should be emphasized in conclusion that both methods, which have been
developed above to describe the normal stresses produced by fluctuations of
statistically independent collisionless particles, are completely equivalent. This
means that either the thermodynamic constituent of interphase interaction or the
normal stresses as defined in equation (42) can be introduced into the momentum
conservation equations for flow of the suspension phases with equal success.
However, it seems wrong to incorporate both the said stresses and the
thermodynamic contribution to mterphase interaction simultaneously. Apart from
these stresses, Reynolds-like stresses may appear as a result of averaging of the
left-hand side of equations (2) and (3), as it has been shown in [32] when
studying stability of fluidized beds.

5. MODELING SHEAR-INDUCED FLUCTUATIONS

Particles of sheared suspensions are well known to exhibit random


displacements both across and in the direction of flow streamlines, due to
relative motion of particulate layers that move in parallel to one another with
different velocities. These displacements are illustrated in Figure 6. They
produce particle self-diffusion in shear flow of macroscopically uniform
suspensions, and also they cause the particles to migrate down a suspension
concentration gradient, if any. Shear-induced particle self-diffusion investigated
in [33], apparently for the first time. After that, it was studied in more detail in
1267

[34] and in some other papers. Shear-induced mutual diffusion of particles was
proven to induce the flow stratification, meaning that a non-tmiform particle
distribution develops in a shear flow as a result of the combined action of
external forces and particle mutual diffusion [35].
There are known many attempts to work out a comprehensive and reliable
model to treat suspension shear flows in general, and to determine particle
distributions specific to various concrete types of shear flow in particular.
Representative examples illustrating different trends of thought can be found in
[36, 37]. The authors of [36] developed a fluid dynamic model of shear
suspension flow, in which it was suggested to describe the shear-induced
migration in terms of a special particle diffusion equation introduced on purely
empirical grounds. In contrast to this, the authors of [37] deduced the steady
particle distribution in a pressure-driven one-dimensional flow in a channel from
the requirement that the total normal stress be constant in directions normal to
the mean flow. The model stemming from this requirement was well supported
by their numerical simulations performed on the basis of Stokesian dynamics
technique, and these simulations were, in their rum, in qualitative agreement with
experiment.
While setting aside a more detailed discussion of these issues until later, we
focus our attention on evaluating the normal stresses that are due to shear-
induced fluctuations within the framework of the approach being developed.
According to equation (46), this means that we have to calculate the tensor of
averaged particle velocity products, or particle velocity variance tensor (w'*w'),
in shear suspension flow as a function of suspension concentration, shear rates,
and particle radius. This has been done in [38], and we reproduce the main
results obtained in that paper.

5.1 Particle Velocity Variance Tensor


As follows from a simple consideration based on the theory of dimensionality,
and is also confimaed by conclusions drawn in [33, 34], random particle velocity
components must scale with ya, y standing for a characteristic shear rate value.
This proves that the particle velocity variance tensor must be quadratic in the
spatial derivatives of the dispersed phase mean velocity. Moreover, this tensor
must be independent of rotation of the suspension as a whole. This means that its
components may depend on those combinations of such derivatives which have a
bearing on the deformation rates in the dispersed phase, but not on combinations
that describe rigid-body rotation of this phase. As such, the desired tensor is
solely dependent on tensor ew that describes strain rates for the dispersed phase
flow. In other words, the wanted tensor depends exclusively on the symmetric
part of the spatial derivatives tensor for the true deformation rates mentioned
1268

above. To obtain a concise expression for this particle velocity variance tensor,
we make use of some tensor covariance methods as characteristic of the rational
mechanics.
The most general rational mechanical representation that satisfies both the
aforementioned requirements is of the form

( w ' * w ' ) = A e w .e w + BI2(ew)l , I:z(ew)=(1/2)ew'ew (47)

/2 standing for the second tensor mvariant, and A and B being scalar
coefficients independent of the dispersed phase mean velocity field. To find a
relation between A and B, and thereby to reduce the uncertainty inherent in
equation (47) to a single unknown scalar coefficient, an analysis of a simple
shear flow sketched in Figure 6 appears to be helpful.
y

1 ly wx--]~

Figure 6. A sketch of particle interaction in shear flow; explanation in the text.

Consider the situation in the vicinity of a spherical particle 1 of a dilute


suspension in a reference frame having its origin at the particle center. Another
sphere 2 that comes from infinity and that passes near the first sphere undergoes
a displacement in a plane containing both the second sphere trajectory and the
line connecting the sphere centers, as shown in Figure 6. The maximal
displacement occurs in the symmetry plane that is normal to the above indicated
plane, and at the same time, contains the first sphere center. The theory of
similarity suggests this displacement to be fully determined by an original aiming
distance r (the far upstream coordinate of the second sphere center) scaled with
particle radius a.
Let us define angle rp as the angle formed by the direction of shear (y-axis on
Figure 6) and a line that is drawn fi'om the reference frame origin (the point in
which the first sphere center is situated) to the point of maximal displacement in
1269

the plane of symmetry. It appears pretty obvious to assume from a simple


syrmnetry reasoning that the maximal displacement is independent of the said
angle, provided the relative velocity of the spheres being the same. Then this
displacement length is expressible as l = 1,(r/a),r/a being the relative aiming
distance. Displacement components along the y- and z-axes in the symmetry
plane equal ly=l cos (p and lz = 1 sin qg, respectively, and the initial relative
velocity of the particle centers is u = 7'a cos (p.
Next, let us focus attention on flow of a macroscopically homogeneous
suspension (~ = const) under a condition of uniform shear (7'-- const). In order to
evaluate averaged moduli of the first sphere's displacement components as
caused by interactions of this sphere with all the other spheres whose centers lie
within a cylinder r < R = a R, and which happen to intersect the symmetry plane
for a unit time, we need the help of the following integrals:

n'/2 R R,
{lyoclz
} ~dq9 fr dr(ycos~)Lsinqgjoc
Ic~176 {re~4}
1/2
fl * ( ~ 2 d na4ocr (48)
0 0

where the integral that stands in the parentheses is independent of any


dimensional parameters.
Since the averaged displacements along the y- and z-axes are attributed to a
unit time, they have to be regarded as convenient measures of corresponding
velocity components for shear-induced fluctuations. Note that a result similar to
that in equation (48) could readily be obtained from the theory of dirnensionality
because there is the only quantity of velocity dirnensionality, y a, that can be
constructed on the basis of available dimensional parameters.
Evaluation of the averaged displacement components as based on equation
(48) bears upon dilute suspension flow in which the chosen sphere's interaction
frequency with all spheres that are passing by is in fact presumed to be linearly
proportional to both suspension concentration and shear rate. This means that
particle fluctuation velocity components in dilute suspensions is actually
calculated by taking corresponding mean displacements of the order of a, which
occur on the average as a result of a single pair interaction, and then by dividing
them by the mean time that elapses between successive interactions of the
chosen sphere with its neighbors. By its very definition, the last time is equal to
the inverted pair interaction frequency.
The indicated physical meaning of equation (48) allows us to generalize this
equation to concentrated suspensions. In doing so, we avail ourselves of a semi-
empirical approach that has already been mentioned in brief in the preceding
1270

section. In a dense gas, as the particle volume fraction grows while the particle
number concentration is held constant, the collision frequency is known to
increase faster than is required by the last proportionality in equation (48). This
increase in collision frequency occurs due to the forcing the particles into a
confined space, and it is usually described with the help of the Enskog factor.
Although relative particle displacements in a sheared concentrated suspension
are affected by long-range hydrodynamic interactions involving many particles,
main contributions to the displacements may again be viewed approximately as
the result of particle pair interactions. As a first approximation, we hereupon
conceive elementary particle displacements as taking place in consequence of
two particles occasionally drawing close to each other, the influence of the other
particles being assumed to cause a secondary effect on pair interactions. This
amounts to viewing the complicated multiparticle interactions involving the
chosen sphere in a sheared concentrated suspension as a sequence of certain
effective pair interactions of this sphere with its neighboring spheres.
Within the scope of such an approximation, the pair interaction frequency in a
suspensions may be replaced by the collision frequency for the corresponding
pseudo-gase composed of hard spheres. Then the same factor Z(~) has to be
used to account for the increase in pair interaction frequency in concentrated
suspensions due to the particles being crowded in a confined space. Thus, the
last quantity in equation (48) has to be multiplied by the Enskog factor.
Obviously, components of the particle velocity variance tensor (47) represent,
by their very definition, averaged products of displacement velocity components.
The last velocity components equal the displacements per unit time, equation
(48), multiplied by the Enskog factor. Hence, we arrive at a general estimate

(Wx2) oc (w~,2 ) oC(Wz2 ) oc~ 2 Z2(~) 2' 2a 2 (49)

If the suspension is macroscopically inhomogeneous and (or) its shear flow is


not uniform, but length scale L characteristic of either the inhomogeneity or the
non-uniformity considerably exceeds particle radius a, the above estimate must
be still valid. Corresponding corrections can be shown to be of the order of a/L
by expanding n and ?, in equation (48) into their Taylor series.
Further, when having applied equation (47) to the plane shear flow illustrated
by Figure 6, we obtain

, 1 2
(50)
1271

whence, taking in account also equation (48),

A+B-(zc/
2 B 4) 1/2
~2
= 4 (51)

Using equations (49) and (51) enables us to reduce equation (47) to the
following form:

(w,,w,)=C~2x2(~)a2[(rc2/4-1)ew .ew +I2(ew)] (52)

where C stands for an unknown coefficient having the order of unity. This
coefficient may slightly depend on the suspension concentration, due to the semi-
empirical nature of the model used to evaluate the shear-induced particle
fluctuation velocity, as well as owing to the fact that multiparticle interactions
have not been considered in an explicit form. A more sophisticated theory is
clearly needed to reveal such a dependence, in which multiparticle interactions
would not be replaced by a hypothetical sequence of pair interactions as far as
their impact on the said fluctuation velocity is concerned.
It is worthwhile to note that, by using simple rational mechanical
considerations, we have arrived at a rather informative inference pertaining to the
longitudinal fluctuations in the flow direction (that is, along the x-axis), in spite
of the fact that these fluctuations did not enter our reasoning in a direct explicit
way. Namely, as follows from equations (50) and (52), the velocity variance for
these fluctuations must be precisely equal to that for fluctuations in the direction
of shear (that is, along the y-axis).
As a reasonable first approximation, fluctuations of different particles can be
regarded as statistically independent. Then, on accord with the arguments
forwarded in the end of the preceding section, the normal stresses generated by
shear-induced fluctuations can be evaluated from equation (46).

5.2 Particle Distribution in Rotational Couette Flow


To provide an application example, and also with a view in mind to compare
theoretical predictions with available experimental data, we now consider steady
flow of a neutrally buoyant suspension between concentric rotating cylinders.
The interphase interaction force altogether disappears in the flow trader
consideration. On the other hand, equation (46) completely determines the
normal stresses produced by particle fluctuations, if the particle velocity variance
tensor is known. In the generalized case where both shear-induced and thermal
fluctuations are relevant, this tensor can be presented as a sum of corresponding
1272

variance tensors for the shear-induced fluctuations and for the thermal
fluctuations when taken alone. Such a superposition approximation corresponds
to an assumption that fluctuations of the types mentioned are mutually
independent, which appears to be quite credible in view of fluctuations of those
types being originated by different physical causes. The radial components of
momentum conservation equation (2) and (3), to which stresses due to thermal
and shear-reduced fluctuations are added, then reduce to

r 2 dr r2r = 0 , r=pfM(~)y, Y=drr


(53)
d---~ r162 +CO (~)(ya) 2 =0

where r is understood as the averaged shear stress in the flow under


consideration, w is the mean suspension velocity which coincides with the mean
velocities of both suspension phases, and C is a new constant that differs from
C in equation (52). Solving the first equation (53) yields

F
Y - M(r
~ ' R1 < r < R2 (54)

where R I and /?2 are radii of the internal and the external cylinders,
respectively, and F is an integration constant that can obviously be related to
the torque applied to the rotating cylinder.
Next, we have from the second equation (53) and equation (54)

~b[1 + 4~bZ(~b)] ~pp+ C M(~b) (Fa)2 ( - ~ ) = const (55)

the constant to be determined from, say, a condition stipulating that the


suspension concentration averaged over the gap between the cylinders is given.
It is not difficult to see from equation (55) that Brownian motion makes for
the flattening of particle distribution in the gap, a distribution which otherwise
may be rather steep. If we confine ourselves to non-Brownian suspensions, the
term with kT/mp must be dropped out, so that constant C in fact disappears
from equation (55) as well. As a result, we arrive at the following equation that
determines the particle distribution within the gap:
1273

R 2 - - t F{~3 [I + 4~Z(~)]} 1/4 M(i~


Z(#~)) 2 (56)

Here q~ is a dimensionless constant that is a single-valued fimction of the


total contents of particles within the gap. Because constant C has vanished from
the analysis, equation (56) does not involve adjustable parameters at all,
provided ftmctions M ( # and Z(~ are given.
In our calculation, fimction M ( ~ as defined in equation (19) and the Enkog
factor as determined by equations (44) and (45) have been employed. Solutions
to equation (56) at different values of constant ~F are plotted in Figure 7.
Experimental dots obtained in [38] are shown in Figure 7 as well.

1 ' 1///oI/~'I'/__ ''/' 1 , , ~_


a
0.8 .' O "

0.6 0.6 9 aa
9 0
9 0

0.4 04 f ~
0.2 I 0.2I 4/ I t I
0.2 0.4 q} 0.6 0.2 0.4 ~0 0.6

Figure 7. Concentration profiles in rotational Couette flow, equation (56), at ~ =


0.65 and with Z(~ defined from equations (44) and (45) (a and b,
respectively), and experimental data by Phillips et al. [36]; a, curves 1 - 7
correspond to q~ = 1, 0.875, 0.75, 0.625, 0.5, 0.375 and 0.25, respectively; b,
curves 1 - 4 correspond to qJ = 0.7, 0.6, 0.5 and 0.4, respectively, and a thinner
line presents the same curve as curve 5 shown in a.

The agreement between theoretical curves and experimental dots is not bad at
all. Nevertheless, the dots apparently corresponds to a dependence of # on
relative coordinate r/R2 which is more gently sloping than the dependence given
by the theoretical curves. The dots are all the more so gently sloping in a range
of high concentrations. These curves conform to the experimental data somewhat
better, if allowance is made for the singularity of osmotic pressure at # - ~ , in
conformity with equation (45). Comparison of the curves plotted in Figure 7, a
with those presented in Figure 7, b proves, however, this singularity to be
1274
important only at concentrations which are sufficiently close to that of close
packing.
It is worth emphasizing once again that no free adjustable parameters have
been used when drawing any curve in Figure 7, but for close-packed state
concentration ~,. The agreement of theoretical predictions with experimental
evidence as witnessed by Figure 7 can therefore be regarded as a weighty
agreement in favor of the model developed. This notwithstanding, this agreemem
is not altogether as conclusive as it might initially seem because of two
approximations used. First, equation (19) has been assumed adequate to
represent the relative suspension viscosity, and secondly, equations (44) or (45)
have been chosen to represent the Enskog factor. One might argue that viscosity
must increase infinitely as the particle volume fraction tends to its close-packed
state value, so that model predictions in Figure 7 may be somewhat fallacious.
At any rate, Brady [26] has obtained a theoretical limit for the relative
suspension viscosity of Brownian according to which M(r --} (1 - #~)-2 at
---} ~ suspensions that proves to excellently agree with experimental data by
many researchers. For instance, the well-known Krieger's formula [41] gives

1-
~m/_1.82 (57)

and the exponent is this formula differs quite negligibly from the exponent of 2
appearing in the Brady's theoretical result.
Moreover, the osmotic pressure function and the Enskog factor must also
diverge in the indicated limit, so that singularities of M ( ~ and Z(~ are likely
to cancel each other out. Moreover, comparison of the curves plotted in parts a
and b of Figure 7 evidences that the effect of divergencies at ~ ~ ~ is not as
high as it might have been anticipated.

5.3 Diffusional Representation


The second equation (53) that stems from the radial component of the vector
momentmn conservation equation for the dispersed phase can be reduced to a
diffusion-like scalar equation. In is not difficult to show that, to arrive at such a
diffusion-like equation, we have to multiply the second equation (53) by the
effective particle mobility for Stokesian sphere in a concentrated suspension, and
then to multiply it also by particle mass rap. The particle mobility is expressible
in terms the hindered settling function, K(~, that represents the ratio of particle
sedimentation velocity in a concentrated suspension to the terminal velocity of a
single particle in an unbounded fluid [27, 40]. Namely, it equals the mobility of a
1275

single particle, (6z,ufa) ~ multiplied by K(O). The hindered settling function is


related to the relative suspension viscosity by an equation [40, 41 ]

K(r - (1 - ~)2 (58)


M(#)

The resultant diffusion-like equation that replaces the second equation (53)
looks as follows (it differs by a factor of (1- ~2 from the equation cited in [38]):

(1-#)2 d f
M(qk) dr 0[1+4r
I kT
6zc-pfa
+ C~b22,2 (~b) (7' a) 2 mp ]1 =0 (59)
6n'pfa 1!
This equation can be rewritten in a more familiar form, that is,

(DR + D,h)dq~
dr =0 (60)

where effective particle diffusivities caused by the thermal and the shear-induced
fluctuations are introduced by means of the following relationships:

DB = #)_____~2
(1- d {q~[1+ 4#Z(q~)]} D~ (61)
M(~) dq~

Ds~ _ ( 1 - ~ ) 2 d
- M(~--------~d~ {#3 Z2 (~)[1 + 4r162 } Ds~ (62)

where the following dimensional quantities are introduced

kT (ya)2mp 2C 7'2a 4
D~ - 6 rc,uf a ' Ds~ = C 6 rc/.tf a = --9 K (63)
vf

Equation (60) involves two terms that must be interpreted as net diffusion
fluxes produced by the thermal and the shear-induced fluctuations. In equation
(63), DR~ is the familiar coefficient of Brownian diffusion for a single fine sphere
immersed into an unbounded fluid, whereas D,h ~ plays the role of a characteristic
scale for the coefficient of particle mutual diffusion caused by shear-induced
1276

fluctuations. Concentration dependence for the coefficients of mutual Brownian


diffusion and of mutual shear-reduced diffusion is described by equations (61)
and (62), respectively.
The dependence of coefficient of mutual Brownian diffusion on concentration
is illustrated in Figure 8, the curves plotted in this figure corresponding to M ( ~
taken from either equation (19) or equation (57) and to different formulae for
Z(r In the dilute limit ~b ~ 0, this coefficient tends to DR~ and there is
altogether no difference between coefficients of Brownian mutual diffusion and
self-diffusion. However, such a difference arises and becomes more and more
pronounced as concentration increases. Suffice it to say that the long-time
coefficient of self-diffusion noticeably decreases with concentration [40, 42],
whereas the mutual diffusion coefficient is proven by the curves in Figure 8 to
substantially increase with concentration, in dilute and moderately concentrated
suspensions.
, , ,, , ,

2
o
2*

. . . . . _ . . . , ,

0.2 r 0.4 0.6

Figure 8 Relative coefficient of mutual Brownian diffusion as a function of


particle volume fraction; curves 1, 2 - M ( ~ is taken from equation (19), and
Z(~) is defined in accordance with equations (43) and (44), respectively; curves
1", 2" - M ( ~ is determined by the Krieger's formula (57), and 2'(~ is again taken
from either equation (43) or equation (44); curve 3 presents the Einstein's
formula (64) for dilute suspensions.

The Batchelor's formula [27] for the Brownian mutual diffusivity

DB/D~ = 1 + 1.45~ (64)

is illustrated by curve 3 in Figure 8. This formula is well corroborated by


experimental evidence. Because curve 1 goes sufficiently close to curve 3, it may
be said that curve 1 is also supported by observations.
1277

Curves 1 and 1" conform to using formula (38) when expressing the nonnal
stress in Brownian suspension flow. They essentially deviate from curves 2 and
2" that correspond to the usage in this context of the osmotic pressure function,
equation (29). Roughly speaking, curves 2 and 2* stem from the model
developed in [25], according to which the assemblage of suspended particles is
regarded as a pseudo-gas. In contrast to this, curves 1 and 1" arise out of our
modeling this assemblage like a system of molecules of a solute. The mere fact
that it is curve 1, but by no means curve 2, that satisfactorily agrees with
experimental data obtained for dilute suspensions serves as an implicit argument
in favor of our model.
As the suspension concentration increases beyond the range of moderately
concentrated suspensions, the mutual diffusion coefficient begins to decrease,
and it tends to zero as ~b~ ~ . This is undoubtedly due to an infinite increase in
viscosity in compliance with formula (57), and to the corresponding decrease of
particle mobility as suspension concentration grows.
Since a relative role played by direct particle contacts in momentum transfer
must grow with suspension concentration, one can justifiably expect that the
actual concentration dependence of the mutual Brownian diffusivity will
gradually undergo a smooth transition from curve 1 to curve 2* as concentration
increases.
It is worthwhile to mention in conclusion that a similar analysis can be
applied to calculation of the Brownian mutual diffusivity in colloids whose
particles exhibit molecular interactions of different origin [43], and also
electrostatic interaction [44]. The theoretical results derived in the papers cited
seem to be in a satisfactory agreement with experimental data. In principle, this
encouraging fact makes it possible to evaluate the particle diffusion flux, and
thereby to approach a difficult task of developing a rheological model for
colloids of interacting particles.
Figure 9 presents similar results for the dimensionless coefficient of mutual
diffusion of particles caused by shear-reduced fluctuations. This coefficient
appears to be rather negligible in dilute suspensions. However, it rapidly
increases with concentration, so that the particle migration generated by these
fluctuations becomes essential for highly concentrated suspensions.
In the case of shear-reduced diffusion, the distinction between the mutual
diffusivity and the self-diffusivity is even more drastic than that for Brownian
diffusion, since it touches upon the basic scaling of these coefficients. Indeed, in
conformity with equation (63), the coefficient of mutual shear-induced diffusion
scales with x 7' 2~vf , x being the density ratio. In contrast to this, the self-
diffusion coefficient is well known to scale with ?'a2 [33, 34].
1278

It has been noted already that there exist a number of attempts to model
particle distribution in shear-driven flows with the help of an equivalent diffusion
equation. Most commonly such attempts are based on an unwarrantable
assumption that the diffusion flux of suspended particles down a concentration
gradient is governed by the self-diffusion. Such an assumption is evidently
wrong, and without doubt, it is the mutual diffusion coefficient that must be used
in the indicated context. Moreover, in view of a so substantial difference
between these coefficients, conclusion drawn from models employing the self-
diffusion coefficient can be anticipated to be inadequate, both quantitatively and
qualitatively. A more detailed discussion of some shortcomings specific to such
models can be found in [38].
102 . . . . . '. . . . ,

2 ~t

ol

10-2

10-4 | I I i ,

0 012 0.4 0)6

Figure 9. Concentration dependence of mutual coefficient of shear-induced


particle diffusion; notation is the same as in Figure 8.

On the whole, the method of attacking shear-induced particle migration as


developed in this section provides a succinct springboard needed to study both
particle distribution in shear-driven suspension flows of different types and
influence of this distribution on the flow hydraulic characteristics. This method is
based on the necessary and quite obvious requirement that the dispersed phase
normal stress must satisfy the components of the dispersed momentum
conservation equation in directions perpendicular to the mean motion, without
invoking an independently stated diffusion equation. In simpler cases, like the
one related to the rotational Couette flow of a neutrally buoyant suspension that
has been considered above, this requirement reduces to the condition of the
particulate stress being constant in the said directions. As such, our model has
phenomenological similarities with the model proposed in [37], although the
normal stress calculation is different.
1279

6. MODELING PSEUDOTURBULENT FLUCTUATIONS

Apart from shear-induced fluctuations, suspended particles are involved in a


hydrodynamically induced chaotic fluctuating motion of a quite different physical
origin. This so-called "pseudoturbulent" motion is specific to fluidized beds and
to vertical suspension flows at large, as well as to pressure-driven suspension
flows in which the mean velocities of the suspension phases happen to be
different, so that a relative fluid slip flow occurs. The underlying physical
mechanism of the pseudoturbulent fluctuations has been extensively discussed in
a number of papers, and in particular, in [45]. In this mechanism, mean relative
fluid flow interacts with random fluctuations of suspension concentration, that is,
with fluctuations of the total number of particles within any given mixture
volume. Such concentrational fluctuations are due to various chance causes, and
they are inevitably present in any suspension flow.
Since the viscous drag experienced by particles is a strongly nonlinear
function of local concentration, concentrational fluctuations are bound to cause
the suspending fluid to exert a fluctuating force on any given particle. This
fluctuating force accelerates the particle either in the direction of relative slip
flow or in the opposite direction. An additional fluctuating force that acts on the
fluid in an external body force field also appears due to density fluctuations
which are closely connected with the concentrational fluctuations. These forces
make the particles and the fluid fluctuate in the direction of slip flow, and also in
the direction of external body force. In vertical suspension flow, such as flows
accompanying fluidization and suspension sedimentation, the two indicated
directions coincide, and only vertical fluctuations initially originate. Due to
fluctuations of fluid pressure and relative velocity, as well as to both
hydrodynamic and collisional interparticle interactions, the kinetic energy of
original fluctuations is redistributed in such a way as to excite fluctuations in the
other directions. As a result, an anisotropic fluctuating motion establishes itself in
both suspension phases.
If suspended particles are large enough for the interparticle exchange by
momentum and energy to be mainly performed by means of direct collisions of
the particles, the pseudoturbulent fluctuating motion is approximately isotropic.
In this case, the particles fluctuate almost as statistically independent entities,
like molecules of a gas undergoing thermal motion. Properties of the
pseudoturbulent motion can then be investigated by analogy with the kinetic
theory of dense gases, as was suggested, seemingly for the first time, by Jackson
[46]. Although study of pseudoturbulent fluctuations in coarse dispersions with
interparticle collisions certainly goes beyond the intended scope of this Chapter,
1280

we point out that various phenomenological approaches exist which are based on
such an analogy and which are primarily aimed at incorporating the effects of the
pseudoturbulent fluctuations on mean suspension and fluidized bed flow. Such
approaches are exemplified in a number of works, and in particular, in [47]. A
more sophisticated analytical kinetic theory approach is discussed in [48].
Another limiting case pertains to quiescent suspensions of fine particles which
exchange momentum and energy exclusively through random fields of ambient
fluid velocity and pressure, and collisions do not play a noticeable role in this
exchange. It is only this case that we are going to address below. A significant
feature of the pseudoturbulent motion generated in finely dispersed suspensions
consists in the fact that suspended particles are by no means statistically
independent. Long-ranged hydrodynamic interactions of the particles result in the
formation of group containing many particles that fluctuate in a correlated
manner. Such groups keep to incessantly originating and then disintegrating, so
that the fluctuating motion can be phenomenologically viewed as an ever altering
collection of intermittent flow patterns which are reminiscent of fluid moles in
one-phase turbulent flow.
In order to find out relevant statistic properties of the particulate groups, it is
necessary to develop an efficient model of multiparticle hydrodynamic
interactions. A generally accepted analytical approach to modeling flow of
quiescent suspensions consists in summing up all such interactions involving
different numbers of particles. In practice, such a summation has been effected
only for pairwise interactions, with the help of a nice renormalization procedure
as first invented by Batchelor to circumvent singularities and to get convergent
results for macroscopically homogeneous dilute suspensions (see, for example,
[8]). Unfortunately, some theoretical conclusions that have been obtained to
date while treating fluctuations by means of different versions of this
renormalization procedure seem to be unacceptable from the physical point of
view. Thus, analytical studies and numerical simulations of a sedimenting
suspension have shown that the variances for particle fluctuation velocity and
coefficients of particle pseudoturbulent self-diffusion in different directions
increase unboundedly with the size of the vessel that contains the suspension, if
the spatial distribution of particles is supposed to be random [49-51]. The last
inference is not supported by experimental data which prove the quantities
mentioned to be insensitive to the container size if it exceeds a certain critical
size [52,53]. These experiments found no systematic variation in the particle
velocity fluctuations with container size at all.
A possible explanation of this controversy might be founded on a stmmse that
a short-range order establishes itself in a sedimenting suspension, so that a non-
random suspension microstructure sets in. In particular, Koch and Shaqfeh [54]
1281

have suggested that changes in pair correlations as induced by the sedimentation


process might lead to a screening of the hydrodynamic interactions. In such a
case, velocity fluctuations could be finite and independent of container size for
sufficiently large vessels. Indeed, these authors have concluded that there is a net
deficit of particles in the neighborhood of any particle, and that this deficit
provides for such a Debye-like screening. However, recent numerical simulations
performed by Ladd [55] have demonstrated that there is no such particle deficit
at all, and correspondingly, that the hydrodynamic interactions of particle pairs
are not screened by the changes in the pair distribution function at long distances.
Moreover, Ladd has drawn a conclusion that the particle velocity fluctuations
essentially depend not only on the size, but also on the shape of the container.
An alternative mechanism, that admittedly results in making the particle
velocity variances and self-diffusivities insensitive to container size and shape,
has been proposed in [56]. According to this mechanism, the pseudoturbulent
fluctuating motion of particulate groups brings about additional viscous
dissipation of kinetic energy. This dissipation gives rise to a corresponding
effective average force that efficiently retards fluctuations of any particle, and
thereby make the fluctuation intensity finite, irrespective of how large is the
container size. The concept of effective force that slows down the fluctuations is
certainly semi-empirical. However, it can be readily understood on the basis of
the familiar ideas specific to various versions of the well-known mean-field
approximation. Below, we shall briefly describe the model as developed in [56]
for vertical suspension flow.

6.1 Friction Force Due to Energy Dissipation


Let us evaluate the viscous energy dissipation that is supposedly caused by
the pseudoturbulent fluctuations of particulate groups. These groups are
separated by changeable boundary interlayers, and they incessantly exchange
particles between themselves. Within the scope of a semi-empirical model that
we are about to disclose, however, such groups may be regarded without loss of
generality as statistically independent, whereas the behavior of particles that are
contained in any group at any given moment of time may be looked upon as fully
correlated. This means, in particular, that the fluctuation velocity of a group is
the same as the velocity of any of its particles. If we make use of a characteristic
particle correlation length as evaluated in [54], we shall arrive at the following
estimates for group's linear dimension, L g , the group's volume, Vg, and the
average number of particles within one group, Ng:

a a 3 1
Lg = c L -~ , Vg = cV ~-~ , Ng = c N 02 (65)
1282

where numerical coefficients of the order of unity are introduced.


Energy is supplied to the pseudoturbulence from the relative fluid flow as it
operates at eoneentrational fluctuations, in compliance with the mechanism
described in [45]. In the steady state, this energy input is exactly compensated by
the viscous energy dissipation, so that the pseudoturbulence intensity is
maintained at a certain invariable level. There are two contributions to this
dissipation: the one that is associated with the fluctuating drag forces acting on
individual particles, and the one due to the random field of variable shear rates as
generated by the fluctuating motion of the groups. The immediate objective of
this subsection is to approximately evaluate the second contribution, for which
purpose we shall use a simple consideration based on the theory of similarity.
It is evident that the fluctuating shear rates are of the order of W/Lg, W
standing for a characteristic value of the RMS particle fluctuation velocity. When
again making use of the effective medium model as formulated in Section 3, we
are able to evaluate the energy dissipated by the random shear motion of the
particulate groups in a unit volume per unit time. By the order of magnitude, this
dissipated energy equals ktfM (W/Lg)2. The total excessive viscous dissipation of
energy, AE, attributed to one group is equal to this last quantity multiplied by
group volume, meaning that its is of the order of/afM W2Lg ~ ~ M W 2 a/qk.
Now, we want to describe this viscous dissipation by means of introducing an
effective retarding friction force, F~ = - a mpw', that acts on one particle on the
average. It is fairly clear that AE must be equal to the work accomplished by
such friction forces experienced by all Ng particles within a group per unit time,
when taken with an opposite sign. Then, using equation (65) yields an order-of-
magnitude equation to solve for unknown coefficient a

AE oc /uf M W 2 -a ocOtmpNg(w'2)ocotppW 2 a3 (66)

whence we obtain, as a final result,

Fr = - o t m p w ' , a = c a ppa2
luf CM(qk) (67)

where c~ is a new coefficient of the order of unity.


Such an introduction of retardation friction forces exerted on the particles is
in accord with standard methods of the mean-field theory. These forces are of
key significance since they can be proven (and will actually be shown below) to
prevent the particle velocity variances and self-diffusivities being divergent in the
1283

dilute limit. It is evident that, as far as the divergence of the said quantities is
concerned, the dilute limit plays the same role at a given container size as an
infinite increase of the container size does at suspension concentration kept
constant.

6.2 Stochastic Equations


The Langevin equation for a group containing Ng particles can be formulated
as

mp Ng dw' = Ng (F' + F r) (68)


dt

where F~ is defined by equation (67) and F' is the fluctuating force


experienced by one particle within the group. The differentiation is performed in
equation (68) along the group trajectory. Summing up the Langevin equations
(68) for all groups within a unit volume of the suspension and using equation
(67) yield the following new equation:

CPP 0---~= nF' - aqkppw' (69)

Quantity w' represents a sum of random velocities of particulate groups in a


unit volume which are assumed to be not correlated. It is quite evident that the
group number concentration can be made as small as is desired, by appropriately
defining the coordinate length scale. Then, if also a coordinate system moving
with the dispersed phase mean local velocity is used, the partial time derivative
can be substituted for the total derivative that appears in equation (68). This has
been actually done when formulating equation (69).
Two supplementary stochastic equations can be derived from the continuum
mass and momenttma conservation equations for the continuous phase as
formulated in equations (1) and (2). These stochastic equations read [56]

+u. v)r = (1- r v. v'


(70)
(1 - qk)pf(~t +u. V) v' =-Vp'+jufMAv'-nF' -Fn'-pfg~b'

where the convective coordinate system connected with the dispersed phase
motion is used again. To simplify the matter, we have ignored time and space
1284

dependence of the mean flow variables when formulating these equations. Thus,
strictly speaking, equation (70) has formally a bearing upon fluctuations
occurring in a macroscopically uniform steady suspension flow.
The fluctuations of mean flow variables that appear in equation (70) are also
sums of corresponding independent random contributions associated with
different particulate groups. In accordance with the central limit theorem of the
theory of probability, their variances equal the corresponding variances for one
group divided by the group number concentration that may be presumed large as
compared to unity. Thus, equations (69) and (70) can be thought of as the ones
related to the fluctuations of a single group.
Since stochastic equations (69) and (70) are linear in fluctuations, it is natural
to employ the correlation theory of stationary random processes, an informative
review of which is given in [57]. According to this theory, all fluctuations are
represented as Fourier-Stieltjes integrals over frequency co and throughout the
entire wave-number space k. Using these representation reduces differential
equations (69) and (70) to a set of linear algebraic equations in which the role of
tmknown quantities is played by random measures that are involved in the
Fourier-Stieltjes integrals. These last equations enable us to express all the
random measures as quantities proportional to the random measure for random
concentrational fluctuations, d Z~. Correspondingly, all spectral densities of
interest turn out to be proportional to the spectral density of the concentrational
fluctuations, q-'r (co,k), with proportionality coefficients being functions of co
and k. Using tile correlation theory of stationary random processes allows
various two-time two-point correlation functions to be found by standard means
characteristic of this theory [57].
An expression for the fluctuation force is to be determined from equations (8)
- (13). For the sake of simplicity, we take into account only the contributions to
the interphase interaction force that are due to viscous drag and to gravity
corrected for buoyancy. After a simple manipulation, we then arrive at the
following formula:

1 dM _ ~x - 1 g~, ] r = - ~2 a 2
nF'=qkPp[~r(v'-w')+~~ (71)
tcr dqk u~' x ' 9 vf

It is also not difficult to get from equation (3)

Fn' : -ppg~k' (72)

Using equations (71) and (72) in the algebraic equations for random measures
that obtain when the aforementioned Fourier-Stieltjes integrals are inserted into
1285

equations (69) and (70), and then solving these algebraic equation, result in final
formulae for the random measures of the pseudoturbulent fluctuations. In
particular, it is not difficult to derive general expressions for the random
measures of random fluid and particle velocity fluctuations v' and w'

o+u.kk
dZv = ck k S- dZ~
H
(73)
_ E { c, o + u . k _k r [s_(S'k)k] }
dZw ico+a+E ,~k k

where scalar and vector functions involved are introduced in accordance with the
following equations:

(io+a)E
e - terM , H = i (1- qkl Pf ( ~ + u "k ) + ttf M k Z + ~kPP i o + a + E
(74)
S= ico+a B+C, B=~u+c.dlnM C=(tc-1)r
ico + a + E dO ~bM

Equation (73) gives an opporttmity to derive equations for the spectral density
tensors of fluid and particle velocity fluctuations which express them as
quantities proportional to the scalar spectral density of concentrational
fluctuations. In order to be able to calculate the variances of particle velocity
components by integrating corresponding components of the spectral density
tensor over co and k, a trustworthy representation for the spectral density of
concentrational fluctuations is necessary. The wanted representation was
discussed at length in [56] where both particle velocity variances and self-
diffusivities were studied.
In what follows, we are interested in the determination of particle velocity
variances, but not of coefficients of particle pseudoturbulent self-diffusion. In
this ease, a detailed discussion of the eoneentrational fluctuation spectral density
proves to be excessive. As will become clear later on, only the variance of
coneentrational fluctuations is actually needed to get reasonable estimates for the
particle velocity variance under certain simplified conditions. The wanted
expression for this variance follows from [58] where the classical
Smoluehowski's combinatorial theory of fluctuations in dilute particulate systems
has been generalized to concentrated systems. This expression reads
1286

(75)

equation (65) having been taken into account when obtaining the second
equality.
At last, let us indicate sufficient conditions which must be fulfilled to allow us
to make use of a simple equation (75), instead of a much more complicated
representation for the concentrational fluctuation spectral density. As has been
discussed in [56], for this purpose in view we have to 1) replace fluid and
particle velocity spectral densities by their values at small frequencies, 2) assume
the integral of the concentrational fluctuation spectral density over frequency to
be dependent on modulus k of wave-number vector k, but not on this vector
direction (that is, not on angular coordinates of a spherical coordinate system in
the wave-number space), and 3) ignore viscosity effects.
The terms that involve frequency may be dropped out of equations (73) and
(74) if pseudoturbulent motion is quasi-stationary in a certain sense. This
condition of psedoturbulent fluctuations being quasi-stationary amounts in fact to
an assumption requiring a characteristic fluctuation frequency scale to be much
smaller than friction force coefficient a determined in equation (67). The
characteristic frequency scale for fluctuations must be of the order of vrMLg-2 ~
vf M~ 2/a2, whereas coefficient a ~ vf M ~ 2/1r 2. Thus, the condition trader
discussion must hold good for suspensions of sufficiently low concentrations, all
the more so, the larger the density ratio. In the generalized case, however, both
frequency scale and friction force coefficient are of the same order of magnitude,
and the said condition presents an approximation the adequacy of which has to
be specially checked.
The second condition requiting the integral of the concentrational fluctuation
spectral density over frequency to be independent of the wave-number vector
direction states that the concentrational fluctuations are nearly isotropic. This
represents a reasonable approximation that has its direct analogies in the theory
of fluctuations in molecular systems. At last, ignoring viscosity effects amounts
to the neglect in the expression for H in equation (74) of the term containing
fluid viscosity gf. This term can readily be shown to effect only quite a
negligible influence on the particle velocity variances.
For simplicity and definiteness, we shall consider the velocity variances only
in a vertical suspension flow (or in a fluidized bed). Vectors S, B and C
defined by equation (74) are collinear in such a flow, and it is easy to understand
that 1) only diagonal components of the fluid and particle velocity variance
tensors differ from zero, and 2) these tensors are axially symmetric with respect
1287
to the vertical axis. Thus, only diagonal components of the corresponding
spectral density tensors are actually necessary to evaluate the vertical and
horizontal velocity variances. Expressions for these diagonal components follow
from equation (73). Under the conditions enumerated above, these expressions
look as follows:

tlJvl'vl((~
{ (-~14(1-~u ) 2k?
2) $2
(76)
q2 qbC + sB - ~bC + s B + i -u~b qJ~,~(co.k)

Wv2,v2(co, k) ~ (1 + fl)2 Ww2,w2(Co,k )

Wwl,wl(CO,k) ~ f12r 1 +q2) B + l-~b

q20C- qbC+s B + i , 0 q~,r

~w2,w2(co,k)~ f1202(k~+q2
1 )
{(/2
s2 1_uo kl2
(77)

q2 r B+ k2 )

where

a
fl=~=cflqk, 2
cfl=-~Ca, s = ~ fl~
, q = - qbpp
- fl E (78)
l+fl o~pf l+fl u

Equations (76) and (77) are formulated in a coordinate system which first
coordinate axis is directed along u, its two other orthogonal axes being
arbitrarily chosen in a plane normal to u. Expressions for the spectral density
tensor components that correspond to the third coordinate axis are to be obtained
1288

from those corresponding to the second coordinate axis, if the third wave-number
vector component, k3, is substituted for the second one, k2.

6.3 Experimental Verification


If the vertical suspension flow under study is uniform, then

u = u, - -(K- 1) (1 - q6)z" 1 - # Vt (79)


m(r

where Vt is the terminal velocity of a single particle falling in an unbounded


fluid. Using this expression in equation (74) yields

,- dlnM 0 u, , c= u, (80)
1-# #(1-#)

Equation (80) enables us to substantially simplify expressions (76) and (77).


To make further calculation of the fluid and particle velocity variances even
simpler, we assume that a characteristic scale of k is much smaller than q. If
we evaluate this scale as ck#a, in accordance with equation (65), then this
condition of q being relatively large can be roughly formulated as follows [56]"

tr - 1 (1 - r ga 3 (81)
~Fr<<l, Fr- 2
lo r vf

When formulating this inequality, the formula for relative suspension viscosity
from equation (4) and equations (78), (79) have been used. Condition (81) is
always satisfied if suspended particles are fine enough, so that their Froude
number, Fr, is small. However, inequality (81) becomes invalid, and it ultimately
transforms into an inverse strong inequality, in the dilute limit, as # approaches
zero.
If inequality (81) holds true, the terms proportional to k~2 in the denominator
and numerator of equations (76) and (77) can be ignored compared to the terms
that are proportional to q2. This simplifies the spectral density expressions
identified in equations (76) and (77) to quite a considerable extent, so that these
expressions can easily be integrated over the wave-number space. In particular,
by making use also of equation (80), we obtain from equation (77) the following
simple approximate formulae for the diagonal components of the particle velocity
variance tensor in a uniform vertical flow [56]:
1289

'
( / u_, ) 2
qJ~,~(co,k)
Wwl,wl(CO,k) ~ ]320 2 1- y ~-~-j 1 r
(82)
ca~ 2
%2,wz(co,k) ~r162 ~2 ) i - r %,q~(co,k), y = ~
l+c,ar

The wanted particle velocity component variances are to be calculated by


integrating over co and k the corresponding expressions for the spectral density
tensor components as defined by equation (82). As has been already pointed out,
the integral of the concentrational fluctuation spectral density over co is assumed
to depend only on modulus k of wave-number vector k. The coefficients at this
spectral density in expressions (82) involve only angular coordinates in the
wave-number space, but they are dependent on neither k, nor co. Furthermore, the
integration of the concentrational fluctuation spectral density over k and co
gives variance (75) of these fluctuations. Keeping in mind these simple
considerations, and also making allowance for equations (4) and (79), we obtain

1/2
vt ~Cw 1 - 3 + 5 (1-~) s/2 1- ~
(83)
1
vt ~ Cw ~ 1- cw = cp r~-~

According to this equation, the ratio between lateral and longitudinal RMS
particle fluctuation velocities is equal to

iw+w211'2 iw ll,2 (
(wf21"2 _ (wf2) 1/2 -_ 15-107'+3y
2y2 2
11,2 (84)

In conformity with the formulae in equation (83), both longitudinal and lateral
RMS velocity components tend to finite limits as ~b comes to zero. Thus, the
pseudoturbulent particle velocity variances do not diverge in the dilute limit.
However, the adequacy of these formulae is conditioned by inequality (81) being
valid, so that they cease to be true in the dilute limit. It can be shown that these
formulae give way in the dilute limit to other relations for the said RMS velocity
components [56]. The last relations prove both lateral and longitudinal RMS
1290

velocities to be proportional to r 2, their ratio 6~ tending to a constant value of


0.067 as r approaches zero. Similarly, the coefficient of particle pseudoturbulent
self-diffusion can be proven to vary proportionally to r at low ~.
Equations (83) and (84) provide an opportunity to check theoretical
conclusions by comparing them with available experimental evidence. Such a
comparison with data obtained for sedimenting suspensions in [53] is illustrated
by Figure 10 and Figure 11. The agreement between our model and experiments
as evidenced by these figures seems satisfactory, in spite of the fact that specific
values have been used for the adjustable factors that appear in theoretical
equations (83) and (84). The deviation of the theoretical curves from
experimental dots at low concentrations is quite understandable because these
equations fail to be valid in the dilute limit, and they have to be replaced by other
equations, as has been already pointed out.
As has been shown in [56], both longitudinal and lateral coefficients of
particle pseudoturbulent self-diffusion as evaluated on the basis of the model
developed also agree with the experimental data of [52,53], and this fact lends an
additional support to the model developed for pseudoturbulent fluctuations in
eollisionless suspensions.
0.8 -..\ . . . . . . . . . . .
0.7

0.6

0.5 9
Cq

0.4 I
~. o.3
0.2

0.1

0 . . . . . . . . . . . . . . . . . . . . . .
0 0.1 0.2 0.3 0.4 0.5

Figure 10. The longitudinal (vertical) particle velocity fluctuation scaled with
particle terminal velocity according to equation (80) at ~ = 0.6 and cp = 0.6;
the upper and lower curves correspond to c:v~/2 = 1.2 and 1.4, respectively; dots,
experiment in [53].

6.4 Pseudoturbulent Stresses


As follows from our model, the chaotic particle pseudoturbulent motion
cannot be viewed as statistically independent fluctuations of individual particles.
Rather it is similar to turbulent fluctuations of fluid moles in one-phase turbulent
flows. The same statement holds true for the pseudoturbulent motion of the
1291

ambient fluid as well. For this reason, equation (42), that has relevance to
fluctuations of individual statistically independent particles which do not collide
but interact only with the fluctuating ambient fluid, certainly cannot be applied to
the description of stresses that originate due to pseudoturbulent fluctuations.
Thus, an alternative way to describe these stresses has to be put forward.

!.8

I hl.6

1.4

1.2

0 0.i 0.2 0.3 0.4 0.5


0
Figure 11. The ratio between longitudinal and lateral standard particle velocity
deviations as a function of suspension concentration; curves correspond, in the
descending order, to equation (81) with c~ = 0.4, 0.6 and 0.8; dots,
experimental data from [53 ].

Such a method can be developed from considerations that are analogous to


those used by Roco [59] in his treatment of two-phase turbulent flows.
According to this method, conservation equations (1) - (3) are to be employed
again to describe the averaged fields of smoothed flow variables pertaining to the
interpenetrating continua that model the suspension phases. After that, these
fields are supposed to fluctuate in just the same way as the fluid velocity and
pressure fields fluctuate in an one-phase turbulent flow, and a new, secondary,
averaging procedure is applied to average these equation with respect to such
quasi-turbulent fluctuations. This second averaging procedure is quite similar to
that which is commonly used while dealing with one-phase turbulent flows and
that leads to the appearance of Reynolds stresses in the newly averaged
equations. If this procedure is adopted, the wanted expressions for
pseudoturbulent stresses should not differ from those for the Reynolds stresses in
turbulent flows.
As a result of applying this averaging procedure to equations (1) - (3), we
arrive at additional Reynolds-like stresses

a -p< - s, = (85)
1292
that are supposed to act in flow of both continuous and dispersed phases due to
their pseudoturbulent fluctuations. The averages that appear in equation (85) can
be calculated from the model of pseudoturbulence as developed earlier in this
section. In particular, while dealing with equations (76) and (77), we can use
precisely the same assumptions that have led to formulae (83). In this case, we
shall obtain, atter a manipulation, expressions for normal pseudoturbulent
stresses that act in a nonuniform vertical suspension flow. We have for the
continuos phase

(1+ Cfl~)2 i1 _ ~m_m/IV(0)+ V(1) z/ u


(86)
~ f = - 15 c fl2CN(1 - qk) u,

and for the dispersed phase

+ --+ pp (87)
15 cfl 2 c N (1 - ~b)2 U,

where j = 1, 2, 3, new coefficients are introduced that are functions of

W l ( 0 ) = 8 - 4 s + 3 s 2, Wl(1)=2s(-2+3s)(h+l), Wl(2)=3s2(h+l) 2
V/(0) = (1 + s) 2 , V(1) = 2s(1 + s)(h + 1), V(2) = s 2 (h + 1)2 (88)

Wi(k)=V ( k ) , i=2,3, k=0,1,2, h=(1-~b) aln------M-M


a0
and expressions for B and C resulting from equation (74) have been employed.
The pseudoturbulent normal stresses defined in equations (86) - (88)
represent fimetions of two flow variables: local volume concentration ~ and
local mean fluid slip velocity u which are in general mutually independent. In
uniform flows, these variables are related to each other by equation (79), so that
the normal stresses are completely determined by a single flow variable. As
follows from equations (86) - (88), normal stresses that act in a uniform flow in
either continuous or dispersed phase of a dilute suspension and that are attributed
1293

to the longitudinal (vertical) direction are eight times as large as corresponding


normal stresses attributed to any of the lateral (horizontal) directions.
It is worth noting in conclusion to this section that there exists an alternative
model of pseudoturbulent fluctuations [60]. In contrast to the model as presented
above, this alternative model implies that particles that undergo pseudoturbulent
fluctuations may be assumed to behave themselves as approximately statistically
independent entities. Such an assumption contradicts the widespread notion that
particles fluctuate as members of groups composed of closely correlated
particles. However, this model findings may also be supported by some
experimental evidence [60]. It seems just possible that the overall
pseudoturbulent fluctuating motion might involve two different constituents that
have different length scales. According to such a scheme, approximately
independent fluctuations of particles, presumably due to the nonlinearity of the
dependence of the interphase interaction force on local concentration, are
superimposed on fluctuations of groups containing many particles, presumably
owing their origin to the interaction of the gravity field with density fluctuations
that are known to be a primary cause of the divergences discussed in [49-51, 54,
55], and also earlier in this paper. Without doubt, this important issue deserves
the most close attention in future theoretical studies.

7. CONCLUSIONS

The two central notions advanced in the present Chapter are actually rather
trivial. The first notion represents a recognition of the simple fact that even
simplest suspension flows are subject to the stratification caused by forces
exerted on suspended particles in the direction normal to the dispersed phase
streamlines. If those forces are not opposed by dispersed phase stresses, this
stratification ultimately leads to the separation of the whole flow domain into
regions containing close-packed particles or altogether devoid of particles.
Consequently, steady suspension flows cannot be adequately and effectively
described with the help of existing conventional fluid dynamic models, if these
models do not allow for concentration-dependent stresses that are specific to the
dispersed phase and that condition this phase stratification.
The second notion intimates that the particulate stresses, which are capable of
hampering the flow stratification in unsteady flows and of bringing it to an end in
steady flows, are due to random fluctuations of the suspended particles. The
action of the particulate stresses results in the establishment of non-trivial
particle distributions that play a paramount role in evolving the fields of all the
flow mean variables and in making up the flow hydraulic characteristics.
1294

Apart from turbulent fluctuations of suspended particles which have not been
considered in this Chapter at all, fluctuations of three distinct types that owe their
origin to different physical mechanisms have been distinguished. These are the
thermal particle fluctuations, and also two types of hydrodynamically-induced
fluctuations: the shear-reduced and pseudoturbulent ones. The fluctuations of all
the types stimulate particle migration which in arm produces a net diffusion flux
down a particle concentration gradient. This flux partially compensate for the
convective flux of the particles caused by the forces directed normally to the
dispersed phase flow streamlines. As a result of the combined effect of these
fluxes, a particle distribution originates that, in steady flows, correspond to
dynamic equilibrium attained between convection and diffusion of particles.
When dealing with a suspension flow, simultaneous allowance for the
relevant fluctuations belonging to different types seems imperative. First, the
coupling of different fluctuations results in the occurrence of non-monotonous
dependencies of meaningful flow characteristics on parameters, such as the
particle size, which can significantly influence various technological processes.
This effect of the coupling is exemplified by a uniform shear layer flow that
has been considered at some length in [38]. A convenient measure of the ability
of uniform horizontal steady shear flow of a given shear rate to suspend particles
is the total mass (or volume) of suspended particles per unit area of the flow
bottom plate. This total mass has been proven to be a non-monotonous function
of particle size that has a minimum at a certain size value, all other physical
parameters being presumed fixed. For smaller particles, the flow suspending
ability increases owing to the corresponding enhancement of thermal
fluctuations. For larger particles, this ability also increases, but due to the
augmentation of shear-induced fluctuations. This conclusion has an immediate
bearing on manifold field-flow fractionation processes in practice.
A similar non-monotonous dependence on particle size can be proven to be
specific to fluidized beds and quiescent sedimenting suspensions of fine
particles. The total mass of particles that form an upper part of a fluidized bed
and that is situated above the approximately uniform bulk region of the bed has a
minimum for particles of a certain radius. This minimum has precisely the same
physical origin as that for the total mass of particles suspended by a uniform
shear flow, except for the fact that the role of shear-induced fluctuations is now
played by pseudoturbulent fluctuations.
Secondly, failure to simultaneously account for fluctuations of different origin
can lead to unwarrantable conclusions that may be erroneous not only
quantitatively, but also qualitatively. A convincing example is provided by a
pressure-driven one-dimensional steady suspension flow in a channel. Even if the
suspension is non-Brownian, neutrally buoyant and finely dispersed, allowance
1295

for the stresses due to both shear-induced and thermal fluctuations is vitally
important. If only shear-induced fluctuations were allowed for, we would come
with a conclusion that a close-packed core region always obtains in the flow, no
matter how low is the particle concentration averaged over the flow cross-section
[36,38]. Indeed, in such a case, the lateral particulate stress disappears at the
channel central axis (or plane) where the mean shear rate turns to zero. Hence it
follows that this lateral stress cannot be maintained uniform over the cross-
section, if the suspension concentration at the central axis (plane) is less than its
value associated with the close-packed state.
In the opposite case where only pseudoturbulem fluctuations were taken into
account, we would conclude, also erroneously, that the particles are uniformly
distributed in each cross-section. It is clearly the conditions of suspension
concentration and fluid slip velocity being invariable throughout any cross-
section that provides for the required tmifonmty of the lateral pseudoturbulent
stress. If the concentration does not vary in a cross-section, the fluid slip velocity
is also uniform. (Indeed, relative fluid flow is caused by the very same
longitudinal pressure gradient that produces the flow. This pressure gradient is
well-known to be uniform in a steady one-dimensional channel flow. Thus, given
that the pressure gradient is fixed, the distribution of fluid slip velocity over flow
cross-sections actually a function of suspension concentration alone.)
Simultaneous allowance for fluctuations of different types becomes even
more as far as general three-dimensional unsteady flows are concerned. In
particular, this is due to the fact that the corresponding particulate stresses
influence not only particle distributions as attained in various steady flows, but
also the rate with which these stationary distributions, and also the steady flows,
establish themselves. Besides, these stresses sometimes play a fundamental role
in making for hydrodynamic stability of steady flows [46, 48, 58].
It is expedient to emphasize in conclusion that much remains to be done in
order to develop a completely reliable, comprehensive and workable theory to
treat fine suspension flows. The point is that the tentative model as developed
and presented in this Chapter is not free of numerous semi-empirical assumptions
the validity of which must be examined in detail. To provide for such an
examination, tangible subsidiary models must be further elaborated and made
more precise for fluctuations of different origin, and also for stresses induced by
these fluctuations. Unavoidably, this task requires very serious efforts to be
applied in a specific research area bordering on mechanics of multiphase flow,
rheology and statistical physics. However, such efforts seem to be fairly justified
and worth being undertaken in view of important practical issues that are likely
to come about in connection with munerous branches of technology.
1296

REFERENCES

1. Y.A. Buyevich, Arch. Mech. 42 (1990) 429.


2. P.G. S ~ a n , J. Fluid Mech. 22 (1965) 385; and 31 (1967) 624.
3. M. Ishii, Thermo-Fluid Dynamic Theory Two-Phase Flow (Eyrolles, Paris,
1975).
4. D.A. Drew, Ann. Rev. Fluid Mech. 15 (1983) 261.
5. R.I. Nigmatulin, R. T. Lahey and D. A. Drew, Chem. Engng Comm. 141-
142 (1996) 287.
6. Y.A. Buyevich and I. N. Shchelchkova, Progr. Aerospace Sci. 18 (1978)
121.
7. Y. A. Buyevich and T. G. Theofanous, ASME Winter Meeting, Dallas,
1997.
8. G.K. Batchelor and J. T. Green, J. Fluid Mech. 56 (1972) 401.
9. G.K. Batchelor, J. Fluid Mech. 83 (1977), 97.
10. Y.A. Buyevich, Chem. Engng Sci., 50 (1995) 641.
11. L.D. Landau and E. M. Lifshitz, Fluid Dynamics (Pergamon Press, Oxford,
1959).
12. G. K. Batchelor, An Introduction to Fluid Dynamics (Cambridge Univ.
Press, Cambridge, 1967).
13. N. Zuber, Chem. Engng Sci. 19 (1964) 897.
14. L. van Wijngaarden, J. Fluid Mech. 77 (1976) 27.
15. J.B.W. Kok, Physica A 148 (1988) 240.
16. A. Biesheuvel and S. Spoelstra, Int. J. Multiphase Flow 15 (1989) 911.
17. B.U. Felderhof, J. Fluid Mech. 225 (1991) 177.
18. D.Z. Zhang and A. Prosperetti, J. Fluid Mech. 267 (1994) 185.
19. R.-H. Jean, and L.-S. Fan, Powder Techn. 72 (1992) 201.
20. G. Astarita, Chem. Engng Sci. 48 (1993) 3438.
21. Y.A. Buyevich, ASME AMD-Vol. 217 (1996) 161.
22. F. Feuillebois, J. Fluid Mech. 139 (1984) 145.
23. A.D. Maude, J. Fluid Mech. 7 (1960) 230.
24. B.S. Yendler, Inzh.- Fiz. Zh. 37 (1979) 110.
25. J.F. Brady, J. Chem. Phys. 98 (1993) 3335.
26. J.F. Brady, J. Chem. Phys. 99 (1993) 567.
27. G.K. Batchelor, J. Fluid Mech. 74 (1976) 1.
28. J.-Z. Xue, E. Herbolzheimer, M. A. Rutgers, W. B. Russel and P. M.
Chaikin, Phys. Rev. Letts 69 (1992) 1715.
29. N.F. Carnahan and K. E. Starling, J. Chem. Phys. 51 (1969) 635.
30. D.L. Koch, Phys. Fluids A 2 (1991) 1711.
1297

31. P. M. V. Resibois and M. de Leneer, Classical Kinetic Theory of Fluids


(Wiley-Interscience, New York, 1977).
32. G.K. Batchelor, J. Fluid Mech. 193 (1988) 75.
33. E.C. Eckstein, D. G. Bailey and A. H. Shapiro, J. Fluid Mech. 79 (1977)
191.
34. D. Leighton and A. Acrivos, J. Fluid Mech. 177 (1987) 109.
35. F. Gadala-Maria and A. Acrivos, J. Rheol. 24 (1980) 799.
36. R. J. Phillips, R. C. Armstrong, R. A. Brown, A. L. Graham and J. R.
Abbot, Phys. Fluids A 4 (1992) 30.
37. P.R. Nott and J. F. Brady, J. Fluid Mech. 275 (1994) 157.
38. Y.A. Buyevich, Chem. Engng Sci. 51 (1996) 635.
39. I.M. Krieger, Adv. Colloid Interface Sci. 3 (1972) 111.
40. W. B. Russel, D. A. Saville and W. R. Schowalter, Colloidal Dispersions
(Cambridge Univ. Press, Cambridge, 1989).
41. Y. Buyevich and I. N. Shchelchkova, Inzh.- Fiz. Zh. 33 (1977) 872.
42. G.K. Batchelor, J. Fluid Mech. 131 (1983) 155.
43. Y.A. Buyevich and A. O. Ivanov, Physica A 192 (1993) 375.
44. S. V. Bushrnanova, Y. A. Buyevich and A. O. Ivanov, Physica A 202
(1994) 175.
45. Y.A. Buyevich, J. Fluid Mech. 49 (1971) 489.
46. R. Jackson, Trans. Instn Chem. Engrs 41 (1963) 13.
47. D. Gidaspow, Multiphase Flow and Fluidization (Academic Press, Boston,
1994).
48. Y. A. Buyevich and S. K. Kapbasov, in: Multiphase Reactor and
Polymerization System Hydrodynamics (Gulf, Houston, 1996), p. 119.
49. R.E. Caflish and J. H. C. Luke, Phys. Fluids 28 (1985) 759.
50. E.J. Hinch, in: Disorder and Mixing (Kluwer, Boston, 1985), p. 153.
51. A.J.C. Ladd, Phys. Fluids A 5 (1993) 299.
52. H. Nieolai and E. Guazelli, Phys. Fluids A 7 (1995) 3.
53. H. Nicolai, B. Herzshafi, E. J. Hinch, L. Oger and E. Guazelli, Phys. Fluids
A 7 (1995) 12.
54. D.L. Koch and E. S. G. Shaqfeh, J. Fluid Mech. 224 (1991) 275.
55. A.J.C. Ladd, Phys. Fluids A 9 (1997) 491.
56. Y.A. Buyevich, Fluid Mech. Res. 22 (1995) 41.
57. A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics, vol. 2 (MIT
Press, Cambridge, MA, 1971).
58. Y.A. Buyevich, Chem. Engng Sei. 26 (1971) 1195.
59. M. Roco, in: Encyclopedia of Fluid Mechanics, Vol. 10 (Gulf Publ. Co.,
Houston. 1991), p. 1.
60. Y.A. Buyevieh, ASME FED-Vol. 231, MD-Vol. 66 (1995) 107.
1299

RHEOLOGICAL PROPERTIES OF CONCENTRATED


SUSPENSIONS

P.J. Carreau, P.A. Lavoie, and F. Yziquel

Center for Applied Research on Polymers, CRASP, Department of Chemical


Engineering, Ecole Polytechnique, Montreal, QC H3C 3A 7, CANADA

1. I N T R O D U C T I O N

Concentrated suspensions and multiphase polymeric systems are encountered


in many industrial sectors. Paints, foodstuffs, pulp and paper, mineral slurries,
filled elastomers and reinforced plastics illustrate very well a large sector of
applications. The processing/morphology/property relationships in suspensions and
in immiscible blends remain poorly understood and a given processing strategy
may result in high value-added products. The dispersion of fillers, the orientation
of fibers as well as the morphology of blends are strongly dependent on the
rheological properties of the components. Rheological properties are needed to
understand the phenomena encountered and changes occurring during processing.
Rheological data are also needed to assess constitutive equations required for
designing equipment and predicting changes under processing. Finally, rheological
methods could be powerful tools to establish relationships between structure,
formulation, processing and properties.
Most industrial suspensions show shear-thinning and thixotropic behavior.
Highly concentrated suspensions do not exhibit the low shear (Newtonian) plateau
depicted by homogenous polymer solutions or melts. The unbounded viscosity or
solid-like behavior at low shear rates or frequencies is the result of particle-particle
interactions and it is referred to as a yield stress. The notion of yield stress is quite
controversial and obtaining a significant value is very difficult, mostly for high
viscosity matrices such as polymer melts for which the high values of stresses mask
1300

the weaker contribution of the particle-particle interactions. Other physical aspects


are related to the degree of solids dispersion in the suspending fluid or matrix. For
highly concentrated suspensions, the solids are not, in general, evenly distributed
and aggregates and flocs may be present, forming a structure which may change
with time due to forces acting on the system. A second aspect is related to the
viscoelastic nature of suspensions. For concentrated suspensions in Newtonian
fluids, viscoelastic effects have been observed only when yield stresses are
detected. Fillers in viscoelastic polymers are believed to reduce the elasticity of
polymers [ 1], but this remains unclear as the definition of elastic depends on the
experiments used to obtain the appropriate data, for example primary normal stress
differences in simple shear flow or elastic modulus in small amplitude oscillatory
shear flow. Measurements in the linear viscoelastic domain, such as small
amplitude oscillatory shear flow, can yield very useful information on the
suspension structure and interactions between the various components. For small
enough strain, the structure of multiphase polymeric systems is expected not to
change and the linear viscoelastic functions (the storage modulus G' and the loss
modulus G") are enough to describe the equilibrium properties. Results of
measurements, however, in the so-called non linear viscoelastic domain, although
very complex, are essential for understanding and predicting morphology changes
during processing.

1.1 Scope
In this chapter, we review the rheological behavior of multiphase systems
(suspensions in Newtonian fluids and polymers filled with or non interactive
particles). First, we present some basic definitions and concepts useful for the
understanding of the rheological properties of concentrated suspensions. Then, we
summarize problems encountered in rheometry, and review the main rheological
models proposed in the literature to describe the properties of concentrated
suspensions. Finally, we focus on non linear effects observed for concentrated
suspensions in small strain experiments using data of our laboratory, and conclude
with challenges and trends for future research.

1.2 Forces acting on particles in suspensions


Various forces affect the rheological behavior of concentrated suspensions,
especially in the case of colloidal (submicron) particles for which Brownian forces
1301

and particle-particle interactions play a major role. The internal forces can be
subdivided in hydrodynamic and non hydrodynamic forces. The hydrodynamic
forces result from the relative motion of the suspending medium with respect to the
particles. These are the only forces considered in the Einstein [2] analysis of the
viscosity of a dilute suspension of rigid spheres in a Newtonian fluid, which led to
the well known result:

]]r - - 1 + 2.5ff (1.1)


11 O

here fir, rl~ and I"1oare the reduced viscosity, the suspension and the fluid viscosity
respectively. The hydrodynamic forces are responsible for migration of particles,
alignment or orientation in the case of non spherical particles and finally structure
break-down in the cases of flocs and aggregates.
The non hydrodynamic internal forces consist namely of the Brownian forces,
responsible for the internal motion of the particles and diffusion, and forces arising
from physical and chemical interactions. These forces can affect the internal
structure of suspensions of colloidal particles, for concentrations well below that
of the maximum packing. Brownian motion allows low range attractive forces to
promote floc building, network and weak gel formation. Such attractive forces are
characterized by the Hamaker constant defined by the Israelachvili equation as
reported by Tsai and Ghazimorad [3]:

A 3 (c, c3/
- + +
3hVe , 2 23 (1.2)

4kT el +e3 16v/-2(n?+n2)

where k is the Boltzmann constant, T is the absolute temperature, c~ et c 3 are the


dielectric constants of the particles and the fluid respectively, n/and n 3 are the
corresponding refractive indices, h is the Planck constant and v~ is the absorption
frequency. For a given concentration, floc formation increases with increasing
value of the Hamaker constant [4]. Flocs can be destroyed under shear stresses,
increasing the shear-thinning behavior. The power-law index has been shown to
decrease with solids concentration. The results of Tsai and Zammouri [4] are in
agreement with those of Gadala-Maria and Acrivos [5], who attributed the
increasing shear thinning to a reduction of the distance between particles with
1302

increasing solids concentration.


Electrostatic repulsion can occur when particles are superficially charged. These
long range interactive forces are characterized by the zeta potential (~) obtained,
for example, from electrophoretic mobility. In ionic fluids, a double ionic layer
masks the repulsive field. Moreover,when the surface is saturated the global charge
is zero and the repulsion range is reduced to the thickness of the double ionic layer,
l D (Debye length). The interesting results obtained by Krieger and Eguiluz [6]
show a drastic reduction of the viscosity for a 40% volume suspension of 0.11 ~tm
polystyrene beads with increasing ionic content (HC1). These results are presented
in Section 3. In pure water, the suspension shows a yield stress whereas at ionic
saturation, Newtonian behavior is observed. Leong and Boger [7] have studied the
influence of the surface chemistry of different coals. The addition of electrolytes
was shown to increase the suspension viscosity until a yield stress appears, in
contrast to the findings of Krieger and Eguiluz [6].
Tsai and Viers [8] have investigated the effect of polarity of non-aqueous
solvents on the rheological properties of graphite and polystyrene suspensions. The
effect was found to be important for the graphite suspensions, but negligible for the
polystyrene systems. This was explained by the large differences in the Hamaker
constant which was found to be ten times larger for graphite particles than for
polystyrene. Tsai and Ghazimorad [3] have correlated the power-law index of the
shear viscosity vs. shear rate and the Hamaker constant for systems of various zeta
potentials. Newtonian behavior was observed when the Hamaker constant was zero
and the zeta potential negligible. Suspensions with significantly non-zero zeta
potential depicted a Newtonian behavior for the higher Hamaker constant values.
That is: Newtonian behavior results from an equilibrium between attractive and
repulsive forces. Repulsive forces can be induced by polymer stabilization with
polymer chains adsorbed on particle surface, which interact to form a network
stabilized by steric repulsion. This is comparable to electrostatic repulsion. Mewis
et al.[9] have examined the rheological properties of suspensions of PMMA
particles stabilized with polyhydroxystearic acids grafted on surface. Their main
conclusion was that the deformation of polymeric layers during flow affects the
rheology of the suspensions. As expected, the effect was more important for
smaller particles with a smaller ratio of particle diameter to stabilization layer. The
critical ratio below which hard sphere assumption is no longer valid is somewhere
between 5 and 30. For a critical ratio of five, the maximum packing was calculated
1303

via the Krieger-Dougherty model to be 0.72 and 0.96 using the values for the zero
shear viscosity, rio, and the high shear rate limiting viscosity, rl=, respectively. The
corresponding values for a suspension of spheres with a ratio 60 were 0.49 and
0.62.
External forces may affect the behavior of suspensions. Buoyancy effects
become important when the particle and the fluid densities are significantly
different. These effects are responsible for sedimentation or flotation, affecting the
stability of the suspensions and resulting in concentration profiles. As illustrated
in the next section, apparent thixotropy can be related to sedimentation. Other
external forces such as that due to an electric field can drastically modify the
rheological properties of suspensions. Table 1.1 summarizes the main forces acting
on colloidal particles and key expressions for calculating their order of magnitude
are presented.

Table 1.1 Forces acting on colloidal suspensions (Adapted from Russel et al.[ 10])

Force Order of magnitude Definitions

kT k: Boltzmann constant
Brownian dP T: temperature [K]
dp: particle diameter

van der Waals A/dp A" Hamaker constant

Viscous rl" fluid viscosity


U: relative velocity of particles

Inertial pp" density of particles

A 9" density difference between


Gravitational dp3Apg fluid and particles
g: gravitational acceleration
1304

1.3 Yield stress


The yield stress is a useful rheological parameter to characterize particle-
particle interactions in suspensions. Classically, the yield stress (Oo) is the stress
value below which a material cannot deform. The simplest model to describe fluids
exhibiting a yield stress is the Bingham model [ 11 ]. Other well known empirical
models are the Casson equation [ 12], the Herschell-Bulkley model [ 13] for power-
law fluids, and the equation of Poslinski et al.[ 14] for Carreau fluids. A better
definition is to consider two distinct deformation modes, as proposed by Oldroy
[ 15]" first an elastic-solid behavior for stresses below the yield value, then followed
by a fluid behavior. Yoshimura and Prud'homme [ 16] and Doraiswamy et al.[ 17]
used the same concept. For example, the Herschel-Bulkley model can be written
as:

o : - G ov, [o1 < o ~ : G ovo


o (1.3)
o = - ( o + m I~t[(~-l)) ~t, Iol > o o

where Go is an elastic modulus. In proposing this model, Doraiswamy et al.[ 17]


assumed that the structure was not modified at low strain (ITI -< To). The
experimental results of Gadala-Maria and Acrivos [5] and our results shown in
Section 4 do not verify this assumption. G ' and G "data are clearly non-linear
functions of strain. The concept of s t r a i n - i n d u c e d structure can partially explain
yield stresses observed by Husband and Askel [ 18] for non colloidal dispersions
in polymeric matrices. Finally, we stress that if many experimental methods
(mostly based on extrapolation of shear rate data to zero shear rate) have been
proposed to determine the yield stress, the determined value depends strongly on
the rheological method or model used [19]. Difficulties in measurements as
discussed below may be other sources of errors.

2. D I F F I C U L T I E S AND MEASUREMENT PROBLEMS

Many difficulties arise in the determination of the rheological behavior of non


homogeneous media. The following criteria are useful in assessing potential
difficulties.

2.1 Basic criteria


A first simple criterion is related to the dimension of the suspended particles
1305

with respect to the characteristic dimension of the measuring element (in cone-and-
plate geometry, gaps of less than 1O0 gm are common)"

dp
<< 1 (2.1)
h

where dpis a characteristic dimension of the particles and h is the gap of the
measuring device. Obviously it is not always possible to respect this criterion and
systematic experimental errors could result.
Inertial and gravitational effects must be negligible. The criteria for both effects
to be negligible can be written as"
4vp
<< 1 (2.2)

and:

texp d2plPP-Pmlg <, 1 (2.3)


h 1]

The first criterion (Equation 2.2) just expresses that the Reynolds number based on
the particle and the average velocity of the fluid, V, has to be very small (creeping
flow regime). The second one relates the experimental time, t~ to a characteristic
time of gravitational or buoyancy effects. With large size particles, other effects are
to be expected such as due to a depletion layer at the wall and/or orientation along
the wall surface in the case of non spherical particles resulting in lubrication effects
or apparent slip.

2.2 Effects of dispersion


During measurement or processing, the degree of dispersion of the solids or the
structure due to particle-particle interactions may be considerably modified with
time. An example of the strong influence of the mixing history of the fillers on the
complex viscosity of a suspension of ruffle (TiO2) particles in a low molecular
weight polybutene (liquid at room temperature) is illustrated in Figure 2.1. The
untreated rutile particles of average diameter equal to 0.6 gm were first mixed to
the polybutene using a Brabender plasticorder until an equilibrium torque was
reached. Then a second operation was applied to some samples, which consisted
of mixing by hand using a thin spatula up to five hours. The behavior of the
1306

complex viscosity curve and its value are drastically changed following the second
step mixing. The power-law behavior with a slope close to - 1 for this 31 vol %
suspension after mixing using the Brabender,is indicative of strong interactions

10 7

~ ' ~ alter 1st mixing


10 6
" O " after 2"d mixing
c~ 10 s

..x- 10 4

10 3

Maron -Pierce ~m= 0.68 ..I


10 2
. . . . . . . . i i i i IiiIll t i i iiilll i i i ii1'11 . . . . . . . . i . . . . . . .

lO-a 10 -2 10 -1 10 ~ 101 10 2 10 3
o [s-q

Figure 2.1 Complex viscosity for a 31 vol % TiO2 in polybutene at 25 o C (From


Carreau et al. [20]).

100 9 9 9 . . . . . i 9 9 9 9 ..=.| 9 9 . 9 . . . . b

II~ ' 9 Manual


I:I Ultrasonic
r'--"l
r~

10-1

10-2

lO-I 10 ~ 101 10 2 10 3
[s-]l

Figure 2.2 Shear viscosity of 6 mass % suspension of fumed silica A200


particles of diameter equal to 12 nm after a manual mixing and after
ultrasonic dispersion (Unpublished data).
1307

between aggregates of rutile particles. The second mixing step broke down a large
portion of the aggregates and the behavior approaches that of a suspension made
of non imeractive particles. The complex viscosity remains, however, considerably
larger than predicted by a Maron-Pierce equation (Equation 3.1 with p = 2 and
assuming ~m = 0.68).
Another drastic effect of dispersion is illustrated in Figure 2.2 for a low
concentration suspension of fumed silica particles in water. These silica particles
have silanol groups at the surface which interact via water through hydrogen
bonding, forming very large aggregates and possibly a network. Following a
preparation via a manual mixing, the viscosity of an aqueous suspension containing
6 mass % of A200 (Degussa) particles (diameter of 12 nm) is shown in the figure
to be highly shear-thinning, typical of a gel-type structure which is broken down
under shear flow. However, the structure can be largely broken down prior the
rheological measurements using an ultrasonic vibrator. The values for the
suspension viscosity are drastically decreased and shear thinning is now very light.

2.3 Effects of adsorption or changes in ionic concentration


For molten polymer matrices, degradation, reentanglement and r are
frequently encountered. For suspensions in polymer solutions, the main problems
are associated with mechanical degradation, solvent evaporation and changes in
solubility of the polymer, i.e., shear induced crystallization and adsorption of the
polymer on the filler surface. Spectacular effects of polymer adsorption reported
by Otsubo [21 ] are shown in Figure 2.3 for suspensions of fumed silica particles
in a polyacrylamide solution. At the lower solids concentrations (up to 5 mass %)
the shear viscosities of the suspensions are shown to be smaller than that of the
suspending solution (0 % solids). The hydroxyl groups at the surface of the fumed
silica particles favor the adsorption of the polyacrylamide chains. Otsubo has
estimated that the adsorption layer is of the order of 8 to 10 nm, which is
comparable to the size of the silica particles. These specular effects could possibly
be caused by other factors such as changes in the ionic content (an example of such
effects is illustrated next). At high solids concentration, the effect of dilution due
polymer adsorption is overcome by the effect of solids loading and the viscosity
of the suspensions becomes considerably larger than that of the suspending fluid.
The behavior becomes quite complex. The rapidly increasing viscosity with
decreasing shear rate at low shear rates is indicative of a gel-type behavior and the
shear thickening at high shear rates suggests flow induced structure.
1308

103
13%
10
~_~ 10 2 7,~

2"~
~'1 0] 3-~

100 9 , I I

10 -3 10 -2 10 -! 10 0 10 l 10 2
[S "1]

Figure 2.3 Shear viscosity of suspensions of fumed silica particles in a 1.5 %


mass polyacrylamide (PAA) solution in glycerine. The average
particle size is 20 nm and the weight average molecular weight of the
PAA is 2.1 x 106 kg/kmol. The particle concentrations are in mass
percent (Adapted from Otsubo [21 ]).

Figure 2.4 reports simple shear viscosity data for suspensions of glass beads
(average diameter of 49 ~tm) in a 1 mass % aqueous solution of gellan (a
polysaccharide of molecular a weight of approximately 10~ kg/kmol). The unfilled
gellan solution is shown to behave as an homogeneous polymer solution,
describing a Newtonian plateau at low shear rate followed by a typical power-law
behavior. Adding 4.76 % per volume of glass beads did not change much the
rheological properties. The viscosity in the plateau region for the 4.76 %
suspension is, however, slightly lower than that of the suspending solution. Adding
more glass beads resulted finally in a large increase of the viscosity in the plateau
region, but strangely enough the viscosity is shown to be unaffected by the solids
content in the power-law region. To discriminate between the effects of polymer
adsorption at the glass bead surface (believed to be negligible here) and effects of
ionic content, we did measure the pH of the solution. It was found to increase from
6.79 for the unfilled solution to 11.53 for the 37.5 % suspension. Obviously, the
chemicals used for the surface treatment of the glass beads interact strongly and
change the polymer conformation in solution. Similar effects have been reported
previously by Carreau [22]. Figure 2.5 shows that the viscosity of the unfilled
aqueous solution of gellan is considerably reduced by increasing the pH of the
solution from 6.8 to 11.4.
1309

10 3 7 ~ .... m ........ m ........ m . . . . . . . "m ........ u ........ m


9 0% glass beads pH = 6.8
*~, [] 6.6% pH=10.5
**, "*._ * 14.6% pH = 10.9
10 2
9 9 ~ 9 28.5% pH=11.3
- A 9149149149 9 37.5% pH=ll.5
101

100

10-] ........ m ........ n ........ m ........ m ........ m . . . mi~mmm]

10-3 10 "2 10] 10 ~ 101 10 2 10 3


[s,]
Figure 2.4 Simple shear viscosity of glass bead suspensions in a 1.0 mass %
gellan solution in water (Unpublished data).

101
O001~ 9
. . . . . . . . I . . . . . . . . I . . . . . . . . I

mmlmmmmmm m 9
[] 9

r 10 0
9 9
9 pH = 6 . 8
9 pH =11.4 9 mo
9 9
lO-1
9 ..mm!
. . . . . . . . I . . . . . . . . I I I ...... I . . . . . . . . I

10 -2 1O 1 10 ~ 101 10 2 10 3
~/ [S -1]

Figure 2.5 Simple shear viscosity of unfilled aqueous 1 mass % gellan solutions
for pH values of 6.79 and 11.4 (Unpublished data).

2.4 Buoyant or gravitational effects


Buoyant or gravitational effects could lead to results which resemble typical
shear-thinning or thixotropic effects. This is demonstrated in Figure 2.6 taken from
Acrivos et a1.[23] for two types of suspensions. The first one consisted of
polymethy methacrylate particles in a mixture of glycerine and water. The particle
1310

diameters ranged from 125 to 150 ~tm and the density differences between that of
the particles and fluid were 0.03 and 0.04 g/mL. The second one consisted of
acrylic particles (diameter between 75 and 106 pm) in a Dow Coming fluid with
a density difference of 0.07 g/mL. A double gap Couette geometry was used for
measuring the viscosity. In the figure, the reduced viscosity as a function of a
reduced shear rate in steady-state conditions is reported for four different solids
volume fractions. The reduced shear rate is based on the fluid viscosity and gives
an order of magnitude of the ratio of viscous to buoyancy forces; ho is the height
of fluid in the Couette geometry. The apparent shear-thinning effects observed in
Figure 2.6 have nothing to do with shear-thinning properties of the suspensions.
Initially due to the small differences in density, sedimentation proceeds leading to
higher stresses in the lower part of the Couette geometry. But at higher stresses
particles are re-suspended by shear-induced diffusion and the particle concentration
reaches an equilibrium profile, which is a function of the applied shear rate. The
solid lines in the figure are the predictions of the model developed by Acrivos et
a1.[23] to account for sedimentation and shear-induced effects.

102
a, r 050

~- 101 0.40

0.30

10~ [_.,.,.,
10 -3 10 -2 10 -1 100 101 102
9rl~'
2h0gAp

Figure 2.6 Apparent reduced shear viscosity for suspensions of non colloidal
particles in Newtonian fluids, due to density differences between the
particles and the suspending fluid. The open symbols are for the first
suspensions; the close symbols are for the second type of suspensions
(Adapted from Acrivos et al. [23]).
1311

These simple results stress how careful one has to be in interpreting unusual
phenomena. Many other difficulties may be encountered, especially when using
polymer matrices. Other major problems are related to polymer degradation when
experimenting with polymer melts at high temperature. In the case of polymer
solutions, solvent evaporation could be a major nuisance as indicated by Acrivos
et a1.[23] when long experimental times are required to reach, for example, steady
conditions. For highly filled systems, it may just be impossible to make reasonable
measurements especially at high shear rates due to wall effects, fracture in the
suspensions, and other complications.

3. M O D E L I N G

In this section we discuss key results of the literature illustrating various


behaviors observed for non interactive and interactive particles in Newtonian and
non-Newtonian matrices. Useful empirical viscosity expressions to describe
properties of concentrated suspensions, as well as recent rheological models for
characterizing thixotropic effects, are presented.

3.1 Hard sphere suspensions in Newtonian fluids


Following the theoretical analysis of Einstein [2] for creeping flow of very
dilute suspensions of spheres, various phenomenological models have been
proposed to predict the rheological behavior of non dilute suspensions of hard
spheres in Newtonian fluids. In the hard sphere concept, the spheres are assumed
to be ideal spherical particles with a Dirac interaction potential. These assumptions
lead to only hydrodynamic and Brownian forces. The Einstein theory neglects the
effects of other particles (particle-particle or/and hydrodynamic interactions).
Interactions with neighboring particles is accounted for in a first approximation by
a quadratic term with respect to the solids fraction, as proposed by Batchelor [24].
Krieger and Dougherty [25] modified the Mooney [26] model to account
differently for crowding effect. The relative viscosity rl~ as a function of the volume
fraction ff is then given by the following simple expression:

(l m) -p
(3.1)
1312

where p = (l)m[]]], (])m is the maximum packing fraction and [11] the imrinsic
viscosity. The Krieger-Dougherty relation reduces to the Maron and Pierce [27]
expression for p = 2, which corresponds to the theoretical case of spherical particles
with [11] equal to 5/2 and ~ = 0.80. However, (I)m is usually kept as a fitting
parameter.
The viscosity models for concentrated suspensions are usually expressed as a
function of the maximum packing fraction ~ , which is a function of shape and size
distribution of the particles. The values reported in the literature range from 0.18
for carbon fibers with shape factor L/dp = 27 to 0.71 for uniform hard spheres.
Chang and Powell [28] have investigated the effect of size distribution using
particles of two well-defined sizes. The results were reported in terms of two
parameters K, the diameter ratio of the small particles, and r the volumetric
proportion of small particles. Their results and those of Storms et al. [29], Poslinski
et al. [ 14] and Shapiro et al. [30] show that for a fixed volume fraction, the relative
viscosity decreases as K decreases and a minimum in viscosity is observed around
= 0.25. The viscosity reduction could be quite spectacular as shown by Poslinski
et al. [14].

103[ .... I

102 ~t_,, 0

10 1

10~ I I I I I i
0.1 0.2 0.3 0.4 0.5 0.6

Figure 3.1 Relative viscosity versus dispersed phase volume fraction for latex
and silica suspensions (combined data of Krieger (1972) and de Kruif
et al. [31]). The solid lines are the best fits using the Krieger-
Dougherty equation (3.1) (Adapted from Barnes et al. [33]).
1313

The value of the maximum packing fraction also depends on the arrangement
of particles. The maximum packing value for monodisperse particles ranges from
0.52 for cubic arrangement to 0.74 for a hexagonal close packed one. The
maximum packing organization varies with flow: de Kruif et al. [31 ] and Krieger
[32] have shown that the maximum packing fraction and the intrinsic viscosity [1]]
vary with shear rate. Figure 3.1 reports the reduced viscosity as a function of the
volume fraction for the combined experimental data of de Kruif et al. [31 ] for silica
suspensions and of Krieger [32] for latex suspensions. The best fit of the Krieger-
Dougherty equation allows to obtain at high shear rates ~m== 0.71 and [rl]= = 2.71
and at low shear rates d~mo = 0.63 and [1]]o = 3.13.
Since only the hydrodynamic and the Brownian forces govern the behavior of
suspensions of colloidal hard spheres, the rheological properties can be scaled as
a function of the radius of the particles, a, the Boltzmann constant, k, and the
temperature T. Krieger [32] proposed a semi-empirical relation based on the Cross-
Williamson [34] model to describe the shear dependence for the viscosity of
concentrated suspensions viscosity. The viscosity is defined by the following
relation:

rl -rl~ 1
(3.2)
11o-r1= 1 +o/o c

26
O BZOH
22- -"~ 9 m-CRESOL "

~"18

14

10 ' , ,
]0-2 ]0-I ]0 0 ]01
O"r
Figure 3.2 Reduced viscosity versus reduced shear stress for 50 vol %
monodispersions of polystyrene latex particles in three different
media (Adapted from Krieger [32]).
1314

This equation describes shear-thinning effects from a zero shear rio to a high shear
limit rl=; Oc is a critical reduced stress for which the reduced viscosity,
(rl - rl=)/(rlo- rl=) - 1/2; %(- oa3/kT) is a reduced shear stress. Figure 3.2 reports the
reduced viscosity as a function of the reduced stress for suspensions containing
particles of two different diameters dispersed in two different media. A unique
curve is obtained, independent of particle sizes and of nature of the suspending
medium.

I I I I

101

rim -~--O
O

l0 0 .

10 -l o / a3G '
kT

[] []

10 -2

10 -3 m

10 "4 I I I

10-1 10 0 101 10 2 10 3
a20~
Do
Figure 3.3 Dimensionless elastic modulus and reduced dynamic viscosity versus
dimensionless frequency for a 46 vol % suspension of silica spheres
in cyclohexane- ffl a = 28 2 nm; (3 a - 76 + 2 nm. The solid lines
are the predictions using the Jeffreys model (Adapted from Mellema
et al. [35]).

Mellema et al. [35] have investigated the linear viscoelastic properties of


suspensions of silica spheres of two different radius values (28 and 76 nm)
dispersed in cyclohexane. They obtained master curves for a dimensionless elastic
1315

modulus G 7cT/a 3 and the reduced dynamic viscosity rl'lrlm as a function of a


dimensionless frequency eoa2/Do where Do is a characteristic diffusivity. The results
are illustrated in Figure 3.3. The dimensionless elastic modulus is shown to be a
quadratic function of the frequency at low frequencies and reaches a plateau at high
frequencies. The reduced dynamic viscosity depicts slightly shear-thinning effects
between a low frequency plateau to a limiting high frequency value, which
corresponds to a high frequency viscosity larger than that of the suspending fluid.
A very similar behavior was observed for the steady shear viscosity. The data are
fairly well described by a simple Jeffreys model.
The electrostatic and sterically stabilized particles are assumed as a first
approximation to form a rigid stabilized layer. The scaling rules for hard spheres
suspensions, as Equation 3.1 or 3.2, can be used provided the thickness of the
stabilized layer is taken into account in calculating the effective radius of the
particles. In electrostatic stabilization, the thickness layer, the so-called Debye
length, is a function of the electrolyte concentration and of the medium. In the case
of steric stabilization, one must include the polymer layer thickness which depends
strongly on the affinity between the polymer and the suspending medium, as
characterized by the Flory-Huggins parameter X- This approximation is valid when
the compressibility of the layer is negligible, that is: for a thickness layer smaller
than the core particle radius and for moderately concentrated suspensions.
Frith et a1.[36] proposed the following modified Krieger equation to predict the
viscosity of soft sphere suspensions:
rl - rloo 1
(3.3)
13o - rloo 1 +(Or/Oc )m
where m is a material parameter. Good agreement between the model predictions
and the experimental data was obtained for suspensions of PMMA latex (sterically
stabilized) particles in Exol.
When the interparticle interactions become dominant, the rheological behavior
of suspensions deviates from the behavior of hard sphere suspensions. Krieger and
Eguiluz [6] observed that electrostatic repulsion between particles can drastically
modify the suspension viscosity at low shear rates and can eventually induce a
yield stress. Figure 3.4 reports the reduced viscosity as a function of the reduced
shear stress for suspensions of PVC latex particles dispersed in water containing
HC1 at different molar concentrations. The electrolyte had been added to suppress
electrostatic repulsion between PVC latex particles. The added counterions screen
1316

off the charges on the particle surface. When electrostatic repulsion is dominant,
an agglomeration of particles is induced. At low shear rates, the viscous forces are
not sufficiently large to break down the structure, responsible for the rapidly
increasing viscosity with decreasing shear rate and for the existence of a yield
stress. At high shear rates, the flow is controlled by the hydrodynamic forces and
the shear viscosity becomes approximately independent of the electrolyte
concentration.

10 6
O
1.9xl04 M HCI ~ O deionized
ut.I O ~ . - -
10 4 D ~o
s..
9 1.9x10"3M HCI O

10 2 9.4xl(I2M HCI 9
A~ m
%Oo
. $ gA~llLemm~
1.9xl0"2M HC! ' ~
I
10 ~
10 .2 10 o 10 2
O'r

Figure 3.4 Reduced viscosity versus reduced shear rate for 40 vol % suspensions
of PVC latex particles in water containing HC1 at various molar
concentrations (Adapted from Krieger and Eguiluz [6]).

3.2 Suspensions in non-Newtonian fluids


In industrial applications, non-Newtonian polymer solutions are frequently used
as the suspending fluids. Generally, adding particles to a shear-thinning fluid does
not alter the general shape of the viscosity curve. The viscosity curve is shifted
vertically upward and the onset of shear thinning occurs at smaller shear rates with
increasing solids loading, as observed by Nicodemo et al. [37]. The power- law
index does not significantly change with solids concentration [38]. For high solids
concentrations or low viscosity matrices, particle-particle interactions may become
significant resulting in an apparent yield stress. Such yield stresses have been
observed by Lobe and White [39] and Tanaka and White [40] for polystyrene melts
filled with carbon black, titanium dioxide and calcium carbonate. Poslinski et al.
[ 14] proposed a modified Carreau model to predict the rheological behavior of
filled polymers:
1317

0o 1] 0
rl - IYI + (1 +(t,~/)2) (1-")/2 (3.4)

where the zero shear viscosity and the characteristic time are described by Maron-
Pierce expressions"

- - (3.5)

and
-2
(3.6)

rim and tlm are respectively the viscosity and the characteristic time of the polymeric
matrix. Poslinski et al. [ 14] obtained good agreement between the proposed model
and the experimental data for glass spheres suspended in thermoplastic matrix, as
shown in Figure 3.5. In this case, the glass spheres are not sufficiently interactive
compared to the viscous forces and no apparent yield stress is observed. The
rheological behavior of the suspensions is very similar to that of the unfilled
polymer. The zero shear viscosity increases and shear thinning occurs at smaller
shear rates as the glass spheres volume fraction increases.

5
lO
9 0%
104 O13%
m26%
r~
t~35%
10 3 A46 %
~x60 %
~" 102

101

10~
10-2 10-1 10o 101 102 10 3 10 4
"~ [S "l ]

Figure 3.5 Shear viscosity of glass spheres at various volume fractions, dispersed
in a thermoplastic polymer at 150~ (Adapted from Poslinski et al.
[14]).
1318
Doraiswamy et al. [17] proposed an extension of the Cox-Merz rule for
concentrated suspensions in polymer exhibiting a yield stress. They have shown
that the steady shear viscosity and the complex dynamic viscosity of a concentrated
suspension of silicon particles in polyethylene can be superimposed if the complex
viscosity is reported as a function of the strain rate amplitude coy~ for oscillatory
shear data obtained in the non linear domain.

3.3 Time dependent and structural models


Most industrial suspensions consist of interactive particles: paints, coating
colors, inks,... The interactions between particles are responsible for the formation
of a structure which is time and shear dependent. Adequate experimental
characterization and modeling of the rheological properties of suspensions,
especially for interactive particles, are very difficult tasks. The reader can find a
rather complete review of thixotropic phenomena in [41 ]. Here, we summarize the
main rheological models which are capable of describing yield stresses and
thixotropic behavior and present our most recent ideas on the subject. The
rheological behavior of suspensions made up of interactive particles is governed
by a competition between micro-structure break-down due to flow and build-up due
to collisions between particles induced by Brownian motion and flow. Hence, the
evolution of the micro-structure is generally described by a kinetic equation of the
following form [41 ]:

- a(1-~)b-c~'~a (3.7)
at

where ~ is the structural function and a, b, c and d are constants.


To describe thixotropy, Quemada [42] used a Krieger-Dougherty expression:

l"lr (1
- - (D /-2 (3.8)
~)m(~)
where the maximum packing factor is given by the following function of the
structural parameter:

~),,,(r _ 1
( 1/~mo - 1/~,,,~,)+ 1/~,,,oo (3.9)
1319

Here the maximum packing factor varies from (~moat low shear rates to (I)mooat high
shear rates. The structural function ~ is obtained by solving the following kinetic
equation:

(3.10)
0t ta td

where t~ and td are characteristic times associated with the aggregation and
desegregation process respectively. The ratio t~/td is assumed to be a function of
the shear rate as follows:
a
t _

(3.11)
ta
where p ( > 0) is an empirical parameter and tc is a characteristic time.The
following expression is obtained for the steady shear viscosity:

1 +(tc~)P ]
(3.12)

where 1]o and ~1=are the limiting viscosity at low and high shear rate respectively.
Good agreement was obtained between the proposed model and experimental data
for blood and latex and silica particle suspensions.
Quemada [43] also proposed the following Jeffreys-type model to describe the
non linear viscoelastic behavior of concentrated suspensions"

1 Oo + ~ 1o - - ~t - ~,(~) O~t (3.13)


G(~) at rl(~) Ot

where G is a modulus and ~ a relaxation time, both functions of the structural


parameter, ~.
De Kee et al. [44] have described the viscosity of yogurt and mayonnaise using
a model based on a structure kinetics. The structural function is obtained by solving
the following equation"
1320

-C [~ l d( ~ - ~equiZ) ~ (3.14)
at

The shear viscosity is then given by the relation"

O - ~( - "~E'qpe -tp ~ ) (3.15)


Leonov [45] has developed a model to describe yield stress and thixotropic
phenomena in filled polymers. He assumed that the total stress contains two main
contributions: a viscoelastic contribution from particle interactions and a viscous
one from the suspending medium"

o - op + o m (3.16)

The total stress tensor consists of a contribution due to particle-particle


interactions, Op, and that of the suspending polymer matrix, o~, considered to be
filled with inert particles. The Leonov model developed for homogeneous polymers
is used to describe this last contribution. A kinetic equation is proposed to describe
floc rupture and build-up in terms of characteristic times. Floc break-down is
related to a critical elastic deformation. Leonov have shown that his model can
describe thixotropic effects in stress growth and relaxation experiments, but it has
been assessed using only steady shear (viscosity and primary normal stress
difference) and and elongational viscosity data for filled polymer melts. Good
agreement has been shown, as evidenced here for the steady shear viscosity of two
filled systems. Figure 3.6 compares the model predictions to the steady shear
viscosity as a function of shear rate for carbon black particles in polystyrene at
various volume fractions (data of Tanaka and White [40]). In Figure 3.7, the
corresponding data for a molten polystyrene filled with calcium carbonate particles
of different sizes (data of Suetsugu and White [46]) are compared to the model
predictions. Figure 3.7 shows that the Leonov model is able of predicting the
drastic increases of the shear viscosity with decreasing size of the particles at
constant solids volume fraction, ~ = 0.3 (comparable results have been reported for
the effective elongational viscosity [45]). However, these data do not allow to test
the model ability to describe time-dependent or thixotropic effects.
1321

106
PS/CB

5
~10
20%
10%
104 ......
10 -2 10-1 10o
'y [S-1]

Figure 3.6 The predictions of the Leonov model for the steady shear viscosity
compared with the data of Tanaka and White [40] for polystyrene-
carbon black (dp ~ 0.045 pro) compounds at various volume fractions
(Adapted from Leonov [45]).

107
PS/CaC03 [

106 1
z~ <--0.07 p m
~ - 10 5 9 _ -~'a~~ -~-" <--0.5 p m
9 9 - " " ' ~ . - - - - . ~ <-- 3 lam'

104 I
10-2 10-1 10o
? [s-l]

Figure 3.7 The predictions of the Leonov model for the steady shear viscosity
compared with the data of Suetsugu and White [46] for polystyrene-
calcium carbonate compounds with various particle sizes at constant
solids volume fraction, ~ - 0.3 (Adapted from Leonov [45]).
1322

Coussot et al. [47] proposed a model similar to that of Leonov to describe the
rheological properties of a granular grease and a clay platelet suspension. The
viscoelastic contribution is given by a modified Maxwell equation:

1 0o 1
P + ~ O - - ? (3.17)
G O at q(~) p

where Go is a constant elastic modulus and rl(~) is a viscosity function for the
viscous contribution of the particles, which depends on the structural factor ~,
determined from the following kinetic equation:

X--~- = (1 -E.j)--yc ---0 (3.18)

with

n
T](~) - T}p (3.19)
(1 - ~ ) - ' - 1

X is a kinetic parameter, yc a critical strain, 0 characteristic time, TIpa characteristic


viscosity and n an empirical parameter.
Yziquel et al. [48] proposed a different kinetic network model based on ideas
ofMarmcci et a1.[49] and Coussot et al. [47] to describe the non linear behavior of
fumed silica suspensions observed during oscillatory measurements. The stress is
described by a modified upper convected Jeffreys model with a single relaxation
time"

6 (3.20)
+ -- I + ~r - q~.
6t G(~) n(~) n(~) ~ G(~)

with

G(r = Gor + G= (3.21)

where Tl= and G= are respectively the viscosity and the elastic modulus for the
1323

suspension with a destroyed structure and Go + G~ is the equilibrium value of the


elastic modulus of the suspension; 6 /6t is the upper convected derivative.
Contrary to Coussot et al., we assume that the modulus depends on a structural
parameter as proposed by Marrucci et al. [49] and Quemada [43]; rl(~) is the
structure-dependent viscosity defined by the relation:

r I(~) - 1]~ (3.22)

1"1ois a characteristic viscosity and fl~) an empirical structural function chosen to


obtain solid-like behavior at very small strains:

1 / (1-n)/2
= 1 (3.23)
I
where ~, the structural parameter is related to the number of interparticle bonds
and ranges from 0 and 1. It is obtained by solving the following kinetic equation:

~'o0~ = (1-~) - k2 (3.24)


z 5-
o Yc

with

= rio (3.25)
0 Go +G oo

k I and ke are kinetic constants for the thermal build-up of the suspension
microstructure and for shear induced break-down respectively, )~ois a characteristic
relaxation time and I/, is the second invariant of the rate-of-strain tensor.
Figures 3.8 and 3.9 compare the experimental data for the elastic and viscous
moduli versus strain amplitude obtained for suspensions of fumed silica particles
in paraffin oil and the values predicted by the Yziquel et al. [48] model. Two types
of particles were used: Aerosil A200 (Degussa) particles have a diameter of 12 nm
and are hydrophilic with surface silanol groups that can participate in hydrogen
bonding. Aerosil R974 (Degussa) particles have also a diameter of 12 nm but are
1324

hydrophobic: a significant fraction of silanol groups have been substituted by


methyl groups. Since the paraffin oil has no hydrogen bonding capacity, the
particles can only interact between themselves. These interactions are stronger for
the hydrophilic particles than for the hydrophobic fumed silica which has less
silanol groups on their surface. The elastic modulus is shown to decrease rapidly
above a strain exceeding 0.03 for A200 particles (Figure 3.8) and 0.008 for R974
particles (Figure 3.9) whereas the loss modulus increases markedly with strain
amplitude, but the increase is less important for the R974 suspensions. For these
suspensions, elastic effects dominate viscous ones, but the A200 suspensions have
the highest elastic modulus (stronger network due to the hydrogen bonding). The
moduli for these suspensions were found to be almost independent of the frequency
[48]. Note that the moduli reported in the non linear domain are apparent values
since the output signal is no longer a perfect sine wave. The model accurately
predicts the elastic modulus decreases with strain as well as the initial increases of
the loss modulus with strain. However, the model predicts subsequent decreases of
the loss modulus at a strain that is smaller than experimentally observed. The model
parameters used to fit the data are listed in Table 3.1 (the parameter k~ was found
to be a non significant role on the predictions). As far as we are aware, no other
models in the literature are able of quantitatively predicting the initial increase of

105 [ - __ _----__ ...... i "~~"~".,.


G' / ' - - - - - - ---_-_- ~-~ i .....
1 04
9 7.0% ----- m o d e l ,,,,k,n~.~
" 9 8.2% ,,-" . . " 9 ""
Q~ " 10.0% " -
~.~" 1 0 3 " 11.5%

102
10 .3 10 .2 1 o-1
~,o

Figure 3.8 Elastic and loss moduli as a function of stain amplitude for
suspensions of A200 fumed silica particles in paraffin oil at different
solids loadings. The solid lines are the Yziquel et al. model
predictions. (Adapted from Yziquel et al. [48]).
1325

105 . . . . . . . . . . . . . . . . . . . . . . .

?
G
104

??
G
103
. . . . | . . . . . . . . ! . . . . . . . .

10.3 10 -2 10-1
3,o

Figure 3.9 Elastic and loss moduli as a function of stain amplitude for
suspensions ofR974 fumed silica particles in paraffin oil at different
solids loadings. The solid lines are the Yziquel et al. model
predictions. (Adapted from Yziquel et al. [48]).

Table 3.1 M o d e l parameters used for describing the data o f Figures 3.8 and 3.9

Suspensions in Go + G= ~.o Y C
nJ(Go + G3 k, / kl n
naraffin oil . (kPa~ ....
, ,, ,
(s) . . . . (S)
7.0 % A 200 15.6 1.19 0.0320 0.0030 0.0085

8.2 % A 200 31.0 1.27 0.0320 0.0025 0.0075


,,

10.0 % A 200 64.2 1.83 0.0320 0.0020 0.0041

11.5 % A 200 140 2.07 0.0320 0.0020 0.0020

8.2 % R974 4.70 0.72 0.0090 0.0620 0.0270 0.08

10.0 % R974 11.5 0.72 0.0070 0.0620 0.0270 0.08

11.5 % R974 19.0 0.72 0.0066 0.0620 0.0270 0.08

13.5 % R974 32.5 0.72 0.0060 0.0620 0.0270 0.08


, , , ,,,

14.2 % R974 51.4 0.72 0.0057 0.0620 0.0270 0.08


1326

the loss modulus with strain. For example, the Coussot model [47] can describe as
well the decrease of the elastic modulus with strain reported here, but the
predictions for the loss modulus are considerably higher than the experimental
values [48].
As mentioned by Leonov [45], the main advantage of these approaches is that
no yield criterion is necessary to describe the transition from solid-like to liquid-
like behavior as in the Oldroyd model, discussed in Section 1.

3.4 Shear thickening


Shear thickening is defined as the increase of viscosity with shear rate. Most
concentrated suspensions exhibit shear thickening at high shear rates, but the onset
and the importance of shear thickening depend on the volume fraction, particle size
distribution and viscosity of the suspending fluid. The increase of viscosity is
attributed to a transition from a two dimensional layered arrangement of particles
to a random three dimensional form as confirmed by Hoffman [50] using light
diffraction and more recemly by Boersma et al.[51] using neutron scattering.
According to Laun et al. [52], this order-disorder transition cannot explain all cases
of shear thickening. Brady and Bossis [53] proposed that hydrodynamic clustering
may even explain shear thickening.
Shear thickening appears generally for volume fractions of solids larger than 50
% and the onset (shear rate at which shear thickening appears) decreases as the
volume fraction increases. The increase in viscosity depends strongly on the shape
and size of particles. According to Frith et al. [54], the data are not sufficiently
accurate to affirm that the onset of shear thickening is a quadratic or a cubic
function of the particle radius, but it seems that a cubic function is more appropriate
for hard sphere suspensions and a quadratic function for soft spheres. Collins et al.
[55] have noted that a large distribution of particle sizes can suppress shear
thickening. The shape of particles is also quite important and particle anisotropy
tends to produce shear-thickening effects at a lower shear rate and smaller solids
fraction. Shear thickening in suspensions of aggregated particles depends strongly
on the shear history. An example is given in Figure 3.10 which reports the viscosity
of a 8 mass % suspension of fumed A300 (Degussa) silica particles in
polypropylene glycol for different pre-shearing histories. The fumed silica are
colloidal particles (diameter of 7 nm) which could form aggregates through
hydrogen bonding via a polar solvent such as polypropylene glycol. Each curve
1327

presented in the figure was obtained using a controlled stress rheometer in the
viscosity mode and steady or pseudo steady viscosity is reported as a function of
the measured shear rate. A pre-shearing at different stress levels, but above the
maximum stress obtained in the viscosity mode, was applied to the sample
sufficiently long to reach steady states and erase previous shearing effects. Then
the subsequent viscosity curve was obtained. We observe in Figure 3.10 that the
low shear rate viscosity, the onset and the importance of the shear thickening
decrease with pre-shearing effect. At low shear rates, the behavior is slightly shear-
thinning. Above a critical shear rate, the viscosity rapidly increases due to
formation of aggregates. Under higher and higher stresses, the hydrodynamic forces
break down the aggregates and the hydrodynamic radius reduces with the
increasing stress level. Therefore, the viscosity is found to be considerably reduced
and the onset of the shear thickening occurs at a larger shear rate. Finally, we
observe that above a critical shear rate, the viscosity begins to decrease. This
subsequent shear thinning is attributed to a further reduction of the aggregate sizes
at higher shear rates.

101

.! , , 9 9 aiiOOOPa
c~

.9 YYvV9 9

' Wv ~Vv v V V v V V V v v v v v v v v v v v v W W Vv

l0 0 ! .... ~ ........ ~ ~ 9 ..~,d ~ ~ ...... ~ . . ~ .... ~ ........ I

10 .2 10 -1 10 ~ 101 102 103


Is-']

Figure 3.10 Pre-shearing effects on the steady shear viscosity for a 8 mass %
suspension of fumed silica (A300) particles in polypropylene glycol
(Unpublished data).

Few models are available in the literature to describe shear thickening. A


flexible one is the viscoelastic model proposed by De Kee and Chan Man Fong
[56], which can predict shear thinning, shear thickening and thixotropy or rheopexy
1328

depending on the choice of model parameters. The following viscosity expression


is obtained:

_ + -IX o l
rI rl~ (1 (bf~ - cfz) e ) (3.26)
tXo

with

r162 = k ( 1 + bf~ - cf2 ) (3.27)

where rio, kc, b, c are contants and f~ and fe are functions of the shear rate. For
(b fl - cf2 ) > 0, the model predicts thixotropy and for (b f / - c f2) < 0 rheopexy is
predicted. For steady state, the viscosity becomes"

n(?) - (3.28)
tz
o

Shear thinning or shear thicknening is predicted if ( b f ( - c f 2 / ) > O or < 0, where


fl/, f2/ are the derivatives with respect to the shear rate off1 and f2 respectively.
Although the De Kee - Chan Man Fong viscoelastic model appears to be quite
flexible, it cannot describe the pre-shearing effects reported in Figure 3.10.

4. L O W S T R A I N H A R D E N I N G PHENOMENON

In this section, we highlight non linear effects encountered at relatively low


strain with concentrated model suspensions and suspensions of industrial interest
(coating colors used in the paper industry).
The model suspensions discussed here consist of non colloidal particles
suspended in a viscous matrix, a polybutene (PB, Indopol H100 of Stanchem), of
low molecular weight, density equal to 890 kg / m 3 and viscosity of 24.5 Pa.s at
25 ~ This material is viscous enough so that the rheological properties are easily
measurable, but not too much in order to highlight mechanical interactions between
the particles. The industrial suspensions consist of coating colors used in the paper
industry. Kaolin particles used for these suspensions are smaller, in the form of
hexagonal platelets and show strong ionic interactions. Because of these strong
particle-particle interactions, elastic effects are much more important, but some of
the characteristics observed with the model suspensions are still detectable.
1329

4.1 Suspensions PVC particles in polybutene


The non interactive PVC (polyvinyl chloride) particles were suspended in a non-
polar viscous matrix PB. The powder was an industrial non formulated PVC, with
an average diameter of 12 ~tm and 50 % of the particles being smaller than 10 ~tm.
A second maximum could be observed at 3 ~tm on the particle size distribution.
Thus, the particle size distribution could be defined as just above the upper limit
of the colloidal domain. Their shape was slightly ellipsoidal with a small to large
axis ratio of 0.6. Suspensions at six different volumetric fractions (~ = 0.09, 0.18,
0.28, 0.37 and 0.47) were investigated, using a Weissenberg rheogoniometer
equipped with parallel plates of 1.2 mm gap, and a Bohlin VOR with a Couette
geometry. The results presented here are extracted and adapted from a previous
publication on theological properties of filled polymers by Carreau et al. [20].
The specific viscosity as a function of the shear rate for the PVC suspensions is
reported in Figure 4.1. For suspensions containing less than 30 vol % solids, the
data fall approximately on a unique curve. The horizontal solid line corresponds to
the theoretical value of 2.5 obtained from the Einstein analysis (Equation 1.1). This
value represents fairly well the data for the most dilute suspensions, mainly at high
shear rates. Using a maximum packing factor of 0.68, the Maron-Pierce equation
(3.1) is shown to describe well the data for the 0.38 and 0.47 vol % suspension
respectively at high shear rates. The shear-thinning effect obviously cannot be
described by this simple relation, nor can be related to settling effects as reported
by Acrivos et al. [23] and shown in Figure 2.6. These particles are non interactive,
large enough so that the viscosity decrease with shear rate could not be associated
with floc size variation nor to changes in the maximum packing fraction. The shape
factor of 1.6 could explain the slight shear thinning observed for the lower
concentrations but not the drastic effect observed for the 0.47 vol % suspension.
For this suspension, the viscosity becomes unbounded as the shear rate goes to
zero, indicating the presence of a yield stress and gel-type behavior, normally
observed for particles having strong physical or chemical interactions. However,
it is not the case here, but the yield stress is believed to be due to particle-particle
steric-type interactions, as discussed below.
As for time dependent models presented in Section 3.3, a structural parameter
can be defined and included in a kinetic equation to follow the particles
organization. With this approach, shear-thinning behavior (under steady state) is
essentially based on effective volume fraction. In other words, when structural
equilibrium is reached at a given shear rate, the structure is stable and can be
modeled using the maximum packing fraction as a single parameter. As shown in
Figure 4.2, the relative viscosity of each suspension is plotted against the volume
fraction for four steady shear rates" 0.156, 1.56, 15.6 and 39.3 s~. The solid lines
1330

100 . . . . . . . . , ,i. .... i..m .i..,....,

80
90 il
9 9 --
[]
Einstein's equation
Maron-Pierce(O., = . )
9% PVC / PB 1
0681
70 A 18%PVC/PB "~
~ - 60 _ Q 28% PVC / PB --I
, 50 "9 9V 437 % P V C / P B ~o 1d
40
30 mmmmmmmmImm~m_m
m
~=47
20
~ ~_ V
_ _" _ ~ ...~]E _ l"Itl?_'~..ll'..~ llt .ay_,~. , _ ~ _ 3-7~176~
10
o ...... , . . . . . . . . , 9 . 7 . . . . . , , ...7...#
10 -2 10-t 10 o 101 10 2
'~[S "1]

Figure 4.1 Specific viscosity for five different volume fractions of PVC/PB
suspensions vs. shear rate. The horizontal solid line at the value of 2.5
represents the Einstein result. The horizontal dashed lines are the
results calculated via the Maron-Pierce equation using a maximum
packing factor of 0.68 (Adapted from Carreau et al. [20]).

35
30 O 0 . 1 5 6 s -1
25 121 1.56 s -!
A 15.6 s-I
~. 20
V 39.3 s "j
15
10
5
0
0 10 20 30 40 50 60
r
Figure 4.2 Relative viscosity of PVC/PB suspensions for four different shear
rates as a function of the volume fraction. The solid lines represent the
Maron-Pierce model predictions using a maximum packing fraction
of 0.575, 0.629, 0.705 and 0.756 for the lowest to the highest shear
rate (Adapted from Carreau et al. [20]).
1331

represent the Maron-Pierce model predictions using the maximum packing factor
as an adjustable parameter. For each shear rate, the optimal values of ~)m equal to
0.575, 0.629, 0.705, 0.756 for each shear rate respectively. Hence, the steady state
viscosity of this PVC/PB system can be simply modeled by this Newtonian
empirical model using an adjustable maximum packing fraction or effective volume
fractions.

0.1
"~n' ...... ' '
! 9%
"<) . . . .4.5%r
. PVC/PB

18% _
28%
.~, 37%
,~ ,=.. I " ~ ~ Ha 47% ,

0.0 " ~ "-


9. " " " " ' ' I I I, . , , , , , . I I I I 9 " ' ' "

10~ 101 10 2 10 3
[s-q
Figure 4.3 Characteristic elastic time (~. - Nl/rlm?2) as a function of shear rate
for PVC/PB suspensions at six different concentrations (From Carreau
et al. [20].

Although the Polybutene matrix was inelastic in the measured shear rate domain,
significant primary normal stress differences, N1, w e r e measured for the PVC/PB
suspensions and increased rapidly with solids concentration. The results are
reported in Figure 4.3 in terms of a characteristic elastic time, ~. = N~/1]m:~2 , a s a
function of shear rate. For concentrations lower than 0.28, the elastic time is not
significantly different from zero. On the other hand, for higher concentrations the
elastic time is large at low shear rates and increases with concentration. However,
drops rapidly with increasing shear rate, as observed before for the specific
viscosity. Because of the nature of these suspensions, the elastic properties have to
be related to particle-particle steric interactions; the particles are large and not
affected by the Brownian motion and the matrix is purely viscous. The 47 vol %
suspension depicts the larger elastic characteristic and the stronger shear-thinning
behavior. To prove that shear thinning of such concentrated suspensions of non
colloidal particles is not due to hydrodynamic effects, we present in the next three
figures oscillatory data for the 47 vol % suspension. All experimems ,sere carried
out using a controlled stress rheometer (CSM of Bohlin) equipped with a 25 mm
1332

parallel plate geometry.


As illustrated in Figure 4.4 for a frequency of 0.628 rad/s, no viscoelastic linear
domain could be detected. At low deformation, G 'and G "decrease with increasing
strain, similarly to a viscoplastic material. The loss (viscous) modulus is much
larger than the storage (elastic) modulus, as expected for non-interactive particles
in a viscous matrix. At a strain equal to 0.25, G 'starts to increase rapidly followed
by a similar, but more modest, increase of G '; at a deformation of about 0.5. Then
at larger strain values, both moduli are shown to decrease. Obviously, for this
system no linear or close to linear domain is depicted. Similar phenomena could be
observed for concentrated suspensions made of different types of solids. Figures
4.5 and 4.6 present two more examples.
. . . . . . 9" I . . . . . . . 1 . . . . . . . 1 . . . . . . . 1 . . . . . . "1 . . . . . . . .

10 3 = 28
S-I.,

02

ro
~,.~" 10 l

10 ~ - -~
10 .4 10 .3 10 .2 10 1 10 ~ 101 102
yo

Figure 4.4 Elastic and loss moduli at 0.628 rad/s vs. strain amplitude for a 47 vol
% PVC/PB suspension (From Carreau et al. [20]).

Figures 4.5 and 4.6 report the elastic and loss moduli as functions of the strain
amplitude at ~ = 0.628 s~. The results of Figure 4.5 were obtained for 47 vol %
dispersion of coarse polypropylene (PP) particles in the same polybutene. The
average diameter of these PP particles was 300 lam and were spherical. To respect
criterion 2.1, a gap of 3.0 mm was used. The particles and the fluid had almost the
same density and measurement was not affected by settling, at least in the time
frame of the experiments. Obviously, particles of this size are not affected by any
type of colloidal forces. In this figure, a viscoelastic linear domain is observed until
the strain amplitude reaches 10%, then a large and rapid increase of G "is observed,
from a value of 13 to 160 Pa. Also G "is seen to start to increase at a larger strain
1333

103 a a I a 9 ! a l a i e 9 I i |

,,,
ff

102

101 . i I I I I I I I I I i i i

10 "1 ?o 10 ~

Figure 4.5 Elastic and loss moduli at 0.628 rad/s vs. strain amplitude for a 47 vol
% PP/PB suspension (Unpublished data [57]).

10 2 .
.
. .
.
.
.
.
. . . . .a .

G"

lO1

10 o
10-1 100
O

Figure 4.6 Elastic and loss moduli at 0.628 rad/s vs. strain amplitude for a 24 vol
% GB/PB suspension (Unpublished data [57]).

amplitude (~ 0.15) and increases from 230 to 370 Pa at a strain amplitude of 0.7,
but the maximum is not reached.
The results of Figure 4.6 were obtained for a 24 vol % of spherical glass beads
(GB) of average diameter equal to 10 gm in the same PB. Here again for the time
frame of the experiments, the effects of settling are believed to be negligible. The
general aspect of the curves remains the same. For the lower strain amplitude a
viscoelastic linear domain is observed, followed by slight decreases of the moduli
up to a deformation of about 0.7, which is the critical strain at which the structure
1334

induction phenomenon is apparent. The strain hardening observed in this figure as


well as the previous ones is believed to be caused by particle-particle interactions
due to crowding or steric effects. From our observations, the critical strain is shifted
to higher strain values for lower solids concentrations. We assume that this critical
strain scales with the distance between particles or eventually with the radial
particle concentration distribution function. The behavior presented here seems
similar to the hardening behavior observed by Gadala-Maria and Acrivos [5].
The results of G 'and G "obtained from oscillatory measurements are based on
the first harmonic of the measured strain signal and do not represent the entire
signal. However, the increases with strain observed for each suspension are not an
artifact due to instrumentation. At the critical and above the strain for which G '
reaches a minimum, the measured (output) strain was slightly skewed compared
to the imposed sinusoidal stress (input signal of the controlled stress rheometer).
During the tests, the edge of the sample was marked with a tracer and was
videotaped using a camera equipped with a macro lens. No slippage at the wall or
fracture in the sample could be detected. Figure 4.7 reports typical normalized
signals for the suspensions which are discussed in Figures 4.5 and 4.6. The strain
response for the Newtonian matrix (PB) is also presemed as a reference. The two
solid lines represents the normalized stress signal (input) and the normalized
amplitude strain response (output) respectively for the Newtonian PB, at 0.628
rad/s. The dot-dashed line represents the strain signal obtained for the 47 vol %
PP/PB suspension and the double dot-dashed line corresponds to the 24 volume %

-1
0 2 4 6 8 10
t[s]
Figure 4.7 Normalized strain and stress signal obtained at a frequency of 0.628
rad/s vs. time for a 47 vol % PP/PB, and a 24 vol % glass beads/PB
suspension. Also shown is the normalized strain output for the
Newtonian PB matrix (Unpublished data [57]).
1335

GB/PB suspensions. In the strain-hardening domain, the deformation of the signal


increases with volume concentration. For both suspensions, no deformation of the
signal could be depicted at very low deformation.
Figure 4.8 shows the influence of frequency on strain hardening for the 47 vol
% suspension of PVC particles in PB, in terms of the complex modulus as a
function of the measured strain amplitude. Strain hardening occurs approximately
at the same strain, independently of the frequency, but because of instrument
limitations it could not be observed at the largest frequency (63 rad/s). This finding
supports our assumption that this strain hardening is due to steric or crowding
effects and not controlled by stress. However, the magnitudes of the energy
dissipated and of the strain hardening depend on the frequency.
The same oscillatory shear data are re-plotted in Figure 4.9 in terms of the stress
amplitude as a function of the strain rate amplitude, ~,~ and compared to the
steady shear stress vs. shear rate data. The low frequency data reveal a very small
elastic yield stress, ao, of approximately 1 Pa. The strain-hardening effect for the
three lowest frequencies is quite visible and it is interesting to note that the stress
amplitude at low frequencies approaches the steady shear stress. At high
frequencies, the curves appear to collapse into a single curve similar to the steady
shear curve, but much lower. Hence, because of strain hardening, the resistance to
flow of concentrated suspensions is considerably greater under steady shear
conditions.

9 " " ~ 1 ~ ------i . . . . . . . . i ~ . . . . . . . i . . . . . . . . i . . . . . . . . i

104 ~" ~

~~i,t,i~,ati,,,,i.m~ 63 s'l

103
6.3 s"l
ak
V ~ . 6 3 s-~
102
V
V
101 V V ~ . 0 6 3 s-~
........ I . . . . I ...I ........ I . ." .**.".'...I ........ I .......

10 -4 10 .3 10 -2 10-1 10 ~ 101 102


0

Figure 4.8 Complex modulus vs. strain amplitude at different frequencies


(0.0628, 0.628, 6.28 and 62.8 s~) for the 47 vol % PVC/PB
suspension (From Carreau et al. [20].
1336

10 4

10 3 constant shear
and model
o 102
D

10~ Jy c o []
A
V
0 6 3 s -1
6.3 s -1
0.63 s 1
0.063 s-1
10o ....................................

10-4 l 0 "3 10 .2 10 -1 10 ~ 101 10 2


~1, (1)y~ "1]

Figure 4.9 Steady shear stress and shear stress amplitude as functions of shear
rate and strain rate amplitude for the 47 vol % PVC/PB suspension
(From Carreau et al. [20].

Creep experiments were also carried out to confirm the strain-hardening


phenomenon. The transient shear viscosity obtained in creep and the complex
viscosity (at 0.628 rad/s) are plotted as functions of the strain and the strain
amplitude respectively in Figure 4.10 for the same 47 vol % PVC/PB suspension.
Inertial effects in creep were negligible and the viscosity obtained is the imposed
stress divided by the instantaneous shear rate calculated for each experimental
point. The viscosity functions, 11 and r I*, show two different distinct pattems. The
complex viscosity, rl*, decreases initially with strain as shown before for the
complex or the elastic and loss moduli. The creep viscosity, rl, obtained at constant
shear stress slightly increases in the beginning because of a shear rate decrease with
increasing strain. The interest of Figure 4.10 is, however, to show that the strain-
hardening phenomenon is observed at about the same critical strain for the two
different experiments. This is another confirmation of our assumption that this
strain hardening is due to steric effects. This can be seen as a structural barrier,
large enough to increase the resistance to flow and to be responsible for an
effective yield stress in non colloidal suspensions. From a rheological point of
view, this limiting case is similar to a shear-thickening effect with a limiting
infinite viscosity.
1337

[] r I , t ~ = 2 5 P a
O rl, cr = 10Pa
10 3
~x rl*, 03 = 0.63 s-1

102

10 -3 10 -z 10-1 10~ 101


~/, yo

Figure 4.10 Complex viscosity and transient shear viscosity in creep as


functions of strain amplitude and strain respectively for the 47 vol %
PVC/PB suspension (From Carreau et al. [20]).

These experimental observations allow us to make the assumption that the strain
hardening effects are included in the steady rheological properties. This
contribution for non colloidal particles could be significant compared to
hydrodynamic and other forces. As shown in Figure 4.9, the steady shear viscosity
of concentrated suspensions in a Newtonian matrix can be modeled by a Maron-
Pierce equation adding the contributions for the elastic yield stress, o o, and the
supplementary strain induced stress, Ooi. The induced yield stress was found, from
oscillatory shear experiments for the 47 vol % PVC/PB suspension, to be
proportional to the frequency at a power 0.65. Hence, we have proposed the
following extension of the Maron-Pierce model:

o - - G oY , 1o1 _< o 0 - GoY ~ (4.1)

o - 9 (o ~ + s) - ~]m(1 - )-2~t, Iol > Oo (4.2)


(~m
Using o ~ = 1 Pa, D= 188 Pa.s 0.65, s = 0.65 and (l)m-- 0.74, an excellent fit is
obtained for the shear viscosity data of the 47 vol % PVC/PB suspension, as
illustrated by the solid line in Figure 4.9.

4.2 C o a t i n g c o l o r s
Many industrial processes are transient from a material point of view. Although
1338

the flow rate or deformation rate could be quite large, the total deformation
experienced by the material may remain relatively small. In general, restricting the
flow analysis of industrial processes by using time independent or steady state
viscosity could lead to a poor understanding of the phenomena envolved and
incorrect modeling. One example of such industrial applications is paper coating,
where a coating color is applied on the paper sheet and excess removed by a thin
trailing blade. The gap between the paper sheet and the blade is in the order of few
microns. As the blade is fixed and the speed of the paper sheet is in the range of
100 km/h, the calculated apparent deformation rate is of the order of 106 s~, but the
corresponding deformation or strain for the flow under the blade is less than 10.
Coating colors are concentrated colloidal suspensions. A typical formulation
consists of water, mineral pigments as kaolin and/or grinded calcium carbonate,
latex particles and/or starch used as binder and of different other additives playing
a role on the rheological properties. Brownian and other colloidal forces affect the
rheological behavior of this type of suspension. The results presented in this section
are extracted and adapted from a publication on rheological properties of coating
colors by Lavoie et al. [58]. The coating colors, used in this work, consisted of 100
parts by weight of pigments and 11 parts by weight of styrene butadiene
carboxylated latex particles of diameter equal to 0.13 lam. The pigments were a
standard #2 US kaolin and a standard delaminated US kaolin. The pH of the colors
was adjusted to 8.0. The Brookfield viscosity of each color was then adjusted to 1
Pa.s (1000 cP) at 100 rpm by adding a "high viscosity" type CMC as dry powder.
The coating colors containing the standard #2 kaolin were prepared at total solids
of 67 and 63 wt.%, and the coating colors made up of the standard delaminated
kaolin were prepared at 66, 64 and 62 wt. % solids. To emphasize the structure
induction (strain hardening) phenomenon presented in the previous section, results
obtained on two kaolin slurries at 68 and 71 wt. % and those of an industrial
coating color of unknown formation are also presented. The rheological properties
of the colors were determined using a strain-controlled Bohlin VOR and a stress-
controlled Bohlin CVO rheometer. A constant temperature of 20 ~ and a bob-and-
cup geometry (25 mm diameter and 1 mm gap) was used at all times. Steady shear
viscosity, transient viscosity (imposed shear rate) as well as strain and frequency
sweep measurements were conducted.
At low strain or low stress, the rheological behavior of a coating color
suspension is dictated by particle-particle (colloidal) interactions, it is strongly
related to the formulation. Figure 4.11 presents the complex modulus as a function
of the shear strain, for the coating color made with the delaminated US kaolin
containing 66 wt. % solids and for the industrial coating color. A linear viscoelastic
domain is observed for the complex modulus, characteristic of solid-like
1339

viscoelastic behavior. The limit of the viscoelastic domain corresponds to the yield
strain, Yo, which is related to the yield stress by Oo/Go. The rapid decrease of the
complex modulus corresponds to plastic flow behavior, and the strain hardening,
as discussed in the previous section, appears at a critical deformation.

1~ ." . . . . . . . . ! . . . . . . ii' ~ . i o ~ & ~ .......


- [] ~ 0 ~~cacr(U-~)
" [] ~ / o L S del.cdcr(0.lI-lz)

10 ! =-- 000 9

Plastic ~ '
Msmelmtic Flow
100 . Dxr~
. . . . . . . . . . . . . . . . I ! i | | . . . .

10 -3 10-2 10-! 10 ~ 10 !
yo
Figure 4.11 Complex modulus as a function of strain amplitude for an industrial
and a 66 wt. % delaminated kaolin (model) color (From Lavoie et al.
[58]).

The oscillatory data for the industrial coating color, in terms of the amplitude of
the complex stress (o ~ G*y ~ as a function of the strain rate amplitude (y~ is
presented on Figure 4.11. Also shown are the steady shear stress vs. shear rate data.
As for the 47 vol % PVC/PB suspension, strain-hardening effects appear for the

101 Model
(5" Steady Shear

10 ~
o

co
~ 10 -! [] 31.4 s -1
cyo A 6.28 s -1
-2
V 0.628 s - !
10
10 -3 10-2 10-1 10 0 101 10 2 10 3
% 7~ [s-q

Figure 4.12 Steady shear stress and shear stress amplitude as functions of shear
rate and strain rate amplitude respectively for the same industrial
coating color as in Figure 4.11 (From Lavoie et al. [58]).
1340

three frequencies, and the stress amplitude at low frequencies approaches the steady
shear stress. The resistance to flow in steady shear situation is larger than that in
high frequency small amplitude shear case because of the stain induced structure.
The induction structure effect for the 68 wt. % standard American delaminated
kaolin suspension is presented in Figure 4.13. The complex viscosity (1"1")at o
equal to 0.628 rad/s and the transient viscosity (1]) in creep are presented as
functions of the strain amplitude (yo) and shear strain (y) respectively. These
rheological properties reflect strong particle-particle interactions. No linear
viscoelastic domain is shown and at low strain the elastic contribution is 5 times
larger than the viscous one, characteristic of a solid-like structure. At a strain
amplitude of 1.2, the structure induction phenomenon from the complex viscosity
becomes apparent. The creep experiments were conducted at three different shear
stresses, one lower than the yield stress (0.5 Pa), a second just above the yield stress
(1 Pa) and finally at a stress of 5 Pa for which the slurry behaves as a fluid. These
curves confirm the different flow regimes observed in oscillation. For a shear
stress of 0.5 Pa, the slurry behaves like a solid. The stress is large enough to deform
the material but not enough to make it flow. The large increase in the viscosity is
the reflect of the structure break-down under strain, as discussed in Section 3 in
terms of the loss modulus for fumed silica suspensions (see Figures 3.8 and 3.9).
For the second experiment (at 1 Pa) a large increase for the creep viscosity is also
observed, but at larger strains, in the domain of the induced structure phenomenon.
After reaching a maximum, the viscosity goes down as the material is beginning to
flow. Finally, the experiment at 5 Pa reveals the same type of behavior, but the
strain hardening effects are attenuated because the hydrodynamic forces dominate.
Comparable behavior would be observed for different types of concentrated
suspensions, including industrial coating colors.
During the paper coating process, coating color experiences a wide range of
deformation rates, the highest being in the nip of a roll or under a blade. However,
at high deformation rates, contact times are very short and, in these conditions, the
total deformation remains very small. Hence, the low strain behavior is as much
important as the steady flow properties for applications such as blade coating. We
stress that the coating process cannot be fully characterized using only steady state
rheological parameters. The magnitude of the deformation for the transient
experiments used to generate the data of Figure 4.13 is comparable to the total
deformation experienced in coating for the flow under the blade. The rheological
behavior under the blade is essentially that of a thixotropic material, which is time
dependent.
1341

|01
o" = 0.5 Pa
W

10~
r Pa
q

lO-1 O- = 5 p a

10 -2 10 1 10 ~ 101 10 2

T,T ~
Figure 4.13 Complex viscosity (60 = 0.628 rad/s) and creep viscosity as functions
of strain amplitude and strain respectively for a 68 wt. % delaminated
clay slurry (Adapted from Lavoie et al. [58]).

5. C O N C L U D I N G R E M A R K S

In this chapter, we have illustrated and discussed various aspects related to the
rheological behavior of concentrated suspensions. Concentrated colloidal
suspensions are characterized by yield stresses, shear thinning and thixotropic
behavior, and for small amplitude shear flow, the elastic moduli could be
significantly larger than the loss moduli. Obviously, colloidal forces and particle-
particle interactions play a major role on the properties. Changes in ionic strength
of the suspending fluid could lead to drastic differences in the viscosity and the
shear-thinning characteristic of colloidal suspensions. Suspensions of non colloidal
particles in Newtonian fluids are expected to be simpler to describe. This is done
using the concept of hard spheres and the increase of viscosity can be accounted for
by using the Krieger-Dougherty, Maron-Pierce or other equivalent expressions, in
terms of a maximum packing factor. For suspensions in non-Newtonian fluids, the
behavior is similar to that of the suspending fluid, provided particle-particle
interactions make negligible contributions to the stresses. The viscosity can then
be described combining viscosity models for polymers and Maron-Pierce
expressions for the filler effects.
We have shown that many experimental problems could lead to erroneous results
or results which lead to wrong interpretations. The dispersion of solids especially
in the case of colloidal particles could result in totally different rheological
properties. For example, an apparent yield stress and shear thinning may be caused
1342

by aggregates which could be broken down by stronger shearing. Buoyant or


gravitational effects can lead to time dependent and shear-thinning properties. The
interesting results of Acrivos et al. [23] reporte d in Figure 2.6 stress how carefully
experiments have to be designed and conducted if one wishes to obtain the true
rheological properties of suspensions. Many other sources of difficulty have been
identified in Section 2 and this implies that considerably more experimental effort
is needed to elucidate the phenomena responsible for the rheological behavior of
concentrated suspensions. Badly needed are transient data which will allow us to
better characterize thixotropy, determine the key parameters which control
thixotropy, and assess constitutive relations proposed to describe their behavior.
Some of the recently proposed rheological models discussed in Section 3 appear
to be quite promising with respect to accounting for thixotropic as well as non
linear and shear-thinning effects. A most interesting feature of these models is that
they are capable of describing a smooth transition from solid-like viscoelastic
behavior to a liquid-like viscoelastic one. However, the specific forms for the
kinetic equations proposed to describe the time evolution of the structure need to
be assessed using more transient data, such as stress growth and large amplitude
oscillatory shear data. For example, in the rheological models of De Kee and Chan
Man Fong [56] and ofYziquel et al. [48], the structure is assumed to be dependent
on the second invariant of the rate-of-strain tensor. In light of our results on strain
hardening discussed in Section 4, we believe that invariants of the strain tensor
need also to be considered, at least to account for the crowding effects in highly
concentrated suspensions.
The rheological behavior of a concentrated suspension is the results of various
contributions which are very difficult to measure and accurately describe. When
dealing with industrial suspensions, other complications may arise from
uncontrollable parameters, such as purity and technical specifications of the
various components, etc... Suspensions are usually considered as homogeneous
materials (a continuum) and the solid particles are assumed to move as the
surrounding fluid (the so-called affine assumption). Close to maximum packing,
the continuum hypothesis is no longer valid as the motion of a particle is restricted
by the surrounding particles. We have shown, using many different concentrated
suspensions, that such steric effects lead to strain-hardening phenomena for strain
values of unity or less depending on the distance between surrounding particles.
This strain hardening has been observed in different deformation modes such as
oscillatory, creep or shear stress growth experiments. This is responsible for the
non linear behavior at very low strains of concentrated suspensions of so-called non
interactive particles, and apparent yield stresses, shear-thinning effects and primary
normal stresses detected for non colloidal suspensions in inelastic fluids. This
1343

strain hardening appears to be an important rheological characteristics of industrial


suspensions in high strain rate but small strain processes such as coating colors in
blade coating. The maximum stress for various coating colors in stress growth
experiments have been successfully correlated to their runnability in coating [58].
We believe that the continuum concept is no longer valid for highly concentrated
suspensions. Hence, in modeling the rheological properties of such materials, one
needs to relax the assumption of affine deformation embedded, for example, in the
model of Yziquel et al. [48]. This can be done empirically using the Gordon-
Schowalter [59] derivative with a slip parameter instead of the usual upper
convected derivative. A better and probably more successful route is to use
computer (direct) simulations of the dynamics of particles injected in a suspending
fluid. Such (2-D) computer simulations have been carried out for a few hundred
long rigid fibers [60] and the qualitative results were shown to be of interest. The
2-D simulations of large size spherical particles, for a number large enough to
account for situations close to maximum packing, should reveal realistic results
that explain the strain-hardening effects discussed in Section 4. We intend to report
such results in a near future and eventually to extend the simulations to 3-D.

REFERENCES

~ Metzner, A.B., J. Rheol., 29 (1985) 739.


2. Einstein, A., Ann. Physick, 19 (1906) 289; 34 (1911) 591.
3. Tsai, S.C., K. Ghazimorad, J. Rheol., 34 (1990) 1327.
4. Tsai, S.C., K. Zammouri, J. Rheol., 32 (1988) 737.
5. Gadala-Maria, F. and A. Acrivos, J. Rheol., 24 (1980) 799.
6. Krieger I.M. and M. Eguiluz, Trans. Soc. RheoI., 20 (1976) 29.
7. Leong, Y.K. and D.V. Boger, J. Coll. Int. Sci., 136 (1990) 249.
8. Tsai, S.C. and B.M. Viers, J. Rheol., 31 (1987) 483.
9. Mewis, J., J.W. Frith., T.A. Strivens, and W.B. Russel, AIChE J., 35 (1989)
415.
10. Russel, W.B., D.A. Saville, and W.R. Schowalter, Colloidal dispersions,
Cambridge University Press, Cambridge, MA (1989).
11. Bingham, E.C., Fluidity and Plasticity, McGraw-Hill, New York, NY (1922).
12. Casson, N, in Rheology of Dispersed Systems, edited by C.C. Mill,
Pergamon Press, New York, NY (1959).
13. Herschel, W.H. and R. Bulkey, Kolloid-Z., 34 (1926) 291.
14. Poslinski, A.J., M.E. Ryan, R.K. Gupta, S.G. Seshadri, and F.J. Frechette, J.
Rheol., 32 (1988) 703; and 751.
15. Oldroyd, J.G., Proc. Cambridge Philos. Soc., 43 (1947) 100.
1344

16. Yoshimura, A.S., and R.K. Prud'homme, Rheol. Acta, 26 (1987) 428.
17. Doraiswamy, D., A.N. Mujumbar, I. Tsao, A.N. Beris, S.C. Danford, and
A.B. Metzner, J. Rheol., 35 (1991) 647.
18. Husband, D.M, N. Askel, and W. Gleissle, J. Rheol., 37 (1993) 215.
19. De Kee, D., and C.J. Duming, in Polymer Rheology and Proceeding, edited
by A. Collyer and L.A. Utracki, Elsevier, New York, NY (1990).
20. Carreau, P.J., P.A. Lavoie, and M. Bagassi, in Macromolecular Symposia,
edited by E. Papirer and J.P. Trotignon, Huthig and Wepf Verlag, Oxford,
CT, 108 (1996) 111.
21. Otsubo, Y., J. Coll. Int. Sci., 112 (1986) 380.
22. Carreau, P.J., in Transport Processes in Bubbles, Drops and Particles, edited
by R.P. Chhabra and D. De Kee, Hemisphere Publishing Corporation,
Washington, DC (1992) 165.
23. Acrivos, A., X. Fan, and R. Mauri, J. Rheol., 38 (1994) 1285.
24. Batchelor, G.K., J. Fluid Mech., 83 (1947) 97.
25. Krieger, I.M. and I.J. Dougerthy, Trans. Soc. Rheol., 3 (1959) 137.
26. Mooney, M., J. Coll. Sci., 6 (1951) 162.
27. Maron, S.H. and P.E. Pierce, J. Coll. Sci, 11 (1956) 80.
28. Chang, C. and R.L. Powell, J. Rheol., 38 (1994) 85.
29. Storms, R.F., B.V. Ramarao, and R.H. Weiland, Powder Tech., 63 (1990)
247.
30. Shapiro, A.P. and R.F. Probstein, Phys. Rev. Lett., 68 (1992) 1422.
31. de Kruif, C.G., E.M.F. van Iersel, A. Vrij, and W.B. Russel, J. Chem. Phys.,
83 (1985) 4717.
32. Krieger, I.M., Adv. Coll. Int. Sci., 3 (1972) 111.
33. Barnes, H.A., J.F. Hutton, and K. Walters, An Introduction to Rheology,
Elsevier, New York, NY (1989).
34. Cross, M.M., J. Coll. Sci., 20 (1965) 417.
35. Mellema, J., C.G. de Kruif, C. Blom, and A. Vrij., Rheol. Acta, 26 (1987)
40.
36. Frith, W.J., J. Mewis, and T.A. Strivens, Powder Tech., 51, 27 (1987).
37. Nicodemo, L., L. Nicolais, and R.F. Landel, Chem. Eng Sci., 29 (1974) 729.
38. Mewis, J. and R. de Bleyser, Rheol. Acta, 14 (1975) 721.
39. Lobe, V.M. and J.L. White, Polym. Eng. Sci., 19 (1979) 617.
40. Tanaka, H. and J.L. White, Polym. Eng. Sci., 20 (1980) 949.
41. Barnes, H.A., J. Non-Newtonian Fluid Mech.,70 (1997) 1.
42. Quemada, D., Rheol. Acta, 17 (1978) 632.
43. Quemada, D., Rev. G6n. Therm. 279 (1985) 174.
44. De Kee, D., R.K. Code, and G. Turcotte, J. Rheol., 27 (1983) 581.
1345

45. Leonov, A.I., J. Rheol., 34 (1990) 1039.


46. Suetsugu, Y. and J.L. White, J. Appl. Polym. Sci., 28 (1983) 121.
47. Coussot, P., A.I. Leonov, and J.M. Piau, J. Non-Newtonian Fluid Mech., 46
(1993) 94.
48. Yziquel, F., P.J. Carreau, and P.J.Tanguy, submitted to J. Rheol. (1997).
49. Marrucci., G., G. Titomanlio, and G.C. Sarti, Rheol. Acta, 12 (1973) 269.
50. Hoffman, R.L., Trans. Soc. Rheol., 16 (1972) 155.
51. Boersma, W.H., J. Laven, and H.N. Stein, AIChE J., 36 (1990) 321.
52. Laun, H.M., R. Bung, S. Hess, W. Loose, O. Hess, K. Hahn, E. H~idicke, R.
Higmann, F. Schmidt, and P. Lindner, J. Rheol., 36 (1992) 743.
53. Brady, J.F. and G. Bossis, Ann. Rev. Fluid Mech., 20 (1988) 111.
54. Frith, W.J., P. d'Haene, R. Buscall, and J. Mewis, J. Rheol., 40 (1996) 531.
55. Collins, E.A., J. Coll. Sci., 71 (1979) 21.
56. De Kee, D. and C.F. Chan Man Fong, Polym. Eng. Sci., 35 (1994) 438.
57. Cotton, F., M.A.Sc. Thesis, in preparation, Ecole Polytechnique, Montr6al,
QC, Canada (1997).
58. Lavoie, P.A., P.J. Carreau, and T. Ghosh, Tappi J., in press (1997).
59. Gordon, R.J. and W.R. Schowalter, Trans. Soc. Rheol., 16 (1972) 79.
60. Thomasset, J., M. Grmela and P.J. Carreau, J. Non-Newtonian Fluid Mech.,
in press (1997).
1347

MODELLING THE FLOW OF FIBER SUSPENSIONS IN


NARROW GAPS

F. Dupret and V. Verleye ~

CESAME, Unitd de Mdcanique Appliqude, Universitd catholique de Louvain,


avenue G. Lemaftre 4-6, B-1348 Louvain-la-Neuve, Belgium
Tel "32 (0) 10 472350, E-mail "fd@mema.ucl.ac.be
a current address 9 TECHSPACE AERO, route de Liers 121, B-4041 Milmort,
Belgium

1. INTRODUCTION

A sustained industrial interest has been focusing on fiber-reinforced polymers,


which combine enhanced stiffness with all the attractive properties of thermoplas-
tic or thermoset materials [1-4]. Thin composite parts are often mass-produced
by injection or compression molding, which result in a reduced cycle time and
automatic processing. However, the thermo-mechanical properties of the end-
product are highly dependent on the fiber orientation distribution [5-7], which is
itself governed by the flow of the compound during processing [1-3,6,8,9]. In
particular, obtaining random orientation (which is convenient in most practical
cases) is almost impossible in view of the strong aligning effect of shear and ex-
tensional flows [10-13]. This is where numerical simulation can play a key role,
the ultimate aim being to control the flow during filling by designing the delivery
system (or selecting the initial load location) and introducing appropriate stiff-
eners, in order to obtain the desired fiber orientation in the final part.
The physical behavior of fiber suspensions is very complex in view of the effect
of fiber-fiber, fiber-flow and fiber-wall interactions. The orientation field is
indeed govemed by the ensemble averaged motion of the compound, but in tum
affects the rheology of the mixture, while the walls play an essential role in nar-
row gap flows. Several material and geometrical parameters must be considered
in order to understand and classify these interactions.
1348

9 The effect of the solvent rheology is certainly essential [14,15]. However, the
available fiber motion models are basically limited to Newtonian solvents.
9 The fibers are first of all characterized by their aspect ratio A r - I/d, where
l and d are the fiber length and diameter [10-12]. In addition, the fiber
thermo-mechanical properties (including the fiber flexibility [16-17]) can play a
non-negligible role. Most theories are, however, restricted to rigid fibers.
9 The volume fraction of the fibers (r directly affects the mixture rheology [18].
General suspensions are classified as dilute (~O<O(Ar -2) ), semi-dilute
(O(Ar-2)<qb<O(Ar-1)) or concentrated (O(Ar-1)<~)), depending on
whether the particles do not interact, have only hydrodynamic interactions, or
interact through both hydrodynamic and direct kinematic effects [3,19-24].
Whereas the dilute and semi-dilute theories have been extensively investigated
[12,21-30], the much more frequent case of concentrated suspensions remains
limited to phenomenological models [1-3,6,20,31-33] and a theory in its begin-
ning stages [34-37].
9 In narrow gaps of thickness 2h, fiber out-of-plane rotation is basically gov-
erned by the ratio I/h (narrow meaning that e-h/L<<l, where L is a
characteristic in-plane dimension [1-3,6-9,38-42]).
9 Other dimensionless numbers characterize the processing conditions.
It is of course totally impossible to elaborate a model encompassing all these
phenomena. Nevertheless, most of the existing theories are based on the use of
orientation tensors [26b,43a] to represent the fiber orientation field and to model
the interaction of the fibers with themselves, the solvent or the walls. A common
presentation of the basic physical model for various kinds of suspensions and pro-
cesses will, therefore, be possible. As the use of orientation tensors requires to
solve the closure problem (i.e., to relate the (n+2)th-order orientation tensor to the
nth-order tensor by means of a sound physical approximation) [26b,40-46], this
issue will be carefully investigated. Another limitation, however, concerns the
numerical solution of the general orientation-flow coupled problem, which entails
significant difficulties [47,48]. In this respect, the recent development of coupled
lubrication models adapted to narrow gap flows [49,50], and of general two-di-
mensional (2D) coupled models with viscoelastic rheology [51] and three-dimen-
sional (3D) coupled models [52], should be emphasized. Hence, important pro-
gress can be expected in the near future concerning the orientation and flow
models that will be available for industrial use.
The objective of the present chapter is to review the state of the art in fiber ori-
entation modelling, with a particular attention bearing on the decoupled lubri-
cation model [2]. The algorithm elaborated by Dupret and co-workers [4,9,39-
1349

42,53-56] in order to predict the flow-induced orientation field in short or long


fiber injection or compression molded suspensions will be analyzed. An impor-
tam part of this work was reviewed by Crochet et al. [9]; subsequent develop-
ments were basically devoted to developing and validating the natural closure
approximation [4,40-42,56,57] and to using the simulation results to predict the
thermo-mechanical properties of the molded part [5-7]. The flow numerical mod-
el will only be summarized, since it is detailed in the companion chapter "Indus-
trial Flows" Polymer Processing and Rheology" Modelling and Simulation of
Injection Molding", that will be referred to as "Chapter IM" in the sequel.
The plan of the chapter is as follows : Section 2 is devoted to the general phys-
ics of fiber suspensions, considering the effect of the flow on the orientation field,
and the fiber-fiber and fiber-flow interactions; the entire Section 3 focuses on the
closure approximation problem; Section 4 investigates the particular theory and
numerical modelling of fiber suspension flows in narrow gaps; finally, some simu-
lation results and experimental validations are presented in Section 5.

2. THE PHYSICS OF F L O W I N G SUSPENSIONS

2.1 The motion of a single fiber in a liquid


The starting point of fiber orientation modelling was the work of Jeffery [10],
who analyzed the motion of a rigid infinitely small particle in an incompressible
Newtonian solvent, with no external forces or torques nor Brownian motion.
From these hypotheses, it results that the flow around the particle is of a creeping
type and that, at the particle scale, the asymptotic velocity gradient at infinity is
uniform at all time. In addition, as the particle (or fiber) was assumed to be an
axisymmetric ellipsoid (Figure 1), Jeffery could prove that any vector Pi of
constant length and parallel to the fiber axis satisfies the evolution equation
[]

Pi - - ~" dkl PiPkPl (PmPm) -1 , (1)


[]

where ~, = (ArL 1)/(Ar 2+ 1) is the fiber shape factor and Pi denotes a mixed
convected derivative of Pi"

a OPi 1 (1 + ~) ~ 1 ~j (2)
Pi = Dt 2 ~ j j P J + -2 ( 1 - A ) -~ii P J "

The Einstein summation convention over repeated indices is used and t , x i and
Vi stand for time, Cartesian coordinates and velocity components, while the
1350

l
e1

Figure 1. Sketch of an ellipsoidal fiber and unit vector representing the fiber
orientation.

expressions DPilDt-OPilOt+vjOPilOx2 and d i j - ~ 2 ( & i l & j +&jl&i)


denote the material derivative of Pi and the rate of strain tensor, respectively.
Equation (1) deserves a short comment. In a general flow, the evolution equa-
tion of an infinitesimal material segment d x i is

dxi - --(dxi ) - dxj -- 0 , (3)


Dt
while an infinitesimal material surface dS normal to the unit direction ni
obeys, in incompressible flow, the equation

dS i - ( dSi ) + dS j - 0 , (4)

with dS i - n i d S . The unit vectors Pi parallel to dx i or dS i (with


- dx,/lldxll or da,/lldSll) satisfy (1) with ~ - 1 or - 1, respectively, corre-
sponding to infinitely slender or flat particles. Hence, while the left-hand side of
(1) is a mixture between upper and lower convected derivatives, due to the effect
of fiber aspect ratio, the right-hand side arises from using normalized vectors Pi
(instead of deforming vectors such as dxi o r d S i ) t o represent the fiber orienta-
tion. Observe also from (1) that % t (PiPi)-0. The presence of a normaliza-
tion term in the right-hand side of (1) causes significant modelling difficulties via
the induced non-linearities.
The Jeffery equation has been extensively studied (cf. [8,58-60]). Bretherton
[11] developed the general solution in simple shear flow, and proved that the
motion of the axis of most rigid bodies of revolution is identical with that of some
ellipsoid. The creeping motion of infinitely small particles in a Newtonian sol-
vent was further investigated by Batchelor [19,61], who developed the slender
1351

body theory, and Brenner (cf. the references in [12]), who highlighted the intrin-
sic tensors governing the motion of arbitrary particles under external forces and
torques. Vincent and Agassant [13] validated the Jeffery model by means of
extensional flow experiments, while Givler et al. [62, 63] and Henry de Frahan et
al. [39] developed numerical tools to solve (1) for plane and multi-faceted thin
shapes. From these investigations, it is worth recalling that, in extensional flow,
the fibers quickly align with the principal stretching direction. In shear flow, each
fiber rotates along a given "Jeffery orbit" at a non-uniform rate with a period
T-2zr (Ar + Ar-1)/y (where ~ ' , a s defined from )?2 2dijdij, denotes the
shear rate); for slender fibers (Ar >> 1), rapid tumbling motions separate long
periods of nearly in-plane orientation; for inf'mitely slender fibers ( Ar = oo ), each
in-plane orientation is an unstable equilibrium position and the motion is no
longer periodic (see also the stability analyses of MacMillan [64] and Szeri et al.
[60], and the investigation of the 3D transition from in-plane to uni-directional
orientation in shear flow in [9]).
The Jeffery model remains useful to evaluate closure approximations or other
effects by averaging the results obtained with a large set of fibers [65-67]. Ingber
and Mondy [59] have shown from exact flow calculations that Jeffery's theory is
robust for rigid fibers, even when weak hydrodynamic interactions with other
fibers or the walls are considered. Similar conclusions can be drawn for semi-
dilute suspensions from the analysis of Koch and Shaqfeh [23] and the experi-
ments of Stover et al. [68]. The theory should be further elaborated in order to
take the solvent viscoelasticity into account, since experiments indicate that vari-
ous behaviors exist, which often disagree with theoretical predictions (cf. Feng
and Joseph [14], Iso et al. [15] and the literature therein).

2.2 Distribution functions; fiber-fiber interaction


The Jeffery model is unable to take into account fiber-fiber and fiber-flow inter-
actions (this being necessary with semi-dilute or concentrated suspensions).
Moreover, orientation must be treated as a random phenomenon, since fibers are
never perfectly aligned in the initial state or in the inlet flow sections. Distribu-
tion functions must therefore be introduced [20-22,27,30,58,68-72]. The distribu-
tion function ~?(p, x,t) is the orientation probability density as a function of p ,
x and t. Hence, dP - ~?(p,x,t)dp is the elementary probability of finding a
fiber whose orientation is located in the infinitesimal zone dp surrounding p.
Note that p is a unit vector, and dp is an element of area of the unit sphere
U . In the 2D case (which applies in the molding of concentrated suspensions of
1352

long fibers), dp is an element of length of the unit circle (which will also be
denoted by U ). The distribution function is normalized by the condition
~u ~ dp - 1, while the symmetry condition W(p) - q ' ( - p ) is introduced to
remove the indeterminacy between fibers oriented in the directions p and - p .
The evolution equation of W has the form [1,73] :

o%t - ,
(5)

where O~/03Pi denotes the constrained divergence operator on U , while Pi


represents the material derivative of Pi expressed from a constitutive equation
as a function of p, Vv and W. The operator tg/Opi is constrained because
(Pl,P2,P3) are not independent variables since Ilpll- . is no problem
because Pi is itself constrained to be orthogonal to Pi when lid- 1, and it is
sufficient to replace everywhere tg/o3Pi by (t)/o3Pi- PiPj t~/Opj) to obtain
non-constrained equations.
The constitutive equation proposed by Folgar and Tucker [20] was based on
adding a rotary diffusion term to the material derivative Dpi/Dt provided by
Jeffery's equation (1), in order to represent the geometrical and random effect of
fiber-fiber interaction :
1 o~vi 1 ~j ~ln~I/
Pi -- -~(1 + / ~ ) ~ j pj ---~(1--/~)~/pj -/], dkt PiPkPt - C 1 9 7 ~
oqPi (6)
where & is the equivalent fiber shape factor of the fiber (taking its real shape
into account) and C I is the (phenomenological) interaction coefficient. The last
term of (6) tends to randomize the fiber orientation by flattening the distribution
function. Due to the factor ~ , the final fiber orientation does not depend on the
global magnitude of the velocity field, but only on its spatial distribution. More
precisely, when velocities are multiplied by an arbitrary uniform positive scale
factor F(t), the final orientation field is not changed provided the time scale is
adapted, because all the terms on the right-hand side of (6) are multiplied by
F (t). On the other hand, reverse motion does not bring back the initial orienta-
tion (as the Jeffery model does), because ~' does not change sign with the velo-
city direction. All this is in agreement with the physics of fiber-fiber interactions
of purely hydrodynamic or kinematic origin. On the contrary, in the case of
Brownian interactions, D t =Ct~, (and not C t ) is a material function of tempe-
rature, volume fraction and orientation distribution [58,69,70]. It should be noted
1353

that Hinch and Leal [69a] def'me the Brownian effect as strong, weak, or very
weak, according to whether DI/~/>> 1, 1 >> DI/~z >> (Ar + Ar -1)-3, or

(Ar + Ar -])-3 >> D , / y . The same scaling could thus be considered for C I .

3.01 0 Experimentali (a)


. Approximated Fit
2.5

=- C I =0 . 0 0 4 4 _
.9 2.0
Aspect ratio
=83
g ].5
Volume fraction
o'1 =0.013
?:51.o

O.5 (3

0
-rr/2 0 77"/2
Orientation Angle, c~

3.0
' ( y : 1.56)
0 Data Points (b) '
0 Data Points (7': 1.56)
(c)
- - Approximated Fit . . . . Initial Orientation (Exp.)
2.5 2.5 Approximated Fit

g~o ~'2.0

t.l. LI.

~0
1.5 C
.91,5
e~ C I: 0.004 .o
-~.
=-
ul

~1,0 -.,,
i:51.0 - CI: aO15 -

0.5
0 0.5 - 1~~
o_ D o o o
0 i 0 I
-~'/2 0 71"12 -71"12 0 ~'/2
Orientation Angle, ~) Orientation Angle,

Figure 2. Experimental validation of the distribution function in shear flow : (a)


steady distribution and best fitted theoretical result; (b) transient data and theoret-
ical predictions with the same parameters as in case (a); (c) same situation as in
case (b), with C / - 0.015. (From [20]).
1354

The most significant consequence of (6) is that, in any steady homogeneous


flow, an asymptotic W exists that can easily be determined by letting DW/Dt
vanish in (5). According to Folgar and Tucker [20], this steady state is more
aligned in elongational than in shear flow (in agreement with experimental obser-
vations in molding processes). Figure 2(a) shows a typical comparison between
experimental orientation measurements and best fitted theoretical predictions in
steady shear flow. Although the agreement is excellent, further comparisons in
transient shear flow with the same value of CI (Figures 2(b) and 2(c)) indicate
that interaction is more important when the fibers are in a random state than when
alignment is high. It should be emphasized that the Folgar-Tucker model of fiber-
fiber interaction is phenomenological and does not rest on a kinetic theory. Ex-
perimental measurements of C/ are described in [1,20,74] (see also [8]), while
the numerical simulation of the motion of many simplified rod-like particles by
Yamane et al. [66] provides estimates of CI whose order of magnitude agrees
with the experimental values. A correlation between CI and the average inter-
fiber spacing in concentrated suspensions is proposed in [34,35]. However, the
only comprehensive theory available concerns semi-dilute suspensions [22-24,28-
30,68,71,72,75]. Very briefly, the principal concept of this theory is the screen-
ing length (Batchelor [ 19]), which scales the short-range screening of hydrody-
namic fiber-fiber interactions and is of the order o[d/2 (r 111~-1)1/2] (it is thus
close to the average inter-fiber spacing) [22,23,28,75]. The interaction model of
Fredrickson and Shaqfeh [28] and Shaqfeh and Koch [22,23] is based on using
multiple scattering expansions of the average Green's function in the Laplace-
Fourier space. Orientational and translational di~sivities, and drift velocities
can be described by this approach [22,71,72], which was validated by the exper-
iments of Stover et al. [68] and the multi-particle numerical simulations of Claeys
and Brady [76] and Mackaplow et al. [29,30] (note that both the dilute/semi-
dilute transition and the semi-dilute regime are very well predicted by a dilute
model with two-body interactions [28b,29,30,75]). In spite of these significant
developments, the resulting models are extremely complex and do not reach the
concentrated regime (except in recent work by Sundararajakumar and Koch [37]).
They thus remain basically useless for the practical simulation of molding
processes. Further research is therefore needed.
More complex constitutive laws than (6) can be imagined, since interaction
depends on the rate of strain and is different when the fibers are aligned and when
orientation is random. Anisotropic effects could be considered by replacing in (6)
the orientation rotary diffusion flux Cti --CtYc)W/OPi by an expression of
the form ~Oli - - K i j ~OtlJ/6~Pj, with a general positive definite interaction ten-
1355

sor Kij. The latter can be developed as a sum involving the 9 base syrrmaetric
tensors constructed from pi and dii (with coefficients depending on their
invariants and W ), and must respect invariance of fiber orientation with respect
to time scale changes. Hence, the general theory is very complicated.

FiA

Figure 3. Conservation of the fibers from isotropic to actual configuration.

A particular class of distribution functions introduced by Dinh and Armstrong


[21] exhibits properties of key importance for the closure problem. This class is
obtained by solving the evolution equation (5) with infinitely slender fibers
(~ = 1) and a vanishing CI, starting from the isotropic orientation state W0,

~o - / J r for 3D orientation, ~o - l~2jr for 2D orientation. (7)

A simple analysis shows that Wc is governed by the deformation gradient Fia


from initial to actual state (FiA - Ox i/OxOA, with capital letters indicating the ref-
erence configuration). Indeed, the relation between initial and final orientations
(pO and Pi ) is directly obtained by solving (1) for there is no fiber-fiber interac-
tion. Moreover, infinitely slender fibers obey the same equation as normalized
infinitesimal vectors, and thus, from the relation dx i - giA dx 0 , it is easy to show

that Pi-FiA pO i..o]l-1 and p~ Pi ]-l.pl1-1. On the other hand, as


C t = 0, the distribution function evolution equation can be integrated with re-
spect to time (Figure 3) :
Wo dpo = Wc (p,x,t) dp , (8)
where qJ0 is given by (7), while dpo and dp denote associated infmdtesimal
elements of area (or length) of U . Therefore, dPo/dp is the Jacobian of pO
expressed in terms of pi. After some calculations, this provides from (7) the
distribution functions one is looking for :
1356

{WC ( P) - l~4zr(Bij-lpi Pj )-3/2


q*c (P) - /l~2jr (Bij-lpi Pj ) -1
in 3 D orientation,
in 2 D orientation,
(9)

where B 6 - Fia FjA is the Finger strain tensor from initial to actual state.
It is clear that W c has 5 degrees of freedom in the 3D case and 2 in the 2D
case, since Bij is symmetric and its determinant is 1 due to the suspension in-
compressibility. When the coefficients Bij are arbitrary numbers constrained by
these requirements, the distributions Wc(p) form a class called by Verleye et
al. [40-42,56] the class of canonical distribution functions, since they considered
that it is the simplest class of distribution functions and used it to build the natur-
al closure approximation by non-linear projection. This class was also used by
Szeri and Lin [46] to approximate the solution of the fiber orientation problem
(with Brownian motion) on the basis of the double-Lagrangian technique of Szeri
and Leal [44]. Lipscomb et al. [27] proved that the same class is generated by
the motion of finite aspect ratio fibers without fiber-fiber interaction, since an
equivalent strain tensor Fi,~q can be defmed from (2) by the relation

Dt = + (~ - 1)dij Ffq ] " (10)

Both 2D and 3D analytical solutions for W without fiber-fiber interaction were


investigated by Altan et al. [73]. See also numerical solutions in [77].

2.3 Orientation tensors


Although W provides a complete description of the orientation state in the sus-
pension, its use is cumbersome since W ( p , x , t ) has 6 independent variables in
the 3D case (and 4 in the 2D case). This is why Hinch and Leal [26b], Lipscomb
et al. [47] and Advani and Tucker [43a], followed by many other authors [2,4,9,
27,39-41,43b,45,48-50,53,65,73,78-88], developed the use of orientation tensors
(which are successive and more and more detailed summaries of W). The n th-
order orientation tensor a n is obtained by averaging the n th dyadic power p|
of p :

an - ~W(p,x,t)p| . (11)
u
From the normalization and synnnetry properties of W, it is clear that a 0 is the
constant 1 and that all odd-order orientation tensors vanish. We will restrict
1357

ourselves to the 2 nd-, 4 th- and 6 th- order tensors aij, aijkl and aijklmn, with

aij - ~W piPjdp , aijkl -- ~W piPjpkPldP , aijklmn = .-.. (12)


u u

The basic properties of orientation tensors are full symmetry and normalization
[43a]. It is convenient to introduce the symmetrization operator S as follows :
for a genetic tensor b, S(b) is the average of all the transposes of b. Hence,
S (bij) and S (bijkl) are defined by

S ( b ij ) - /~2 ( b ij + a f t ) , (13)

S(b(ikl) = /~24(bijkl +bijlk +bikjl +bjikl +...) (24terms), (14)

with a similar expression (with 720 terms) for S(bijklmn ) . The full symmetry
property, which is a direct consequence of (12), writes as :

aij - S ( a i j ) , aijkl - S ( a i j k l ) , aijklmn = S(aijklmn ) . (15)

On the other hand, the normalization condition results from the fact that Pi is a
unit vector. Hence, from (12),

aii - 1, aijkk -- aij , aijklmm -- aijkl 9 (16)

This shows that all the informations provided by aij are contained in aiju , and
the same for aijkt and aijklmn. Nevertheless, the deviatoric parts of the orienta-
tion tensors can be defined by the expressions [43a]

I,ta ij - a ij
_iS. .
N tj , (17)
3
d -- 6------~-S ) "k- S(~ij(~kl),
aijkl - aijkt 4 + N (Sijakt (4 + N)(2 + N)

and so on with N = 3 or 2 depending on the problem dimension, in such a way


that their traces vanish :
d d d
aii - O , aijkk -- O , aijklmm -- 0 . (18)
d d d
The tensors aij, aijkt, aijklmn ... are independent and W can be recovered
from the expansion [89]
d )d d )d
W(p) - A o + A1 a(i ( P i P j + A 2 aijkl(piPjPkP l + ... , (I9)
1358

where ( p i p j ) cl , ( p i P j P k P t ) d ... are the deviatoric parts of P i P j , PiPjPkPt ...


d d
(defined in the same way as aij, aijkl ... in (17)), while A0, A1, A2 ... are
coefficients which depend on N (see [43a]). It should be pointed out that the ap-
proximations obtained by limiting the number of terms in (19) are not very accu-
rate since the expansion is linear in the powers of Pi (see [90]).
Once the orientation tensors are defined, it is very easy to establish their evolu-
tion equations by orientation averaging. According to Advarli and Tucker [43a],
this is performed by multiplying (6) by the successive odd dyadic powers of Pi
and integrating the result over U . After some calculations involving careful inte-
grations by parts since c)/Op i is constrained, this provides the equations

ij - -2~aijktd~a - 2 C 1 ~ ' N a q ,
( d 3 N ) (20)
[aokt - - 4 X aokt,,, , dmn - 4 C r y ( N + 2)a01 a + ---S(a~6kl
+4)N "

I-1 rl

The mixed convected derivatives a ij and aokt are defined in the same way as
rl

P i in equation (2)"
[] V A

v a (21)
ijkl -- ~2(1 + Z) aijkt+/~2 (1 - X)aiykt,
V V A A
where a ij , a ijkt, a q and a ijkt denote the upper- and lower-convected deriva-
tives of aq and aqk l . In particular,

[] Daij l(l+,~)(O~ia Ova) 1 (~. oars)


aij = Dt 2 ~.Ox k kj + aik + -~(1 - ~) -'--t akj + aik . (22)

The 1~t and 2 "a equations (20) are embedded, in the sense that the trace of the
2 nd equation is exactly the 1st equation, whose trace is itself vanishing. Hence,
only one equation need be considered in a given model. Most authors select aij
as the basic variable (with 5 unknowns in 3D problems) [2,9,26b,27,39-41,43,45-
49,53,78,79,81 ], while Altan et al. [65,80,82,83,86,87] investigated the 2 "a option
(i.e., to solve for aijkt, with 14 unknowns in the 3D case), in view of the direct
effect of aijkt on the thermo-mechanical properties of the product [7].
1359

Solving the problem in terms of orientation tensors has the drawback that the ev-
olution equation of the 2nO-order orientation tensor involves the 4th-order tensor,
whose equation itself involves the 6th-order tensor, etc. To circumvent this diffi-
culty, a closure approximation expressing aijkl algebraically in terms of aij (or
aijklmn in terms of aijkt ) is required. In view of its importance, this issue is ex-
amined in a separate section.

2.4 Rheological aspects; fiber-flow interaction


Detailed analysis of the theology of fiber suspensions is not the object of this
chapter, since basically orientation-flow decoupled simulation models are consid-
ered. Nevertheless, the relevant literature will be reviewed in view of its impor-
tance for future developments. Following Prager [91], Batchelor [25], Brenner
[12], Hinch and Leal [26b], Lipscomb et al. [27] and Ausias et al [92], the extra-
stress zij can be expressed for a Newtonian solvent in the generic form [2] 9

~'ij -- 21"ld ij + 2 Tl ~) [ A a ijkl d kl + B ( d ik a kj + a ik d kj ) + C d ij + 2 O a ij ] , (23)

where 7"/ is the solvent viscosity, while A, B, C and D are coefficients which
primarily depend upon the suspension concentration and the aspect ratio and
degree of alignment of the fibers, but might also be functions of the invariants of
ai# l and dij [93]. The dilute theory is well-established [10,12,25,27,58,61,69,
94] while, starting from the pioneer work of Batchelor [19] and Dinh and Ann-
strong [21], the semi-dilute theory has been developed by Shaqfeh, Koch, Fred-
rickson and co-workers [28-30,37,75], and validated by means of experiments
[68,92,95] and numerical simulations [29,30,37]. Mass, momentum and heat
transfer have been investigated. It is worth noting that, when A is put in the
form A - 1 / 2 ar 2 [ln(X/d)] -1 , the length X (which is thereby defined) scales
as the fiber length l for dilute suspensions, and as the screening length for semi-
dilute suspensions. Hence, as explained in Section 2.2, X is in this latter case
of the order of magnitude of the average inter-fiber spacing (and not the average
shortest distance between two fibers as in [21]), for any orientation distribution.
The coefficient B is negligible for slender fibers [27], while C = 2 and D
vanishes in the absence of Brownian motion.
No complete theory exists for concentrated suspensions. Semi-empirical values
of A, B, C and D are proposed in [96]. Phenomenological models combining
extensional and shear viscosities were developed by Gibson and co-workers [6,
31-33] (see also [97]), while Toll and M~nson [36] and Sundararajakumar and
1360

Koch [37] investigated the effect of non-hydrodynamic mechanical fiber-fiber


contacts. Dynamic simulations of suspensions of rigid or flexible fibers were per-
formed by Yamamoto et al. [17b] by modelling each fiber as an array of spheres.
Another important issue concerns the coupled numerical solution of the orienta-
tion and flow problems, which was investigated in planar or axisymmetric 2D sit-
uations by Lipscomb et al. [47], Papanastasiou and Alexandrou [77], Rosenberg
et al. [48], Ranganathan and Advani [35], and Azaiez et al. [51] (see also [82,83,
86]). The complex numerical effect of the wall no-slip condition, which causes
periodic fiber tumbling and generates very sharp orientation boundary layers that
cannot be resolved by mesh refinement, is discussed in [47,48]. Recent work by
Kabanemi et al. [52] indicates that coupling the flow and orientation predictions
becomes possible in 3D free-boundary flows by using the VOF method. The
coupled models adapted to narrow gap flows are reviewed in Section 4.
Finally, it should be noted that several authors are currently investigating the
combination of viscoelastic and fiber-induced stresses, without however taking
into account the local effect of viscoelasticity on the fiber motion (cf. [51,98]).

3. THE CLOSURE PROBLEM

The use of orientation tensors to simulate fiber suspension flows represems the
best modelling methodology as long as an appropriate closure approximation is
developed (Hand [93], Hinch and Leal [26b], Advani and Tucker [43]). This
issue brings about the question of whether the (2n+2)th-order orientation tensor
can be expressed as a given function of the (2n)th-order tensor in such a way that
the predicted values of the 2nth-order tensor obtained by solving the associated
momem equations provide good results "in most circumstances of practical in-
terest" This question expresses the situation well. Exactly in the same way that
orientation tensors are summaries of the distribution function, the moment equa-
tions completed by a closure approximation represent summaries of the distribu-
tion function evolution equation, and summaries never provide full information.
Hence, the closure problem is not precisely posed in a mathematical sense, be-
cause this would necessitate to define the probability of the different states of the
system (each being represemed by a given distribution function), which is impos-
sible for there is a lack of sufficiem knowledge about the problem physics and
also because of the tremendous difficulty of the mathematical problem. There-
fore, heuristic methods, which combine theoretical and physical considerations
with numerical and real experimems, remain the only way to tackle the question.
1361

3.1 Basic concepts


Some basic theory is necessary. Details can be found in [43]. For the sake of
simplicity, only the 4th-order tensor closure will be investigated. Both the 2D and
3D problems will be addressed in view of their technological importance.
According to the Cayley-Hamilton theorem, any 3D closure approximation can
be expressed as a sum of products of the 27 base tensors ((~ij(~kt, ~ik~jt ,...
t$ijakt .... aijakl .... t$ija~amt , ... a i m a m j a ~ a n t , ...) by scalar functions of the invar-
iants of aij,
-1
P - det (aij) , D - P aii . (24)

In the 2D case, the sum is limited to 12 terms and a single invariant needs to be
considered:
O - det (aij) . (25)

Note that symbols are selected in such a way that equations (24) and (25) match
when a 3D orientation state tends to becoming 2D.
Optimally, closure approximations should satisfy the f u l l normalization and sym-
metry conditions (16.2) and (15.2). However, this is not a constraint [26b,43b]
and it is sufficient to require p a r t i a l normalization and symmetry :

aiikl -- 1///N (aiijj - 1)t$kl -- akl , (26)

aijkl - ajikl -- aklij , (27)


since only the product aijkl dkl plays a role in the moment equations (20), while
partial symmetry guarantees that aijkl has the symmetries of an elasticity tensor
and can be used for thermo-mechanical property predictions [7]. To achieve par-
tial symmetry, the number of base tensors in the expression of Ctijkl reduces to
12 in the 3D case, and 6 in the 2D case, while full symmetry is obtained when
aijkl is expressed in the form

a ijkl -- [~1 S ( ~ij ~kl ) + [~2 S ( ~ij a kl ) + ff~3 S ( a ij a kt ) +


+ r4 S(t$ijakmaml) + r5 S(aijalonaml) + r6 S(aimamjaknanl) , (28)

or aijkl -- 131S(tSijgkl)+ 132 S(~ijakl ) + 133 S(aijakl ) , (29)

depending on whether the 3D or 2D case is considered. The coefficients fll-6


or fll-3 are functions of the invariants P and D , or D , respectively.
1362

Normalization imposes the conditions

I101!1I1 I12o 4r
0
0
7
0
132 = -
133
0
7
1-6D
5
4(P-D)
2(3- 4D)_]
]Iil + , (30)

or I80 [-4D
depending on the problem dimension.
The first closure approximations that were successively proposed by various re-
search groups are listed below. All these formulae, which are discussed in [43b],
satisfy full normalization.
9 The linear closure (Hand [93]) writes as
3 6
aii~t = (4 + N)(2 + N) S (6iy6 ~ ) + 4 + N S (aij6~t ) , (32)

in such a way that a~.kt- 0. This fully symmetric closure approximation is


exact when the fibers are in a random state. It can lead to unacceptable physi-
cal inconsistencies both in shear and elongational flows [4,40,43b].
9 Two composite closure approximations (named HL1 and HL2 in [43b]) were
developed by Hinch and Leal [26], in order to fit properly both in the strong
and weak Brownian motion limits. These closures behave well in some cases
(especially HL1), but have strong deficiencies (with possible pathological be-
havior) in other cases [43b,44b,45,99].
9 The quadratic closure aijkt- aijakl was proposed by Hinch and Leal [26b],
Doi [70] and Lipscomb et al [27,47], and has only partial symmetry. It is exact
when the fibers are perfectly aligned, but can provide poor (and even non-
physical) results in transient flows or with a non-vanishing C t [4,40,43b,44b,
73,84].
9 The hybrid closure approximation was further developed by Advani and
Tucker [43] in order to remove the problems associated with the linear and
quadratic closures. In its best formulation [43b], it writes as

aijkl -- f , I 6 1
3 + N) S(t~ijt~kl)+ 4 + N S(aijt~kl ) + ( 1 - f ' )
(4 + N)(2 aijakl 1 (33)

where f ' - N N det (aij) . The hybrid closure is exact when the fibers are in a
random state or perfectly aligned, and never behaves badly (except that it
1363

accelerates the orientation transients) [4,43b,73]. Moreover, it can be very


easily implemented.
9 Quadratic and hybrid closure approximations for aijklmn in terms of aij~l are
investigated in [35,65,67,80,82,83,86,87].
In spite of these results, several groups pursued their investigations in parallel
since the problem had not received a really satisfactory solution. In the sequel,
we will focus on the natural closure approximation [4,40-42,57], while the prin-
cipal aspects of the best fitted orthotropic closure approximation [45] and the
double-Lagrangian technique [44,46] will be summarized.
,r/ 1
I --,,--- SimpleShear .-. BiaxialElong.
-E?r- Shear/StretchA o L UniaxialElong.
0.8
- O - Shear/StretchB

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1

Figure 4. Left : space of possible orientation states in principal values and fitting
domain. Right: points generated by calculation of the distribution function and
used for the orthotropic fitted closure. (From [45]).

3 . 2 0 r t h o t r o p i c closures
The term "orthotropic" closure was introduced by Cintra and Tucker [45] to de-
note a 3D closure approximation of aijkl that has the same principal axes as aij.
Since aiykl must take all its directional information from aij, any closure ap-
proximation must be orthotropic. Hence, the closure relation expressing aijkt in
terms of aij can be considerably simplified. Letting ~, 7/ and ~" denote the
principal values of aij, it is easy to show by s ~ e t r y considerations that, for any
orthotropic closure satisfying the normalization and partial symmetry conditions
(16.2) and (27), there are, in principal axes, exactly 6 independent components
aij~l. This number reduces to 3 when full symmetry (equation (15.2)) is required.
Cintra and Tucker [45] selected allll, (/2222 and a3333 to provide these basic
1364

1.0 . . . .
all
/ [ an
/ ,--...-.- DFC ........... "~

0.8 rl._ - - ORF ..-"......... !

o.~ .... .<, i~I i


I-/ \\'.
'" " ';i --- ,,, I
L .... "'"*"" ..~/. ~.~
................

0.2

............. ~"" o.o L~. . . . .


0 5 10 15 20 25 30 o 5 lO 15 20 25 30

Gt = reduced time Gt = reduced time


0.3
a12
az3 ----- DFC
...... , DFC
ORF

/ ".. oaJ= .................................\


0.2 t " " ~.'K NAT

i "',,... ~, ! " ____" . . . . .L1


,~" ........... ------ HL1
i . . . . HL2

~x.~: "- .... H,~ 0.1 '.~~\ ................HY8


-,~,.., ............... . , ,
X. ~%.
0.0
\./

. . . . i . . . . 9 . . . . ! . . . . ! . . . . ! . .
' . . . . ' . . . . * . . . . . 0.1 . . . . . ' . . . . ' . . . . . * . . . . ' . . . . * . . . .
5 10 15 20 25 30 0 5 10 15 20 25 30
Gt = reduced time Gt = reduced time
Figure 5. Tensor components obtained by means of the best fitted orthotropic
closure. The flow starts with simple shear, switches to shearing/stretching flow,
and finally shearing continues with a strong stretching in the 3 rd direction (DFC"
integration of distribution function; ORF" orthotropic fitted closure; NAT" nat-
ural closure approximation, 1 st version; HL1 and HL2" Hinch and Leal's 1 st and
2 no composite closures; HYB 9 hybrid closure). (From [45]).

expressions. As ~, 7/ and ~" are linked by the condition (16.1),


~+r/+~"-1 (O<~,q,~'<l) , (34)
two independent variables (~ and O ) can be chosen and, due to symmetry, it is
sufficient to consider the domain 1 > ~ > 7/> ~" > 0 (Figure 4). Hence, 3 func-
tions (a1111, a2222 and a3333 ) m u s t be evaluated in this region.
In order to obtain a fitted orthotropic closure, the 1st step consists in selecting an
appropriate finite dimensional space for the functions a~l~l, a2222 and a3333.
1365

Quadratic polynomials in ~ and 7/ were selected in [45]. The 2 nd step consists


in choosing a set of flows that generate a wide variety of orientation states. For
that purpose, the simple shear, uniaxial and biaxial elongational flows, together
with two sheafing/stretching flows were selected in [45] (Figure 4). In the 3 rd
step, the polynomial coefficients of a l l l l , a2222 and a3333 are determined by
least square best fit, with the exact values obtained by integrating the evolution
equation (5). Impressive results were obtained by this technique (Figure 5), ex-
cept for very low values of the interaction coefficient C I .
In conclusion, the best fitted orthotropic closure approximations can easily be
tuned to general or particular classes of flows as long as the fitting objectives are
clear. Indeed, there is no unique solution to the closure problem and, if a general
approximation is desired, there is a limit accuracy that cannot be improved on.
Nevertheless, the formula proposed by Cintra and Tucker [45] today certainly
represents the best choice in solving a very broad class of problems in view of its
accuracy and ease of implementation. Further investigations could bear on select-
ing an optimal approximation space for allll, a2222 and a3333. A related
question is to know whether constraints should be introduced in order to avoid
discontinuities and derivative discontinuities when the solution is extended to the
whole triangular domain of orientation states (Figure 4).

3.3 Natural closures


The natural closure approximation of Verleye et al. [4,40-42,56,57] is defined
by considering the set of canonical distribution functions introduced in Section
2.2. As this class has 5 degrees of freedom in the 3D case and 2 degrees of free-
dom in the 2D case, while the positive semi-definite 2nO-order orientation tensor
itself has 5 or 2 degrees of freedom, respectively, since it is synunetric and its
trace is l, there is a one-to-one correspondence between the set of 2nd-order ori-
entation tensors and the subset of canonical distribution functions. The natural
closure approximation is generated by non-linear projection onto this subset.
More precisely, the relation (12.1), which links aij to ~ , can be inverted and
thus aij~t, which is a tensor functional of q~, can be expressed in terms of aij,
and this in principle provides the natural closure. The existence of this theoretical
closure was demonstrated by Lipsc0mb et al. [27]. However, no attempt was
made by these authors to directly investigate and exploit the resulting relation
between aijk! and aij.
In theory, the natural closure approximation is completely specified by the
above definition. However, direct explicit calculation of aij~:t as a function of
1366

aij is only possible (to date) in the 2D case [40]. Tuned numerical methods are
therefore required in the 3D case to obtain an accurate approximation of this rela-
tion. Hence, two levels of approximation must be considered, viz., the theoretical
natural closure, which provides an approximation of the 4th-order orientation ten-
sor aijkl, and its numerical approximation (which is thus an approximation of an
approximation). To avoid any misunderstanding in the sequel, the term "natural
closure" will concem the theoretical closure, while the term "natural closure
approximation" will be used for the numerical approximation of this theoretical
closure only.
It should be emphasized that the natural closure concept is based on establishing
a relationship between aij and aijkl by neglecting the effect of fiber-fiber inter-
action and starting from isotropic orientation in a hypothetical flow. These as-
sumptions are only necessary to define the closure, which is further used in situa-
tions where they are no longer valid. There is some physical evidence, which is
corroborated by the investigations of Verleye et al. [4,40-42,56,57] and Cintra
and Tucker [45] (see Figure 5), that the natural closure behaves well in many
circumstances, especially for molding problems. However, this cannot be always
true, and in particular the natural closure should not be used when two (or more)
preferential orientations of the fibers are present. These states are highly unlikely
in practical situations.

3.3.1 The 2D natural closure


In 2D problems [40], the full symmetry and normalization conditions impose to

,5 i i i i ,I i i i

1.4 a,, Linear closure ........ .


Quadratic closure . . . . . . .
1.3 Hybrid closure "
1.2 Natural closure
.'" Distributionfunction
1.1 .

1.0 - .....:.,'-- -- ---"_-z_-':"-". . . . . . . . "

0.9 . " " / . 9' " " - ~ ~ ~ ~ ' ~


08
0.7 ~
ot"
0.6 ,,-'~E , , , ,Reducedtime ~: = ~/~ t
0.5 , , ,
0 1 2 3 4 5 6 7 8 9
Figure 6. Prediction of 2D fiber orientation in elongational flow. (From [40]).
1367
write the natural closure in the form

aijkl - ~8 ((l + 4D)~3 -1)S(6ij61d) - (~3 -1)S(aij6kl ) + ~3 S(aijalcl ) , (35)


which results from (29) and (31). On the other hand, the deformation gradient
F~A from fictive initial to actual time can be written in principal axes. Hence,

[FiA]_
[ J
F
0
0
F -1 '

where F is a non-vanishing positive constant. Finally, the current orientation


(36)

vector Pi can conveniently be expressed in polar form.


Therefore, combining (36), (10) and (9.2) with the definition (12), the compo-
nents aij ( - ~, 71 in principal axes) and aiju are easily calculated as a function
ofF:

- , (37)
F2+l 1

[a1111 a~122]_ 1 [F2(2F2 + l) F2 1 (38)


a1122 a2222 2(F 2 + 1)2 F2 2+ F2 '

and thus [ allll a1122]- 1 I~(l+~) ~F] 1 (39)


a1122 a2222 2 ~r/ 77(1 + 7/) "
Expressing (35) in principal axes and comparing the result with (39) shows that
/33 - 1, from which the 2D natural closure is obtained"

aij t - S(a j t) 9 (40)

The 2D natural closure performs very well when compared with other 2D clo-
sures [4,40,51]. Figure 6 illustrates this behavior in simple elongational flow
(v 1 = ~ x 1, v2 = 0 , v3 = - ~ x 3 ) , starting from isotropic 2D orientation
(all =a22 = 0 5 , a12 = 0 ) and using the value C I =0.01. Impressive experi-
mental validations have been obtained in the prediction of Bulk Molding Com-
pound (BMC) injection molding with the 2D natural closure (see Section 5).

3.3.2 The 3D natural closure


When the fibers can rotate in all directions, a 3D closure is needed. Unfortu-
nately, the problem of determining accurate approximations for the coefficients
1368

fll to t6 in (28) in terms of P and D turns out to be extremely difficult in


view of the presence of complex singularities in several expressions. Therefore,
many different formulations were investigated until the present theory was elabo-
rated. To summarize this work, let us first establish the following non-linear
partial differential equation-

aijkq (~pm + a p#a (~im + a ipgq 3jm -- a pjmq tSik - a ipmq r -- aijmp t~qk +

a ijkp 3 a ijmq 3 a ijmq Oa ijkp -- 0 ( 41 )


+2
Oars arsmq - 2 Oars arskp + 2 Oaks aps - 20ams aqs ,

which is a direct (but not obvious) consequence of equations (20.1) and (22) and
the definition of the natural closure. Indeed, from the identity
D F i A / D t - ~ V i / ~ X j FjA, and taking the conditions ~ - 1 and C t - 0 into
account, (20.1) and (22) can easily be combined as follows in differential form :

daij - dFiA FAT a kj - aik FAT dFjA - -aijkl (dFkA FAll + FAT dFia ) , (42)

where aijkt is a function of aij, which is itself a function of Fia. The key point
is to observe that the differential daij in (42) must be exact when aijkl depends
on aij through the natural closure. More precisely, the derivatives
o32aij/CgFkloqFmn and o32aij/cgFmn3Fkt provided by (42) must be equal for any
value of FiA when aijkt is defined by this closure. After quite long calculations
(which must be performed with care for aij is constrained by (15.1) and (16.1)),
this condition provides equation (41), which is the comerstone of the numerical
theory of the natural closure approximation. Also, (41) can be extended to
higher-order closures and, in particular, it could help in defining a closure approx-
imation for the 6th-order orientation tensor in terms of aijkl (by extending the
space of canonical distribution functions, since 14 degrees of freedom are re-
quired). This theory is under investigation.
At this stage, since the invariance properties of indicial tensor notation have
been used to provide (41), the calculations can be pursued with the tensors aij
and Fim in diagonal form. The following notations are used (Figure 7) :

(a11,a22,a33) = (~,rl,() , (a2233,a1133,a1122) = ( X , Y , Z ) , (43)


while, from (24), the invariants P and D read as
(44)
1369

r/=l

~-1/8
m

r r/=O ~=1
Figure 7. Left'triangular domain T of eigenvalues of aij. Right" isovalues of
Z-a1122 in T (the maximum Zmax is 1/8). (From [4]).

Moreover, the normalization condition (16.2) becomes


a1111 + Y + Z - ~ , a2222 + X + Z - 77 , a3333 + X + Y - ~" , (45)
and a l l l l , a2222 and a3333 can be eliminated in terms of X , Y, Z, ~, 77
and ~" (the natural closure is a particular orthotropic closure [45]).
Furthermore, from (28) and (34), the expression of Z writes as

Z - 112/31 + (~ + r/)/32 + 2~r//33 + (~2 + r/2)f14 + (~ + r/)~r//35 + 2~2r/Zf16], (46)


6
with similar formulae for X and Y. Combining these relations with (44) pro-
vides (X + Y+ Z), (X~ + Yr/+ Z~') and (X~ 2 -k-Y/72+Z~ 2) in terms of P ,
D and 131 to 136. After some calculations, and with use of (30), these relations
can be cast in the form :
7 1
1A0 4
2
0 0 --10 /~6
1
-4P 6P 2D -2 10 0
0 X+Y+Z
- 6P 1-4D -1 -2D 1 7 o + + , (47)
2D -1 -3 (D-15P) -2 21 -35 X~ 2 + g r / 2 + Z ~ . 2
1370

with A - - 2 7 P 2 + 18PD + D2 -4D 3 -4P . (48)


Equation (47) proves extremely useful in evaluating the natural closure.
Finally, equation (41) can be transformed after very long calculations (involving
trace operations) into the following set of equations"

[ (4Z + X - r/)(2Z- ~) 1
+ (rt- r162 r L_(4 z + Y - , ~ ) ( 2 Z - r l ) _ ]

X - Z)(2Y - ~)(7/- ~') 1 - 0 (49)


(Y - Z)(Z X - rl)(r - ~) J

with t5 - (~ - 7"/)(7/- ( ) ( ( - ~), and A - ~ 2 . Similar pairs of equations govem


the derivatives of X and Y. It can be proved that no other independent rela-
tionship than these 6 equations can be derived from (41). It should also be ob-
served from (34) that ~, 7/ and ( are are not independent variables and only
the constrained partial derivation operators (tg/0~- 0/tg(), ( 0 / 0 r / - 0/0~) and
( 0 / 0 ~ - 0/0r/) can provide meaningful results. Indeed, when a given function of
and r/ (or ~ and ( , or r/ and ( ) is regularly extended to a function of
~, 7/ and ( (considered as independent variables), each of these operators
provides a unique derivative along the subset of admissible values of (~, r/, ~'),
as defined by (34), whatever extension is performed.

3.3.3 Numerical approximations o f the 3D natural closure


In the 1 st strategy selected by Verleye et al. [41,56] to obtain a closed approxi-
mation of the natural closure, rational expressions were determined to accurately
approximate the coefficients fll to f16- For that purpose, the canonical distribu-
tion function q'c was integrated numerically for many values of the cumulative
deformation gradient FiA, thereby providing a set of pairs (aij, aijkt) from
which a rational approximation of fll to f16 respecting (30) was further obtain-
ed by best fit. The drawback of this method (which nonetheless produced good
results) is that the error decreases only slowly when the degrees of the numerator
1371

0.14 9 . . . . . . .

0.12 Analytical Z ~
Approximate Z - -
/
0.1
~,, / Anal.ytical Y - -
0.08
0.06
0.04
0.02
0.0
-0.02 i i i i i i i i

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0.5 -
0.4 Analytical dY/d~
Approximate dY/dr
0.3 -
0.2
Analytical dZ/dr
0.1 Approximate dZ/d~ - -
0.0
-0.1
-0.2 ,,
-0.3
Figure 8. Comparison between the theoretical and approximated values of
Z -- a1122 and Y - a l l 3 3 (top), and dZ/d~ and dY/d~ (bottom), along
- 0 . The 2 "d strategy was used. (From [4]).

and the denominator of the approximants are increased (in [41,56], these degrees
were 8 and 1). This results from the singularity of the natural closure along the
boundary 3/" of the triangular domain T of admissible (~,r/,~') (Figure 7).
Altemative approaches, taking into account the theory of the natural closure,
were thus investigated.
In the 2 "a strategy [4], the following decomposition was introduced"

Z-/~2~r/+PZ , (50)

where Z is analytical and PZ tends towards 0 when P tends towards 0.


This form is justified by the fact that ~r//2 reduces to the 2D natural closure
1372

(40) when P - 0 . An approximation of Z was obtained from 3 basic


properties of the natural closure. First, numerical experiments showed (and this
should be confirmed with the help of (49)) that
2- 3/10 , (51)

when ~ r / - ~,2 or, equivalently, when


p _ D3 _ ~3 , D - ~ . (52)
Secondly, the following derivatives of X , Y and Z can be calculated from
(49) along the boundary aT when ~" - 0 :

ae, g-r z - - ~ n , ar ac X - a~ ~) y = 1
2
(53)
a a)z_l~r
an ar
(aa~ ara] x - a(an 1
2

with similar equations when ~ = 0 or 7/= 0. Thirdly, the axisymmetric natural


closure [4] (Figure 8) is easy to calculate :
9 either by solving the differential equation provided by (49) along the bisectors
of the triangular domain T :
(1-3~)(8Z-(1-~))dZ/d~+40Z 2 + Z ( 8 ~ - 7 ) + 1/4 (1-~')2 - 0 ; (54)

9 or in parametric form from (42) :

_ (4 cosh 2 S - - 1) cosh s cosh 2s (2 cosh 2s + 1)


,6s s
_

s+

coshs
16sinh4s (55)
S ~ ~ " if o<r (~<s<O),
sinh3s sinh2s

(4COS2S- 1) COSS COS2S(2COS2S + 1)


. . . . . . . . . s-k-
16sin 5s 16sin 4s (56)
coss 1
_:-~_3-s+~2-, if )~<~'_<1 (O<s_<~).
sin s sin s

In both cases, (34) provides the relation ~ - 7 / - ~ 2 ( 1 - ~'), while it can also be

proved that X - Y - ~ (1 - u

8Z).
1373

The approximation (X ap, I?ap, Zap ) tested in the 2 nd strategy was of the form

lnD +(~
Zap ~
I l)I ll
1+/7 lnP In
(57)

with symmetricexpressions for f~ap and I~ap. In (57), P(v) and G(v) are
polynomials in 77= lnD/lnP, which vanish when "c = 0 and are selected in order
(i) to respect condition (51) when equation (52) holds :

and (ii) to provide a very close approximation of the axisymmetric natural closure
when ~ - U . This was achieved by constraining X ap, Yap and Zap to present
the same expansions (up to a given order) as )(, I7 and Z around ~ - 0,
= 1/3 and ~=1. In particular, complex calculations around ~ = 0 produce the
following expansions :

Z= 1 ( 1+ +..., X -Y- 1+ +..., (59)


8 4

which provided conditions for the selection of polynomial coefficients in (57).


Equations (59) illustrate the singularity exhibited by the natural closure in the
case of a nearly 2D orientation.
When the expressions of f~ap, Yap and Zap are selected, it becomes easy to
calculate Xap + Yap + Zap, ~ Xap + 71Yap + ~ Zap and ~2 Xap + 02 Yap + ~2 Zap
in terms of P and D from (44) and (50), and thus, in principle, to obtain ap-
proximate formulae for the coefficients fll to [36 by means of (47) and (30).
However, a significant difficulty appears at this stage, since the determinant A
defined by (48) vanishes identically along the bisectors of the domain T. There-
fore, the equations cannot be solved in the axisymmetric case. This problem is
worse than it seems, since the fight-hand side of (47), which vanishes identically
when the exact natural closure is used and A - 0, never did vanish when an ap-
proximate closure was introduced by means of the 2 nd strategy (a simple formula
satisfying this condition was impossible to obtain). Hence, artificial singularities
were introduced for nearly axisymmetric orientation states. To circumvent this
difficulty, either the rational approximants determined by means of the 1st strategy
or the expression (57) were used for low values of A.
1374

Figure 9. Isovalues of (Zap -Z)/Zma x in % (2 ~d strategy). (From [4]).

Whereas very good results were obtained by using this 2 nd method [4], the
solution adopted to remove the artificial singularities was not completely satisfac-
tory, since a closed analytical expression was expected for the closure approxi-
mation. Moreover, since in molding processes the fibers are most often strongly
aligned by shear and elongational effects, the vicinities of the comers of the do-
main T must be given particular attention. In particular, overlapping (and thus
non-regular) solutions must be avoided. This is the reason why a 3 rd strategy,
which is still under development [42], was investigated. Recent work also
suggests that non-linear algebraic relations linking fll to /56 can be obtained by
equaling the cross 2 "d derivatives of Z (obtained from (49)). Hence, a very high
accuracy should eventually be reached.
The results obtained by using the 1~t strategy (with rational approximants) were
compared by Cintra and Tucker [45] to their best fitted orthotropic closure (see
Figure 5). They showed that this natural closure approximation generally behaves
very well, whereas slightly better results are obtained with the optimal orthotropic
closure. We believe (but this needs confirmation) that this can be related to the
lower quality of the 1~t version of the natural closure approximation in the vicinity
of the boundary (and especially the comers) of the domain T. Further numerical
experiments are needed. On the other hand, the natural closure approximation
does not oscillate in the same way as the best fitted orthotropic closure for low
values of C I [45,88].
The isovalues of the approximate solution Zap obtained with the 2 nd strategy are
represented in Figure 7 along T. The good quality of this approximation is illus-
trated by Figure 8, which depicts the values of X, Z, X ap, Zap and their
1375

derivatives along the bisector ~ = 7/, and by Figure 9, where the relative error
(Zap- Z)/Zma x (in %)is represented over T. Near to the comers, the error is
more important because another approximation was used.

3.4 D o u b l e - L a g r a n g i a n technique; deformation tensor model


In order to avoid the non-physical behaviors exhibited by the quadratic and HL1
closures, Szeri and Leal [44] elaborated the "double-Lagrangian" computational
technique, which consists in solving microdynamical equations of the form (6) in
a space that is Lagrangian for both the coordinate (x ~ and orientation (pO)
variables. This method requires to express W in the Lagrangian form [44a]
~F(x ~ p~ t) - v~o /det(0pi/oap 0) , (60)

where 0pi/Op ~ (x ~ p0, t) represents the Jacobian matrix of the solution of (6),
while q'0 - W ( x~ p0,0). Hence, a system of ordinary differential equations is
found, which can be solved numerically by generating a large number of Lagrang-
ian fibers. Details are in reference [44b]. This serf-adaptive scheme is of particu-
lar interest for those problems where a good closure approximation is not availa-
ble, or when steep orientation gradients are present in the boundary layers.
In a subsequent work, Szeri and Lin [46] developed an accurate method to solve
the fiber orientation problem by tracking an equivalent deformation gradient
tensor FleAq (x,t) obeying the equation

Dt

where D/
- -~j + ( ~ - l)dij + 3hDI (BiJq(-1)-3B~c(-1) ~iJ) F ~
1 ,

is Brownian diffusivity, while Bi~q is defined by Bi~q-FieAq Ff~,


(61)

and h-h(Feq(T).Feq) is an average value of p~176 (pOEU),


which is designed to match selected asymptotic states. The tensors aij and
aijkl, which are recovered from the relation Pi -- Fleaq pOA[~'P0[I-1 by exact or
Gaussian quadrature, are thus linked by the theoretical natural closure [4,40-42,
56]. This method extends the theory of Lipscomb et al. [27] (equation (10)) to
the case D I ~ 0 (see also [77,94,100,101]). First tests [46] indicate that the
"deformation tensor model" provides slightly less accurate results in shear flow
than the orthotropic fitted closure [45] or the natural closure. Additional experi-
mental comparisons are necessary.
1376

4. A P P L I C A T I O N TO SUSPENSION F L O W S IN N A R R O W GAPS

Investigating the flow of fiber suspensions in narrow gaps is of particular in-


terest in view of the model simplifications resulting from this dimensional hypoth-
esis ( e - h/L << 1 ), and the direct applicability of the model to injection and
compression molding simulation [ 1-9,38-42,49,50,53-56,74,78-81,85,87,88,102-
104]. It must, however, be noted that the gap between model and actual process
remains quite wide, especially concerning the effect of the polymer rheology and
the fiber concentration and flexibility on the flow and fiber orientation. Hence,
only rather idealized simulation tools are currently available.
In an important paper, Tucker [2] identified 4 suspension flow regimes in nar-
row gaps, considering rigid slender particles and an incompressible Newtonian
solvent. His classification was based on defining 2 dimensionless numbers,
namely, the particle n u m b e r N p ( w h i c h scales the macroscopic extra-stresses
caused by the particle resistance to flow stretching effects) and the order of mag-
nitude of out-of-plane orientation 6. Neglecting the Brownian effect and recast-
ing equation (23) in the form

~ij -- 27/(1 + CO) [dij + Np aijkl d u + N s (dikakj + aikdkj )] , (62)

provides a convenient way to define and understand the role of N o (the shear
number N s is negligible for slender particles). Tucker [2] proved that Pi, aij,
aiju ... are all of the order O(6S), where s is the number of out-of-plane com-
ponents among the indices i , j , k .... while, from (20.1), 6 - m a x ( e , C ~ i 3 ) .
(Note that aiju is scaled in [2] from the definition (12.2), without considering
the effect of the closure approximation. First inspection reveals that classical
closures such as the quadratic, hybrid ... closures, but not the linear closure, are
consistent with this scaling.)
The flow regimes identified by Tucker [2] are :
9 I. The decoupled lubrication regime ( N p S 2 << 1, S > e), which allows to
separate the flow and orientation calculations since the gapwise shear stresses
are governed by the solvent viscosity (without orientation effect) and dominate
the in-plane momentum balance.
. II. The coupled lubrication regime (Np 62 > _ 1 , e << 6 << 1), where the gap-
wise shear stresses dominate the in-plane balance, with orientation effects,

~'3a - 2r/(1 + CO)[d3a + 2Np a3ap3 dp3 ] (63)


1377

(Greek indices, from 1 to 2, indicate the in-plane directions).


9 lIl. The general narrow gap regime ( N p S 2 _>1, S = E ), where few simplifica-
tions are allowed.
9 IV. The plug flow regime with shear boundary layers (NpC 2 >> 1, ~ << c),
where the velocity profile is flat across the gap and generates two wall lubricat-
ing layers of thickness [3-O(LNpl/2), while both in-plane (fiber-induced)
stretching stresses and wall shear stresses are significant.
A different approach was selected by Barone and Caulk [38] who, considering
the compression molding of Sheet Molding Compound (SMC), assumed a priori a
plug flow in the cavity, with in-plane stretching stresses and wall friction layers.
When in-plane orientation is random, this model is equivalent to the regime IV of
Tucker [2]. However, while the latter was rigorously derived for rigid fibers, the
former is phenomenological and can be considered whenever the suspension is
concentrated and the fiber length prevents out-of-plane tumbling. With long flex-
ible fibers, the lubricating layers (of thickness 13) must be considered as regions
where less fibers are present (due to the wall effect) and whose rheology is gov-
erned by the solvent viscosity, while the plug flow results from kinematic fiber-
fiber interactions. The applicability of this model depends on the solvent rheolo-
gy and the real size of the fibers, which can be affected by the gates, the walls ...
Experiments [33,102,105] indicate that injection molded long-fiber composites
have a significantly thicker core and a thinner shear layer than short-fiber parts.
Finally, it should be noted from (62) that the ratio between the powers dissipated
by out-of-plane shear stresses and in-plane stretching stresses is of the order
O(L2fl-lh-lNpl). The latter can thus be neglected in the in-plane momentum
balance when L is large, since the order of magnitude of ]3 does not increase
with L for flexible fibers as in regime IV. This justifies the thin charge approx-
imation of Barone and Caulk [38].
The remainder of this section will be devoted to decoupled models. The coupl-
ed lubrication regime II of Tucker [2] was investigated by Chung and Kwon [49,
50], who extended the Hele Shaw model to that case (see Section 5).

4.1 Decoupled mathematical and numerical models


Two problems are posed, considering a generalized Newtonian rheology and an
incompressible suspension.
9 P1 : decoupled lubrication flow of a short-fiber suspension [2-4,9,41,42,49,50,
56,80,81,87,88,104]. The Hele Shaw approximation is used and orientation is
3D (using a 2D model produces erroneous results; if simplifications are
1378

required, imposing constant out-of-plane fiber orientation components is


possible, but their selection is critical [88]). This model applies rigorously
when the suspension is dilute.
9 P2 9plug flow of a long-fiber suspension with shear layers [2-7,9,38,40,78,79,
85,103]. The thin charge model of Barone and Caulk [38] is used and the fiber
orientation field is 2D and decoupled. This regime occurs when the suspension
is concentrated and the in-plane stretching stresses can be neglected with re-
spect to friction in the skin layers.
For both problems, the kinematics are governed by the pressure gradient and the
fluidity. Letting v-~ denote the gap-averaged in-plane velocities, the volume
flow rate can be written in the form
2hV a - - S Op/Ox a , (64)
where the fluidity S is calculated differently in P1 and P2. The pressure field is
governed by mass conservation"

+ ~ S - 0 (65)
& 3x~
where 3h/oat is included to provide a single model for injection and compression
molding.
The Hele Shaw approximation, which is used to model problem P1, has been
presented in Chapter IM. Some points need to be recalled :
9 The pressure field is approximately constant across the gap.
9 Assuming a power-law viscosity 77- moe b p - a T ~ n - 1 , where T denotes tem-
perature while n, m0, a and b are material constants, the shear rate can
be approximated from the relation ~2 _ (oava/o3z)2 (with z - x3).
9 The fluidity is given by the expression

S - 2h [[Op/&,~l[~/n-~e-bp/" m~ ~/"~ , (66)

where ~ denotes the gap-averaged "profile function" 09, which is propor-


tional to the velocity profile in the gap 9

-
jz (e- z0) le- z0I1/n-1 eaT(~)/nd2 (67)

with the wall no-slip condition o g ( - h , T ( . ) ) - 0 . In (67), z0 stands for the


level at which ~ ' - 0. The profile function depends upon the temperature pro-
file T(.) only. Its average is calculated by integration by parts 9
1379

rhl eaT(z)/nd z
2h~- j_hlz -- Z01Vn+l . (68)

In-plane velocity components can be recovered from the relation"

-- ~ e-bp/nmoVn(o(z,T(.)) . (69)
va - 3x a Oxt~

In problem P2, the gap-integrated in-plane momentum equations read as


- 2 h O p / O x a + "c3a ( z - h ) - "c3a ( z - - h ) - 0 . (70)

Letting 13+ and 13- denote the upper and lower lubricating layer thicknesses
(with ~ - / h , ~+/h << 1), the velocity gradient and shear rate in the skins are"

{ ~v~l~z
~v=/az
- - ~1~+
- ~-~/~,
, +
i, - I I ~ l l / ~
e -I1~=11/~
<h-~ + _<z<_h>,
(-h_<z_<-h + ~ ) ,
where v~ stands for the constant in-plane core velocity. Furthermore, assuming
a power-law skin viscosity, and using (70) and (71) with the approximation
"r3a - 7"10v a I O z , the core velocity norm I1~11 ca. be expressed in the form

][Oqp 1In le_aT+ e_aT-I-Vn


II~ll- (2h)Vn -~a moVne-bp/n -~n + ~~-n ' (72)

where the temperatures T and T- in the upper and lower skin layers are
supposed to be constant (and this is a restrictive assumption). Introducing (71)
and (72) into (70), the core velocity is therefore approximated by the expression

V'-'a= ~ t I1~11
]I~ (2h)l/n mo-1/ne-bp/nle-aT+
~+~+n e-aT- I ' (73)

which is of the form (64), with

s I1~11
n-1(2h)l/n+lmo-Vne-bp/n~+~-~~+n
/ea+ / (74)

From this discussion, it results that the same equations (64) and (65) will be
used to predict the kinematics in P1 and P2. Thermal effects can play a non-
negligible role. The effect of chemical reactions could be easily considered (cf.
also [ 104]). The solution of these flow problems was investigated in Chapter IM,
1380

to which we refer the reader for any details.


Once the flow is known, fiber orientation must be calculated. It is sufficient to
consider the 2~ tensor aij, whose evolution is governed by (20.1). In the
framework of the finite element technique used in the MOLDSYS software (see
Chapter IM, Section 4), aij is a 2D or 3D additional field, which is integrated in
a similar manner to T. Details will not be repeated and we will focus on some
particularities of the algorithm related to the tensorial nature of equation (20.1).
To cast the equations into a weak f o r m , the residual of (20.1) is multiplied by an
arbitrary tensor test-function aij (defined over the filled zone f~(m) of the mid-
surface) and the result is integrated over f~(m) :
r-I
d "
(aij + 2~,aijud u + 2Ct ~' N aij ) a(i dS - 0 . (75)
~(m)

Recalling that aij is constrained by equations (15.1) and (16.1), while aij and
dij must have an equal number of degrees of freedom in order to produce an
equal number of discrete unknowns and equations, d o is itself constrained 9

aij -- d j i , aii - 0 . (76)


On the other hand, when the average mass equation (65) is put into a weak
form, integration by parts is carried out o v e r ~,-~(m) and the weak solution
p(xu,t) belongs to the Sobolev space H l ( f ~ ( m ) ) . The velocities provided by
(69) or (73) belong to L2(f~ (m)) and theft in-plane derivatives are measures
[106]. Hence, whereas (75) does not violate any distribution product rule when
aij and dij themselves belong to H l(~(m)), integrating this equation is diffi-
cult and much better results are obtained by integrating by parts every term in-
volving velocity in-plane derivatives (Verleye [4], Crochet et al. [9]).
This operation is delicate in 3D problems when 2h is not a constant over
f~(m), since the different fluid layers are no longer parallel to the midsurface (par-
allel layers are defined by z - c o n s t a n t , while effective layers obey the equation
z/h - constant ). To solve this problem, let f denote a generic quantity and de-
free a*/ ana a*I /ax by fixing z/h (instead of z ) when derivation is
performed"

a*I oy ah oy cg*f _ oaf t- zOhOf .


_

(77)
& & h&&' 0x a oaxa h 0Xa 0z
1381

Splitting the out-of-plane velocity component w as follows :

W - - - - +V a +w , (78)
h
it results from mass conservation and the wall no-slip condition that
9 3" 3" h
W ---~~hd2+<Sz zV~d2 . (79)

Combining (77), (78) and (79) shows that

Df _- c)*f + v a c)*.__._f_+f w* -oaf


- (80)
Dt Ot oaxa cgz '
which is a convenient form because in-plane advection is considered along effec-
tive layers. (In practice, due to numerical stability concerns, w is often approxi-
mated by setting w* to zero in (78), which corresponds to imposing a locally
convergent, divergent or parallel velocity profile; see [107] and Chapter IM, Sec-
tion 3.) Hence, in order to integrate by parts f Og/oaxa (where g is a 2 nd ge-
netic quantity), the following formula must be used :

[.fOg a*f z Oh Og (81)


~2(m) <dS -- S ( g+f ) dS + S f g nads '
~-~( m ) ~ 0~ h O~Xo~ ~ Z ~-~(m )

where n a denotes the in-plane outer normal along o ~ (m) . In the fight-hand
side of (81), the 1st term is easy to calculate in a layer-by-layer procedure, while
the 2 nd term is indispensable.
To complete the weak formulation of (20.1), boundary conditions must be treat-
ed. Along the gates, aij is specified by means of essential conditions. Along
the side walls, no treatment is mandatory because the characteristics of (20.1) are
tangent to the side walls. However, the normal in-plane velocity may be usefully
set to 0. The conditions to be applied at the flow front and along the singular
lines of the midsurface (namely, the edges, abrupt changes of thickness, bifurca-
tions ... of the part) are examined in a separate section.
After the evolution equation of aij has been put into a weak form and the
boundary conditions have been treated, the semi-discrete scheme is built up in
two steps. First, aij is approximated by the discrete sum

aij - ~.~ r (xo~) aij (m) (z,t) , (82)


m
1382

where aij (m) (z,t) and qb(m) (xa) denote the nodal values of aij and the shape
functions. The tensor a ijkt is discretized in the same way"

aijkl -- Z ~)(m) (Xot ) aijkl (m) (z,t) , (83)


m
without loss of accuracy [9] (evaluating the closure approximation of aijkl at the
sole mesh nodes is very efficient). Secondly, in order to obtain an equal number
of equations and unknowns, the test-functions aij are also discretized taking
constraints (76) into account"

[ ~n) 0 0 0
^
0 0
o , o
o -r 0 -r

o 8(n) 0 0 0 ~(n) 0 0 0
o ; 0 o 0 (84)
o ~,., o o ~(n, 0

o -~n)' ~~ '
for a 3D or 2D problem, respectively. In (84) and (85), and (82) and (83), the
^

functions r and q~(l) differ, due to the effect of applying a modified SUPG
technique (cf. Chapter IM, Section 4).

4.2 Effect of the fountain flow; abrupt changes of thickness and bifurcations
In 3D Hele Shaw flows, the fountain effect results from the deflection of the
core fluid towards the walls of the narrow cavity when it approaches the flow
front, as a consequence of the fluid-wall no-slip condition. Modelling this effect
is an important issue, since appropriate temperature and orientation conditions
must be imposed at the front to obtain correct simulation results. Both the flow
and its influence on the fiber orientation must be considered.
The effect of the fountain flow on the kinematics has been thoroughly investigat-
ed in Chapter IM and only the principal results will be stunmarized. Basically,
the Hele Shaw approximation generates a singular perturbation problem [108],
whose outer zone comprises most of the flow domain and is governed by the
1383

\\\\\\\\\N\\~z\\\\\\
\\\\N\\
. . . . !

j--)
i
i
I
dx!2
\\\\'N\\\\\\\\
dZl~z ' 7 ...... a 2: . . . . . . . . . . . . . . . . .

dZl~z
I z z~ ~ ~ - ~ dz 2
__ l_l _~ /////// ,I t11111111111'1/
x1

Figure 10. Simplified models for the effect of fountain flow (left), and for the
flow across an abrupt change of thickness (fight).

lubrication approximation, whilst the inner front zone has a characteristic dimen-
sion of the same order of magnitude as the gap thickness. Matching the inner and
outer zones imposes to select precise front boundary conditions. The fountain
flow thermal model developed by Dupret and Vanderschuren [107] was thus
based on the theory of matched asymptotic expansions. The following points
should be stressed :
9 The Hele Shaw approximation governs the whole flow domain.
9 The front region is infinitely thin and consists of straight segments perpendicu-
lar to the midsurface.
9 Heat diffusion and viscous heating are negligible in the front zone. The core
material points entering the front at a given level leave the front at another level
while keeping the same temperature.
This model provides a front thermal boundary condition for each level in the gap
where the material points move slower than the front (according to the Hele Shaw
model). This agrees with the fact that the characteristics of the simplified energy
equation enter the flow domain in these layers [9].
Once the flow is modelled, its effect on the fiber orientation can be investigated.
This problem is complicated, since orientation is governed by the motion, rota-
tion and deformation of the elementary material volumes that cross the front re-
gion. Ideally, the effect of both the inner and outer asymptotic flows should be
taken into account. However, calculating the inner flow means solving a free
boundary problem at each time and position along the front, whereas a unique
inner solution can be determined once and for all only if the viscosity is a con-
stant. Approximating the viscosity by a constant in the inner zone and predicting
the fiber orientation by matching techniques thus represents an attractive solution,
which has not yet been implemented.
The method developed by Verleye and Dupret [4,9,40,53] (which is very close
to the model of Bay and Tucker [81 ]) extends the thermal model of [ 107]. Shear
1384

deformation is neglected when the suspension crosses the front region. Consider
in Figure 10 an elementary parallelipiped entering the front at level z 1 and leav-
ing it at level z 2 , and let (dx 1,dy 1,dz l) and (dx 2 ,dy E ,dz a) stand for the initial
and final side lengths of this material volume (which remains a parallelepiped
since shear deformation is neglected). The deformation gradient F~/(l-a) from
z I to z a can be expressed in terms of the profile function to and its average
defined by (67) and (68). Indeed, from mass conservation, the relation link-
ing the associated pairs (Zl, z2 ) is Sz 2 ( ~ - (3.)(Z))dz - O, from which it results
1

that (~-(.o(z2))dz 2 - ( c o ( z 1 ) - m---)dZl, while mass conservation in the 2D front


flow also produces the relation dx 2 dz 2 - dx 1 dz 1. Hence,

-- ~" -- ('O(Z2) 0 0
(.O(Zl) - ~
r__,lF,~l_2)/ _ 0 1 0 (86)
L -J j
0 0 - o~(zl)-~
~" _ (.D(Z 2 )

where the negative signs account for the 180 ~ rotation.


In order to approximate the relation relating a/~2) to a hi) , Verleye and Dupret
[40] analyzed in a 1~t step the effect of the fountain flow in a theoretical problem
where in-plane orientation was assumed, while the evolution equation of aij was
supposed to be governed by the 2D natural closure (see Section 3.3). In that
case, the squares of a~1) and a~2) can be expressed as functions of the defor-
mation gradients F/(a1~ and F/(A2) from fictive initial to actual time"

amnUnm a/(]) a , ( 1 ) - ( F f f ~ ) F i ~ Fi(A1) F)A ) ,


~:J - --1 (87)
[~am(2)...(2)]
Unm a[~ ) akj ~'kB r kB ) F,'2
since this latter relation is valid in principal axes (from (36) and (37)) and satis-
fies frame indifference (because Fi(A1) F),~) and Fi(A2) F),~ ) are the Finger strain
tensors from fictive initial to actual state). Therefore, assuming that fibers are
slender and that fiber-fiber interaction is negligible during the short time the fibers
undergo fountain flow, the relative deformation gradient Fij (1-2) is related to
F/(2) and F(1-2) F ( ~ ) - F i (2)
F/(A1) by the relation "ij Eliminating F/(2) thereby
1385

from (87.2), and then FiCA 1) from the result by means of (87.1) provides after
some calculations the relation
a (z) = (F(1-Z).a(1).a(1).F (1-2) T)l/2
(88)
tr[F ]
where the conditions (15.1) and (16.1) have been taken into account. The foun-
tain flow model of [40,53] consists in using the jump condition (88) as an orienta-
tion front condition for the layers moving slower than the front. Whereas quite
restrictive assumptions were introduced to establish (88), this model can efficient-
ly be used in more general situations, in particular when the fibers are not parallel
to the midsurface or when fiber-fiber interaction is not negligible.
The above model can also be used in the presence of other singular regions,
such as abrupt changes of thickness or midsurface bifurcations (Figure 10). Con-
sidering the theory of singular regions developed in Chapter IM, this methodolo-
gy provides a convenient way to treat all the boundary and jmnp conditions asso-
ciated with the prediction of fiber orientation in Hele Shaw flows.
On the other hand, there is, in principle, no fountain effect when the 2D plug
flow model applies. Nonetheless, bifurcations and abrupt changes of thickness
can be treated by using the above orientation jump model. As an example, the
deformation gradient tensor Fij(1-2) across an abrupt change of thickness is
0 0
/]7(1-2)/ -- 0 1 0 (89)
t'ij _

0 0 h2/h 1
where 2h 1 and 2h 2 denote the upstream and downstream thicknesses. The ex-
pression of Fij (1-2) provided by (89) can be introduced into (88) for calculating
the jump of aij. Applying this model to long-fiber suspensions however requires
some caution. In particular, experiments indicate that the velocity profile is not
completely flat in some circumstances [67,102,105], with a consequently low, but
non-vanishing fountain effect (in which case the fibers in the front zone can be
bent across the whole gap [33,105]), while a plug flow is clear in other cases
[38]. The model to be selected therefore depends on the material (SMC, BMC,
GMT ...) and the processing conditions. Additional modelling effort is needed.
1386

5. E X A M P L E S O F N U M E R I C A L SIMULATIONS AND EXPERIMEN-


TAL VALIDATION

Few publications report quantitative data describing the orientation field of


injection molded fiber reinforced composites and, in addition, most results are
presented in such a way that it is difficult to compare the observations with theo-
retical predictions and to draw general conclusions. Most often, the publications
focus on the effect of the orientation state on the suspension rheology or the

r
0.8
I
" Point A
Num. a~ m
Exp. al,

0.2
2h = 3.18 mm
0.0

-0.2
-0.4
-1.0 -0.6 -0.2 0.2 0.6 1.0
m
, 1.2 | i |

(b) . . . . . . Num. a t , - - (C) / . . . . . Num. a , l - -


Exp. ajj ~-~
Exp. a,, ~-' t 1.0
Point B Point C
t tt 0.8
t t 0.6
0.4

0.2 -~t .]t Num.


~r ' . E x p_'a,3~--f
_
9 9 9 9 9 9 9 - tl i , . I . taF 0.0 , ,,

-0.2
i i i i i | I I I -0.4 i i i i I I I I i

- 1.0 -0.6 -0.2 0.2 0.6 1.0 - 1.0 -0.6 -0.2 0.2 0.6 1.0
z~ z~
Figure 11. Filling of a short-fiber reinforced nylon disk" sketch of the part, and
numerical and experimental profiles of a l l ( = arr ) and a13 ( - arz ) across the
gap 9 (a) r - 19.7 m m ; (b) r - 36.2 m m ; (c) r - 47.5 m m . (From [4]).
1387

1.2 ..... i 9 i i ' " i '

(a) Num. a , l - -
W = 25.4 m m 1.0 Exp. al~ ,*--,

0.8
L = 20_3._2 m m

B
C 0.6 t

0.4

.......... 0.2 Num. a~3m


Exp. a i 3 ~..-,
Xl 2h = 3.18 m 0.0 ~

-0.2

-0.4 I I I I ,,,

0.0 0.2 0.4 0.6 0.8 1.0

(b) Nurfi. a,, (C) um. a l l w


Exp. a~l ,*--,
Exp. a l l
t t |
" i
t"

Point B

Num. a ~ 3 m Num. al3_. -


Exp. al3 ,.,---, Exp. a ~3 ,-,-.-,

9 I___
9 9 " " I L_ \' ~f~----' I ' ! I
9 !
, -0.2 ~ - - - - - - - - ~ - - - - - - - -
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
z~ z~
Figure 12 Filling of a short-fiber reinforced nylon strip" sketch of the part, and
numerical and experimental profiles of al 1 (-- axx ) and a13 (-- axz ) across the
gap 9 (a) x - 9.1 mm; (b) x - 54.1 mm; (c) x - 167.0 mm. (From [4]).

thermo-mechanical properties of the product. However, since the physical behav-


ior of the solid part can be quite well predicted by means of homogenization
methods based on the orientation field description (Lielens et al. [7]), it is of
prime interest to validate the orientation prediction models.

5.1 Filling of short-fiber reinforced parts


In this section, we analyze the filling of a center-gated disk (Figure 11) and a
film-gated strip (Figure 12), using experimental results provided by Davis [109]
and Bay and Tucker [81 ]. The reader is referred to these publications for further
1388

detail as to the experimental conditions and thermo-physical data. The injected


material is nylon 6/6 reinforced with glass fibers (~ = 43 % weight ratio, average
l = 210pro, average d = 11 pm). It should be noted that a rather complex analy-
sis is required to characterize 3D non-planar orientation fields.
The filling of a center-gated disk represents a useful benchmark, since it is com-
monly encountered in mold filling operations. As the flow decelerates radially,
the in-plane extensional velocity gradient tends to align the fibers perpendicular to
the flow direction in the core of the cavity. Across the thickness, however, shear-
ing competes with extensional effects and tends to align the particles with the
flow in the shell layers. Other factors that complicate this analysis are the radial
dependency of the velocity gradient, and the fountain flow effect (which
generates the skin layers).
This problem was solved by Bay and Tucker [81] and Verleye [4,42], using the
non-isothermal generalized Newtonian Hele Shaw flow model with exponential
and power-law dependence of the viscosity upon temperature and shear rate, res-
pectively. Fountain flow was taken into account, while the packing stage was not
considered. Decoupled orientation predictions were performed using either the
hybrid [81] or the natural [4,42] closure approximation, and an interaction coeffi-
cient C~ = 0.002. Altan and Rao [87] also studied this problem by means of
steady Newtonian flow simulation and decoupled prediction of the distribution
function. The interaction coefficient was adapted to achieve best fit with experi-
mental data.
The numerical predictions of Bay and Tucker [81b] look very well, but do not
display satisfactory quantitative agreement with the experimental results through-
out most of the disk. In particular, far from the inlet gate, the shape and thickness
of the core region and its dependence upon the radial distance are not correctly
predicted. The small out-of-plane fiber orientation is overpredicted and the core-
shell transition is placed too close to the mid-plane (see Figure 8 in [8 l b]). This
discrepancy results from using the hybrid closure approximation, as confirmed in
[4,42,87], since the exact transient or steady-state orientation fields cannot be
obtained by this way [87]. Much better agreement could certainly be obtained
from the best fitted orthotropic closure approximation [45]. The orientation
results obtained by Altan et al. [87] are in slightly better agreement with reality.
Unforttmately, their simulations were performed without considering the moving
front effect, and the orientation structure is not accurately described in the regions
affected by the fountain flow.
The natural closure approximation behaves much better and allowed Verleye [4,
42] to obtain very good agreement between experiments and numerical predic-
1389

tions, except near to the gate. In Figure 11, the thickness of the core region and
its evolution are quite well predicted. Furthermore, the correct fountain flow and
frozen layer treatments induce very good prediction of the orientation in the skin
and shell layers (whose mastery is essential for realistic prediction of the product
thermo-mechanical properties [7]). Around the gate, orientation was never pre-
dicted accurately since, as the Hele Shaw model is no longer valid in its vicinity,
the temperature, pressure and orientation fields were approximated as inlet
boundary conditions into the cavity. Fortunately, the effects of shearing and elon-
gational flow rapidly erase the discrepancies downstream from the gate.
Bay and Tucker [81] also investigated the injection of a film-gated strip (Fig-
ure 12). The general orientation behavior was well captured by the simulations,
and some orientation components were correctly predicted, while the correlation
between experiments and simulation was of lower quality for the other compo-
nents (see Figure 5 in [8 lb]). The origin of this discrepancy must again be found
in the use of the hybrid closure. For the same problem, Verleye [4] obtained
excellent agreement by using the natural closure approximation, the only discrep-
ancies being observed near the gate (in point A of Figure 12).
It should be noted that the layered structure obtained by simulation and experi-
ments is similar to that of the center-gated disk, except that there is no longer a
skin layer. Along the mid-plane, the core orientation is not significantly modified
after the fibers exit the gate. Near the walls, in the shell layers, shearing aligns
the fibers in the flow direction. The thickness of the different layers changes
along the flow, this change being a function of the flow rate, the gap thickness
and the polymer material properties. Near to the cavity extremity, Bay and
Tucker [81b] predict a core layer thinner than the experimental one, and this
might be related to the fountain flow model implemented. However, the effect of
the packing stage, and possibly the in-plane viscosity, which are not considered in
[4,42,81 ], should be analyzed before drawing final conclusions.
The parameter analysis of Bay and Tucker [8 lb], confirmed by Verleye [4], pro-
vides key information, for various shapes, about the influence of the injection and
material parameters on the orientation field. The process parameters that
significantly affect fiber orientation are identified, together with the physics to be
included in the model for good orientation prediction. The influence of the wall
temperature proves to be limited [49,81b,104], although it has some impact that
will not be detailed. The inlet temperature has a reduced effect because it can
only vary over a narrow range. The effect of the flow rate is much more signifi-
cant, since it govems the thickness and orientation of the skin layer, which could
even disappear in very fast filling processes [81b,110,111]. This last conclusion
is directly drawn by comparing the fast filling of the film-gated strip and the slow
1390

filling of the center-gated disk. Conversely, the transition between the core and
shell layers is less stiff in the case of fast filling than for long injection times. Bay
and Tucker [81b] finally observed that the sensitivity of fiber orientation to the
velocity profile and the polymer rheology is high, basically since the orientation
profile cannot be accurately predicted if the velocity profile is inaccurate.
With a view to investigating the role of the suspension rheology, Chung and
Kwon [49,50] developed a simulation program to predict the injection of short-
fiber reinforced thermoplastics in arbitrary 3D mold cavities. Their lubrication
flow model was based on combining the Hele Shaw approximation with the Dinh-
Armstrong model, in order to couple the velocity and orientation predictions. The
hybrid closure approximation was used. The filling of a tension test specimen
and a center-gated disk was analyzed. The presence of different layers across the
thickness was well predicted, and the same conclusions as Bay and Tucker [8 l b]
were obtained from parametric analysis.

5.2 Filling of long-fiber reinforced parts


Jackson et al. [74] analyzed the fiber orientation in thin compression molded
parts by integrating the distribution function evolution equation with a non-van-
ishing interaction coefficient. Very good qualitative correspondence between
predictions and experiments was obtained by assuming a plug flow across the
thickness (without fountain flow effect) and considering a 2D orientation field.
Later, Reifschneider and Akay [85] studied the compression molding of a truck
head with multiple charges, whose thermo-mechanical properties were predicted
by means of the Halpin-Tsai and Scharpery equations.
Verleye [4] analyzed the filling of a BMC box (Figure 13). Although the poly-
mer was a thermosetting resin, curing was neglected since it takes place mainly
after filling. The coupling between flow and orientation was also neglected
(whereas it is known that in-plane viscosity can play a significant role in some
cases). Experimental data were provided by the Owens Coming company. The
simulation was performed by adapting the power index and using the 2D natural
closure. As a result, the relative orientation state was very well predicted, in par-
ticular near to the gate, the welding line and the last flow front (where agreement
is difficult to achieve), while the pressure drop was less accurately estimated.

Figure 13 (next page). Filling of a thermoset box: (a) finite element mesh; (b)
successive fronts; (c, d) predicted fiber orientation from 2 viewpoints; (e) experi-
mental orientation. (From [4]).
1391
1392

Figure 14 (next page). Compression molding of a container" (a) initial load and
successive flow fronts; (b) final fiber orientation field; (c, d) comparison between
the mechanical effect of the flow-induced orientation (c) and a reference case
with planar isotropic orientation (d) (isovalues indicate the von Mises stress, and
w is the camber); (e) example of temporary mesh; (f) transient orientation field
during filling. (From [7]).

5.3 Calculation of thermo-mechanical properties

This section will be closed by recalling that the ultimate aim of fiber orientation
prediction in molding or other forming processes is to correlate the thermo-me-
chanical properties of the product with the final orientation distribution in the part
and the processing conditions. A single example, from the work of Lielens et al.
[7,112], will be considered to illustrate this issue, which is beyond the scope of
the chapter. See also [5,6]. Using the predicted fiber orientation distribution
represented by the 2nd-order tensor aij, the authors developed a homogenization
theory in order to evaluate the stiffness, thermal stress and thermal conductivity in
multiphase composite parts with inclusions of arbitrary orientations. It is worth
noting that orientation averaging over the aggregates was directly performed by
using the aij and aijkl tensors, the latter being estimated by means of the natural
closure approximation.
The simulation results, which concern the compression molding of a 5mm thick
container, are depicted in Figure 14. In view of the part's symmetry, only one
quarter of the geometry is represented. In this problem, the abrupt contraction
located at the border of the upper horizontal face, and whose thickness ratio
changes when compression progresses, has a strong impact on the downstream
orientation distribution, which in tum direcly affects the thermo-mechanical
properties in the molded part. Using these results, it was possible to calculate the
reaction of the container when clamped on its lower horizontal face and loaded
with an internal pressure of 1 bar [6,7,112]. The computation software SAMCEF
was used for that purpose.

6. CONCLUDING REMARKS AND A C K N O W L E D G M E N T S

This chapter was first devoted to analysing the general behavior of fiber
suspension flows, with a particular attention focusing on the decoupled prediction
of the fiber orientation evolution in a Newtonian solvent. Since orientation
1393
1394

tensors represent the best approach to model fiber orientation, the closure approx-
imation problem has been investigated, and the theory of the natural closure ap-
proximation has been thoroughly developed. The second objective of the chapter
was to review the models and numerical algorithms that are available to predict
the fiber orientation distribution in injection and compression molded parts. The
particular techniques implemented in the MOLDSYS software in order to per-
form decoupled orientation predictions have been detailed (the flow problem is
addressed in a companion chapter). Several examples, including successful
experimental validations, demonstrate the quality of the models and numerical
methods that have been elaborated
Besides the authors of this chapter, several people participated in this research,
including Marcel Crochet, Alain Couniot, Laurent Dewez, Hubert Henry de Fra-
han, Bernard Languillier, Gregory Lielens, Pascal Pirotte and Luc Vanderschuren,
whom the authors wish to thank here for their contributions. The work was
carded out within the framework of the "Multimat6riaux" project of the Walloon
Region of Belgium, the program of Intenmiversity Attraction Poles initiated by
the Belgian state and the COST 512 European project, and in collaboration with
the Shell Research and Technology Center in Amsterdam (the Netherlands) and
the Owens Coming company in Battice (Belgium). Grants from the IRSIA
(Belgium) are acknowledged. The authors wish to thank Natasha Van Rutten for
the drawing of several figures, and Jacques G6rard, Emile Franck and Jacques
Bleyfuesz for their help and advice in performing the molding experiments. The
efficient page setting work of Victor Vermeulen was also appreciated.

REFERENCES

1. W.C. Jackson, F. Folgar and C.L. Tucker 1II, in C.D. Han (ed.), Polymer
Blends and Composites in Multiphase Systems, American Chemical Soc.,
Washington (1984) 279.
2. C.L. Tucker lJ/, J. Non-Newtonian Fluid Mech., 39 (1991) 239.
3. C.L. Tucker Ill and S.G. Advani, in : S.G. Advani (ed.), Flow and Rheology
in Polymer Composites Manufacturing, Elsevier, Amsterdam (1994) 147.
4. V. Verleye, Ph.D. Thesis, Universit6 catholique de Louvain, 1995.
5. V. Verleye, G. Lielens, P. Pirotte, F. Dupret and R. Keunings, Proc.
Numiform'95, Balkema, Rotterdam (1995) 1213.
6. A. Couniot, L. Dewez, F. Dupret, R. Keunings, G. Lielens, P. Pirotte, J.
Bland, A.G. Gibson, G. Kotsikos, R. Gebart, J. Krispinsson, F. Vahlund, S.
Toll, J.-A.E. M~ason and C. Servais, Proc. MMSP'96 General COST 512
Workshop, Davos, Switzerland, European Commission (1996) 99.
1395

7. G. Lielens, P. Pirotte, A. Couniot, F. Dupret and R. Keunings, Composites


Part A, 29A (1998) 63.
8. M.C. Altan, J. Thermoplastic Composite Mat., 3 (1990) 275.
9. M.J. Crochet, F. Dupret and V. Verleye, in : S.G. Advani (ed.), Flow and
Rheology in Polymer Composites Manufacturing, Elsevier, Amsterdam
(1994) 415.
10. G.B. Jeffery, Proc. Roy. Soc., A102 (1922) 161.
11. F.P. Bretherton, J. Fluid Mech., 14 (1962) 284.
12. H. Brenner, Int. J. Multiphase Flow, 1 (1974) 195.
13. M. Vincent and J.F. Agassant, Rheol. Acta, 24 (1985) 603.
14. J. Feng and D.D. Joseph, J. Fluid Mech., 324 (1996) 199.
15. Y. Iso, D.L. Koch and C1. Cohen, J. Non-Newtonian Fluid Mech., 62 (1996)
115; 135.
16. S. Yamamoto and T. Matsuoka, J. Chem. Phys., 98 (1993) 644; 100 (1994)
3317.
17. S. Yamamoto and T. Matsuoka, Polym. Eng. Sci., 35 (1995) 1022; 36 (1996)
2396.
18. A.B. Metzner, J. Rheol., 29 (1985) 739.
19. G.K. Batchelor, J. Fluid Mech., 46 (1971) 813.
20. F. Folgar and C.L. Tucker III, J. Reinforced Plastics and Composites, 3
(1984) 98.
21. S.M. Dinh and R.C. Armstrong, J. Rheol., 28 (1984) 207.
22. E.S.G. Shaqfeh and D.L. Koch, Phys. Fluids A, 2 (1990) 1077.
23. D.L. Koch and E.S.G. Shaqfeh, Phys. Fluids, 2 (1990) 2093.
24. D.L. Koch, Phys. Fluids, 7 (1995) 2086.
25. G.K. Batchelor, J. Fluid Mech., 41 (1970) 545.
26. E.J. Hinch and L.G. Leal, J. Fluid Mech., 71 (1975) 481; 76 (1976) 187.
27. G.G. Lipscomb II, M.M. Denn, D.U. Hur and D.V. Boger, J. Non-Newtonian
Fluid Mech., 26 (1988) 297.
28. G.H. Fredrickson and E.S.G. Shaqfeh, Phys.Fluids A, 1 (1989) 3; 2 (1990) 7.
29. M.B. Mackaplow, E.S.G. Shaqfeh and R.L. Schiek, Proc. R. Soc. Lond. A,
447 (1994) 77.
30. M.B. Mackaplow and E.S.G. Shaqfeh, J. Fluid Mech., 329 (1996) 155.
31. A.G. Gibson and G.A. Williamson, Polym. Eng. Sci., 25 (1985) 968,
32. A.G. Gibson, Composites, 20 (1989) 57.
33. A.N. McClelland and A.G. Gibson, Composites Manufact., 1 (1990) 15.
34. S. Ranganathan and S.G. Advani, J. Rheol., 35 (1991) 1499.
35. S. Ranganathan and S.G. Advani, J. Non-Newtonian Fluid Mech., 47 (1993)
107.
1396

36. S. Toll and J.-A.E. M~-ason, J. Rheol. 38 (1994) 985.


37. R.R. Sundararajakumar and D.L. Koch, J. Non-Newtonian Fluid Mech., 73
(1997) 205.
38. M.R. Barone and D.A. Caulk, J. Appl. Mech., 53 (1986) 361.
39. H. Henry de Frahan, V. Verleye, F. Dupret and M.J. Crochet, Polym. Eng.
Sci., 32 (1992) 254.
40. V. Verleye and F. Dupret, Proc. ASME'93, New Orleans, Louisiana, AMD-
Vol. 175 (1993) 139.
41. V. Verleye, A. Couniot and F. Dupret, Proc. ASME'94, Chicago, Illinois,
MD-Vol. 49, HTD-Vol. 283 (1994) 265.
42. F Dupret, V. Verleye and B. Languillier, Proc. ASME'97, Dallas, Texas,
FED-Vol. 243, MD-Vol. 78 (1997) 79.
43. S.G. Advani and C.L. Tucker 111,J. Rheol., 31 (1987) 751; 34 (1990) 367.
44. A.J. Szeri and L.G. Leal, J. Fluid Mech., 242 (1992) 549; 262 (1994) 171.
45. J.S. Cintra and C.L. Tucker III, J. Rheol., 39 (1995) 1095.
46. A.J. Szeri and D.J. Lin, J. Non-Newtonian Fluid Mech., 64 (1996) 43.
47. G.G. Lipscomb II, R. Keunings, G. Marucci and M.M. Denn, Proc. I X th Int.
Congress on Rheology, Acapulco, Mexico, Elsevier, 2 (1984) 497.
48. J. Rosenberg, M. Denn and R. Keunings, J. Non-Newtonian Fluid Mech., 37
(1990) 317.
49. S.T. Chung and T.H. Kwon, Polym. Eng. Sci., 35 (1995) 604.
50. S.T. Chung and T.H. Kwon, Polym. Composites, 17 (1996) 859.
51. J. Azaiez, R. Gu6nette and A. Ait-Kadi, J. Non-Newtonian Fluid Mech., 73
(1997) 289.
52. K.K Kabanemi, J.F. H6tu and A. Garcia-Rejon, Intern. Polym. Proc. 12
(1997) 182.
53. V. Verleye and F. Dupret, Proc. Numiform'92, Balkema, Rotterdam, (1992)
389.
54. V. Verleye, A. Couniot and F. Dupret, Proc. ICCM/9 (9th Int. Conf. on
Composite Materials), Madrid, Spain (1993) 642.
55. A. Couniot, V. Verleye and F. Dupret, Proc. FPCM'94 (Flow Processes in
Composite Materials), Galway, Ireland (1994) 227.
56. V. Verleye, A. Couniot and F. Dupret, Proc. CADCOMP'94 (CAD in
Composite Material Tech.), Southhampton, United Kingdom (1994) 303.
57. F. Dupret and V. Verleye, in preparation (1998).
58. L.G. Leal and E.J. Hinch, J. Fluid Mech., 46 (1971) 685.
59. M.S. Ingber and L.A. Mondy, J. Rheol., 38 (1994) 1829.
60. A.J. Szeri, W.J. Milliken and L.G. Leal, J. Fluid Mech., 237 (1992) 33.
61. G.K. Batchelor, J. Fluid Mech., 44 (1970) 419.
1397

62. R.C. Givler, in M.M. Chen, J. Mazumder and C.L. Tucker (eds.), Transport
Phenomena in Materials Processing, ASME, New York (1983) 99.
63. R.C. Givler, M.J. Crochet and R.B. Pipes, J. Composite Mat., 17 (1983) 330.
64. E.H. MacMillan, J. Rheol., 33 (1989) 1071.
65. B.N. Rao, S. Akbar and M.C. Altan, J. Thermoplastic Composite Mat., 4
(1991) 311.
66. Y. Yamane, Y. Kaneda and M. Doi, J. Physical Soc. Japan, 64 (1995) 3265.
67. K.A. Ericsson, S. Toll and J.-A.E. M~nson, J. Rheol. 41 (1997) 491.
68. C.A. Stover, D.L. Koch and C1. Cohen, J. Fluid Mech., 238 (1992) 277.
69. E.J. Hinch and L.G. Leal, J. Fluid Mech., 52 (1972) 683; 57 (1973) 753.
70. M. Doi, J. Polymer Sci. : Polymer Physics Ed., 19 (1981) 229.
71. M. Rahnama, D.L. Koch, Y. Iso and C1. Cohen, Phys. Fluids, 5 (1993) 849.
72. M. Rahnama, D.L. Koch and E.S.Go Shaqfeh, Phys.Fluids, 7 (1995) 487.
73. M.C. Altan, S.G. Advani, S.I. Giigeri and R.B. Pipes, J. Rheol., 33 (1989)
1129.
74. W.C. Jackson, S.G. Advani and C.L. Tucker, J. Composite Materials, 20
(1986) 539.
75. E.S.G. Shaqfeh, Phys. Fluids, 31 (1988) 2405.
76. I.L. Claeys & J.F. Brady, J. Fluid Mech., 251 (1993) 411; 443; 479.
77. T.C. Papanastasiou and A.N. Alexandrou, J. Non-Newtonian Fluid Mech., 25
(1987) 313.
78. C.L. Tucker III, in : T.G. Gutowski (ed.), The Manufacturing Science of
Composites, ASME, New York (1988) 95.
79. S.G. Advani and C.L. Tucker, Polym. Composites, 11 (1990) 164.
80. M.C. Altan, S. Subbiah, S.I. Gtigeri and R.B. Pipes, Polym. Eng. Sci., 30
(1990) 848.
81. R.S. Bay and C.L. Tucker III, Polym. Composites, 13 (1992) 317; 332.
82. M.C. Altan and B.N. Rao, Proc. ASME'92, Anaheim, California, AMD-Vol.
153, FED-Vol. 141 (1992) 93.
83. M.C. Altan, S.I. Giigeri and R.B. Pipes, J. Non-Newtonian Fluid Mech., 42
(1992) 65.
84. M.C. Altan and L. Tang, Rheol. Acta, 32 (1993) 227.
85. L.G. Reifschneider and H.U. Akay, Polym. Composites, 15 (1994) 261.
86. L. Tang and M.C. Altan, J. Non-Newtonian Fluid Mech., 56 (1995) 183.
87. M.C. Altan and B.N. Rao, J. Rheol., 39 (1995) 581.
88. B.E. VerWeyst, C.L. Tucker m and P.H. Foss, Intern. Polym. Proc. 12
(1997) 238.
89. E.T. Onat and F.A. Leckie, J. Appl. Mech., 55 (1988) 1.
90. C.V. Chaubal and L.G. Leal, J. Rheol., 42 (1998) 177.
1398

91. S. Prager, Trans. Soc. Rheology, 1 (1957) 53.


92. G. Ausias, J.F. Agassant, M. Vincent, P.G. Lafleur, P.A. Lavoie and P.J.
Carreau, J. Rheol., 36 (1992) 525.
93. G.L. Hand, J. Fluid Mech., 13 (1962) 33-46.
94. B.N. Rao, L. Tang and M.C. Altan, J. Rheol., 38 (1994) 1335.
95. M.A. Bibbo, S.M. Dinh and R.C. Armstrong, J. Rheol., 29 (1985) 905.
96. N. Phan-Thien and A.L. Graham, Rheol. Acta, 30 (1991 ) 44.
97. R.B. Pipes, J.W.S. Hearle, A.J. Beaussart, A.M. Sastry and R.K. Okine, J.
Composite Materials, 25 (1991) 1204; 1379.
98. A. Ramazani, A. Ait-Kadi and M. Grmela, J. Non-Newtonian Fluid Mech.,
73 (1997) 241.
99. C.V. Chaubal, L.G. Leal and G.H. Fredrickson, J. Rheol., 39 (1995) 73.
100. P.K. Currie, J. Non-Newtonian Fluid Mech., 11 (1982) 53.
101. A.J. Szeri and L.G. Leal, J. Fluid Mech., 250 (1993) 143.
102. R. Blanc, S. Philipon, M. Vincent and J.F. Agassant, Intern. Polym. Proc., 2
(1987) 21.
103. T. Matsuoka, J.-I. Takabatake, Y. Inoue and H. Takahashi, Polym. Eng.
Sci., 30 (1990) 957.
104. B. Friedrichs, S.I. Gii~eri, S. Subbiah and M.C. Altan, J. Mat. Process. &
Manufact. Sci., 1 (1993) 331.
105. S. Toll & P.-O. Andersson, Polym Composites, 14 (1993) 116.
106. L. Schwartz, "Th6orie des distributions", Hermann, Paris (1966).
107. F. Dupret and L. Vanderschuren, AIChE Journal, 34 (1988) 1959.
108. M. Van Dyke, Applied Mathematics and Mechanics, Vol. 8, Perturbation
Methods in Fluid Mechanics, Academic Press, New York, 1964.
109. R.B. Davis, M.Sc. Thesis, University of Illinois, Urbana, 1988.
110. S. Kenig, Polymer Composites, 7 (1986) 50.
111. P.F. Bright, R.J. Crowson and M.J. Folkes, J. Mat. Sci., 13 (1978) 2497.
112. G. Lielens, Ph.D. Thesis, Universit6 catholique de Louvain, 1998.
1399

R E C E N T A D V A N C E S IN T H E R H E O L O G Y OF FLUIDIZED
MATERIALS

S. I. Bakhtiyarov and R. A. Overfelt

Space Power Institute, 231 Leach Center,


Auburn University, Auburn, AL 36849-5320, USA

1. INTRODUCTION

Flows of particulate materials are encountered in numerous commercial


industries including metallurgical, chemical, agricultural, pharmaceutical,
plastics, and food processing. Extensive applied research in these areas over the
years has developed a large database of empirical correlations that are used by
designers to engineer practical reactors and devices. However, the empirical
nature of the correlations often limits the applicability of the data to specific
situations.
The successful application of computational techniques to accurately simulate
single-phase flows in complex flow geometries makes their application to multi-
phase flows also highly desirable. Although rigorous two-phase flow
algorithms have been demonstrated for simple geometries, excessive
computational times preclude their widespread application to complex industrial
problems. Single phase algorithms must be modified to enable routine design
calculations. Central to the successful exploitation of single phase codes is the
measurement of appropriate transport properties (e. g., viscosity) that represent
the complex rheological behavior of these multi-phase fluids.
In this review we consider the works related to achieving an understanding of
the hydrodynamics of aggregative fluidization (gas-solid beds) with special
emphasis on rheological aspects of the problem. The review part of this work
covers the literature that was published in 1985 or later (also unpublished) and a
few earlier materials concerning the rheology of fluidized beds which were out
of the scope of previous major review works [ 1, 2].
1400

2. FLOW REGIMES IN FLUIDIZED BEDS

In both particulate (fluidization with a liquid) and aggregative fluidizations


there are several hydrodynamic flow regimes primarily related to the
characteristics of the solid phase, the fluid phase or the design of the equipment.
To understand the flow regimes in aggregative fluidization Grace [3] applied
the analogy with gas-liquid two phase flow. The following flow regimes have
been identified
9 delayed bubbling regime,
9 bubbling regime,
9 slugging regime,
9 turbulent regime,
9 fast fluidization, and
9 pneumatic conveying.
Bi and Grace [4] presented flow regime maps for gas-solid fluidized beds
(fixed bed, bubbling regime, slugging regime and turbulent regime) and gas-
solids upward transport (dilute-phase flow, fast fluidization or turbulent flow,
slug/bubbly flow, bubble-free dense-phase flow and packed bed flow). The
flow regime maps are shown with Re / Ar u3 plotted against Ar u3. Practical flow
regime maps are presented as ~ plotted against Qp [ pp.
In delayed bubbling regime (fixed bed) the particles contained in the bed are
motionless and are supported by contact with each other. The pressure drop -
flowrate diagram obtained by Bakhtiyarov and Overfelt [5] for fluidization of
silica sand by air is a good indication of flow regimes in aggregative
fluidization (Figure 1). At low air flow rates the pressure drop increases with
air flow rate. This branch of the curve pertains to air flow through the fixed
bed. At a certain air flow rate the pressure drop decreases rather suddenly,
which is obviously essential for predicting the onset of bubbling regime. The
value of the pressure drop at bubbling onset significantly depends on the initial
fixed bed height in the fluidization chamber. As the air velocity continues to
increase the pressure drop increases slowly but cannot reach the pressure drop
at onset of channeling. Qualitatively these curves are in good agreement with
the data of earlier experiments [6]. There are several approaches to predict the
superficial velocity at minimum fluidization; however, most equations are of the
form [7]

Remf = (C12 + C2 Ar) 0"5 - C l , (1)

where Remf = pg dp Umf / ~g and Ar = pg (pp - pg) g dp3 ] ~j,g2 .


1401

Most experimental studies of bubbling phenomena are interested in bubble


size, velocity, shape and flow patterns. Jackson [8] concluded that the theory of
stability of uniform fluidized beds reveals the basic instability without which
the process of bubble growth cannot begin. Many researcher confirmed that the
bubbles in fluidized beds can be developed from a small perturbation.
Experimental data show that increasing pressure causes bubbles of smaller sizes
resulting in smoother fluidization [9, 10]. Based on their experimental data
with fluidized cracking catalyst and polypropylene powder, Piepers and
Rietema assume that the elasticity modulus is a significant property of the
homogeneous and heterogeneous fluidized beds [10]. Gautam, Jurewicz and
Kale [11] applied laser Doppler anemometry to non-intrusively measure the
through-flow component in single isolated bubbles in two-dimensional
fluidized beds. They observed that the bubble velocity increases with distance
from the air distributor and the through-flow inside the bubble (at a given height
above the air distributor) is constant. Recirculation zones of gas were not
observed inside the bubbles. Gera and Gautam [12] reported that the bubble
aspect ratio plays an important role in predicting an accurate gas flow through
the bubble. They also observed that the through-flow velocity component at the
nose of an elongated bubble is higher than that in a flattened bubble. Utilizing a
newly developed capacitance imaging system, Halow et al. [13] reported that
the aspect ratio of bubbles appears to increase linearly with bubble diameter up
to a bubble to tube diameter ratio of about 0.5, and remains constant up to a db/
d of 0.7 - 0.9, then increases rapidly as db --- d. Using X-rays, Yates, Cheesman
and Sergeev [ 14] examined air-fluidized beds (solid particles of A and B groups
of the Geldart classification) and found that the rising bubbles are surrounded
by an expanded "shell" of particles in which the void factor is significantly in
excess of that of the emulsion phase remote from the bubbles.
Based on Lavrentyev's turbulent-eddy flow model, Bakhtiyarov and Overfelt
[5] predicted the friction coefficient during the bubbling regime

= 12 c / Re [1 - ot2(3 - 2o0 + 6o~2(1 - o~) n/(n+l)], (2)

where c, n and ~ are dimensionless parameters depending on particle and gas


properties and the design of the fluidization chamber.
The experimental results, obtained employing combined capacitance-pressure
probes on the apparatus shown in Figure 2 with silica sand fluidized by air [5,
15], confirmed the predictions of equation (2) (Figure 3). As seen from the
comparison of analytically predicted and experimentally obtained data, there is
good agreement at a wide range of Reynolds number values.
1402

Figure 1. Variation of the pressure drop as a function of the average air velocity
and height of the fixed silica sand bed [5].

Slugging is a flow regime in which gas bubbles increase to the diameter of the
fluidization chamber. The slugs of solid particles will move upward in a
pistonlike manner, reach a certain height, and then rain through the gas phase in
the form of aggregates or as individual particles. In group A of Hovmand-
Davidson's [ 16] classification, air slugs rise through the bed of particles which
rain down through the void to allow its upward motion. In group B, the upward
movement of the interface is slow and is caused by particles raining down
uniformly through the disperse regions. A slug flow of group A is analogous to
the flow of the two-phase (gas-liquid) system. Therefore the group A slug flow
has been studied thoroughly. A slug flow of group B occurs at certain values of
the bed height to diameter ratio (H/D) and the superficial velocity of gas. To
predict the slug flow regime some investigators introduce an "equivalent bubble
diameter" which is about 1/3 - 1/2 of the bed diameter.
The effect of slugging regime on the energy loss in fluidized beds has been
investigated analytically and experimentally by Bakhtiyarov and Overfelt [15].
Based on Meshchersky's model of motion of a body having variable mass and
the B lake-Kozeny-Carman equation of porous media flow, the following
expression for friction coefficient as a function of the Reynolds number, void
factor, gas velocity, sizes of the slug and fluidization chamber has been
obtained
1403

Figure 2. Apparatus for study of hydrodynamics of fluidized bed [5, 15].

10000
0 h=7cm
o h=18.4cm
!_
A h=33.65cm
o - - " ' e q u a t i o n (2)
0
t~
lie
C" 1000
o
,m

~0
!._
lie

100
10 100
Reynolds Number

Figure 3. Friction factor as a function of Reynolds number in bubbling regime


of fluidized silica sand bed [5].

X = 321/{Re [1- [~4 -I- (1- 1~2)2 / In 13 + 8 e3 132/ 5 R z S 2 (1- e)2]}, (3)

where ~ is a design parameter of the fluidization chamber.


1404

Also experiments were run in the fluidization chamber (Figure 2) with silica
sand at different values of the fixed bed height, pressure and air flow rate [15].
The study of the slug motion dynamics demonstrated that both the drag
coefficient and the resistance factor decrease with increasing the Reynolds
number and the porosity of slug. At a given Reynolds number, the values of the
friction coefficient and the resistance factor are higher than for fluidized bed
with large value of the fixed bed height (Figure 4). There is a qualitative
agreement between the experimental data and predictions, with better agreement
at lower values of Reynolds number.
One of the major complications in the fluidization of materials is a tendency
towards radial segregation in the bed. Segregation is a state of the fluidized bed
in which a larger proportion of solid particles is found in one region than in
another one. In many experimental studies of gas-solid particles two-phase
flow the radial segregation phenomenon has been observed, where a fluidized
bed comprises a rapidly rising core with a low concentration of solids and an
annulus region near the column wall where the solids concentration is higher.
The research literature related to segregation phenomena and published before
1985 was reviewed by Nienow and Chiba [17]. Shape factor, density, moisture
content, size and size distribution of the solids are the main factors that affect
segregation tendencies.
Nienow, Naimer and Chiba, [18] concluded that the distributor design has a
significant effect on segregation tendencies. According to their results,
standpipe and perforated plate distributors provide better mixing than a porous
plate at the same superficial gas velocity. To predict the presence of
segregation during fluidized bed operations Delebarre, Pavinato and Leroy [19]
introduced an index based on pressure losses along the fluidized column.
Introducing suspended particle distribution coefficients, Hong and Tomita [20]
presented a model for high density gas-solid stratified pipe flow, in which the
particle-particle interactions between the suspension and the sliding bed were
taken into account. They found that particles begin to drop out of the gas phase
at the saltation point. The impact of the inlet configuration on the rate of
segregation of particles to the wall and the internal recirculation during the
steady, developing flow of gas-solid suspensions in a vertical tube were
investigated numerically by Pita and Sundaresan [21]. Examining the three
inlet configuration (uniform, core-annulus and circumferential), these authors
conclude that a circumferential injection of gas has a favorable effect on the
flow in the sense that it can decrease the extent of internal recirculation, and
hence, the segregation of particles to the wall. Dasgupta, Jackson and
Sundaresan [22] analyzed the time-smoothed equations for the motion of dense
1405

suspensions to demonstrate the impact of density fluctuations on the occurrence


of segregation. It was shown that the solid particles will congregate in regions
where the kinetic energy of fluctuations associated with the particles is small.
100

0 h=18.4cm

9 [] h=33.65cm
!....
o 10
.0 . . . h=18.4cm (eqn.3)
t~
tl. 9
~: h=33.65cm (eqn.3)
,,i.I
0
=
9 "~
,,_ 1

N
0.1 . . . . . . . . ~ ~- . . . . . .
10 100 1000

Reynolds Number

Figure 4. Variation of friction factor with Reynolds number at slug flow regime
[15].

Rhodes, Zhou and Benkreira [23] and Bakhtiyarov and Overfelt [24]
developed semi-empirical models for radial segregation of solid particles in a
circular tube. According to the models proposed, the radial distribution of
solids is dependent on the average solids concentration as

(1 - 8) / (1 - 8w) = C1 (r/R) n, (4)

(1 - ~) / (1 - ~av) = C2 (r/R) n, (5)

where C~, C 2 and n are constants (C~ = 1, [14, 15]; C2 = 2, [14, 15]; n = 2, [14];
n = 3, [15]).
In Figures 5 and 6 the radial solids concentration profiles predicted by
theoretical analysis [23, 24] are compared with experimental data of different
researchers [24-29]. As seen from these figures, there are qualitative and
quantitative agreements between the experimental data and the theoretical
predictions.
Spouting is a state of fluidized bed in which a fluid jet at high rate pierces
through the bed and a central channel is formed. The spouted bed is limited to
relatively large solid particles and mainly is affected by construction of the
1406

9 U=3.8m/s [27]
A
I
2.5
.= [] U=4.0m/s [28]
~o
!
2 9 U=3.3m/s [26]
A U=O.2m/s [25]
1.5
A
0 U=1.3m/s [24]
r! 1 0 U=l.7m/s [24]
qlI
[] U=2.3m/s [24]
~" 0.5
"~" --"theory [23]
theory [24]
0 0.2 0.4 0.6 0.8 1

rlR

Figure 5. Radial profile of solids concentration.

fluidization chamber. Bakhtiyarov and Overfelt [30] reported the results of the
theoretical and experimental analyses of the spouted bed concerning the
influence of the design parameters of the fluidization chamber on the
hydrodynamic characteristics of the fluidized bed. The friction factor was
expressed as a function of Reynolds number, as following

~, = 8 13/ k Re, (6)

where ~ is a design parameter of fluidization chamber and k is a function of the


permeability of the spouted bed.
The experimental data obtained with silica sand on the apparatus described in
Figure 2 compare well with the analytical predictions (Figure 7).
In the past years experimental and theoretical studies of the gas-solid turbulent
flow have attracted considerable interest. Gidaspow et al. [31] and Miller and
Gidaspow [32] measured the particle average velocity and concentration of gas-
solid flow in a vertical tube. It was shown that the main differences between
dilute and dense gas-solid turbulent flow are the mechanism of exchange of
momentum and fluctuation kinetic energy between the particulate and the fluid
phases. To establish a thermo-mechanical formulation for turbulent gas-solid
flows, Ahmadi and Ma [33] utilized a phasic mass-weighted averaging
technique. The velocity, void factor and fluctuation kinetic energy of different
phases were predicted. This model was used by Cao and Ahmadi [34] to
analyze the steady fully-developed, dilute and dense, gas-particle turbulent flow
between two parallel plates. The model predictions for the particulate and gas
phases were compared with the experimental data of Miller and Gidaspow [32]
and Tsuji, Morikawa and Shiomi [35] and good agreement was found.
1407

Introducing a new mean drag law for the solid particles, Singh and Joseph [36]
proposed one dimensional model of fluidized beds where the fluid and solid
momentum equations are decoupled. The linear stability analysis of the model
shows that uniform fluidization is unstable even when the force acting on a
solid particle is assumed to depend on the area fraction. The results of this
analysis agree with the experimental results.
A theoretical model for the fully developed flow of gas-solid suspensions in a
vertical tube was developed by Sinclair and Jackson [37]. Over the whole range
of co-current and counter-current flows, they predicted the relation between gas
pressure gradient and the flow rates of the two phases. A drag force between
the gas and the solid particles is a function of their relative velocity. The
existence of the mutual interactions between the solid particles through inelastic
collisions has been reported. The model has been expanded by Pita and
Sundaresan [38] for gas-solid suspensions flow in vertical tubes of different
diameters.
The flow of gas-solid suspensions in ducts of arbitrary inclination was
considered by Ocone,Sundaresan and Jackson [39].These authors conclude that,
as a result of the compaction due to gravity, it is necessary to take into account
forces transmitted between particles at points of sustained, rolling and sliding
contact.
Tu and Fletcher [40] reported the results of numerical computations and
comparisons with the LDV measurements of earlier experiments for turbulent
gas-solid particle flow in a 90 ~ square-sectioned bend. The comparison of the
mean velocity profiles of both the gas and particulate phases gave good
agreement with the LDV measurements. Authors predicted the localized high
particulate concentration near the outer curve of the bend that occurs at large
Stokes numbers.

3. R H E O L O G Y OF FLUIDIZED MATERIALS

There are several different experimental approaches to the study of the


transport properties of fluidized materials. The problem of the rheologist is the
interpretation of the flow behavior of a fluidized material in terms of its
physical and chemical properties and its state of fluidization. The first
measurements of an "apparent viscosity" of a fluidized material were reported
by Matheson, Herbst and Holt [41]. They used a modified Stormer viscometer
with a paddle 1.9 cm wide and 3.81 cm high. It was immersed in an aerated bed
of solids 7.62 cm deep. A viscosity was obtained at a sensor speed of 200 rpm
by comparing a liquid of known viscosity. They reported that the fluidized bed
1408

> 2.5 I xI
r
Q,)

i
2
qlm

1.5
A

I"

0.5
0
0 0.2 0.4 0.6 0.8 1
rlR

F i g u r e 6. Radial Profile of solids concentration. [29]" 9 - dp = 0.054 m m , pp =


930 k g / m 3, U = 3.58 m/s; II - dp = 0.075 mm, pp = 608 k g / m 3, U = 3.58 m/s; ~ -
dp = 0.043 mm, pp = 2003 k g / m 3, U = 3.58 m/s; A - dp= 0.054 m m , pp = 930
k g / m 3, U = 1.40 m/s; "ff - dp = 0.054 mm, pp = 930 k g / m 3, U = 3.15 m/s; 9 - dp =
0.054 m m , pp = 930 k g / m 3, U = 2.8 m/s; O - dp = 0.054 mm, pp = 930 k g / m 3, U
= 2.16 m/s; X - dp = 0.075 mm, pp = 608 k g / m 3, U = 2.16 m/s; [24]" dp = 0.32
mm, pp = 2593 k g / m 3, A - U = 1.25 m/s, ~ - U = 1.7 m/s, .I, - U = 2.25 m/s;
theories" m m [23], ~ [24].

1000

It.

100

10 i i * * . * 9 .

100 Reynolds Number 1000

F i g u r e 7. Variation of friction factor with R e y n o l d s n u m b e r at spouted flow


r e g i m e [30].
1409

was significantly less viscous as the gas flowrate was increased and particle size
was decreased. The addition of particles of large size had little effect on the
viscosity, but the addition of negligible amounts of fine particles significantly
lessened the viscosity. Trawinski [42] suggested an interesting model to
explain this phenomena. According to his model, fine particles act as "ball
bearings" between moving surfaces. Later, Kramers [43], Furukawa and Ohmae
[44], Shuster and Haas [45] and Grace [46] reported viscosity data also using
Stormer viscometers obtained with fluidized beds. The viscosity data actually
reported by a Stormer viscometer are the weights required to spin the immersed
paddle at a certain angular velocity. A different approach using a falling sphere
was applied by Peters and Schmidt [47] and Trawinski [42]. Applying the hole
of the liquid state, Trawinski [42] also analytically described the fluidized bed.
The following formula has been proposed to predict the effective viscosity of
the fluidized system

'1]a=[TIg+64(1-Emf)Pp(pp-pg)gdp 3] [U/(Umf E m f ) - K ( U [ U m f ) ~ 0"5 - 1 ].

Viscometers based on Hagen-Poiseuille laminar flow have also been used,


among the first by Siemes [48] and Siemes and Hellmer [49]. The main
advantages of this method are that (a) the flow process has no effect on the state
of the fluidized material and (b) the fluidized material flows in inclined
channels like a Newtonian liquid. However, Wu et al. [50] observed an
interesting effect in sand flow through orifices. At small particle diameters
(0.01 to 0.1 mm) the flow of solids is steady in the narrow range of the orifice
diameter to particle diameter ratio (2 to 12). At larger values of this ratio, the
oscillatory flow of sand through the orifice has been observed. Using a two-
phase fluid flow model based on kinetic theory for granular flow, Manger et al.
[51 ] developed a numerical model of the oscillatory flow of solid particles. The
simulations show qualitative agreement with the experimental data of Wu et al.
[50].
Based on the behavior of gas bubbles (shape, pressure drop, size, etc.) some
researchers attempted to predict the apparent viscosity of fluidized materials.
The method based on shape of bubbles in fluidized beds to predict the viscosity
was proposed by Grace [46], Davies and Taylor [52] and Mendelson [53]. It
was found that the bubble shape factor is a function of the Reynolds, Weber and
EtStvos numbers. Using the "included angle" (0) (angle of supporting surface
from horizontal) Grace [46] proposed the following empirical formula to
compute the apparent viscosity
Tla = 0.0435 dB UB pp (1 - a m f ) exp (0.004 0).
1410

Estimated values of the apparent viscosity were found to coincide with those
obtained in [46, 54].
Murray [55] predicted the apparent viscosity of a fluidized materials by
estimating the friction factor for rising bubbles. A method based on pressure
measurements as a bubble rose through a fluidized bed was reported by Stewart
[56]. Row and Partridge [57] estimated the apparent viscosity of a fluidized bed
through the bubble size measurements.
According to Sch~gerl, Merz and Fetting [54] only the viscosity data
measured by a torsional pendulum oscillating at low amplitudes or by Couette
viscometer at low angular velocities of the inner cylinder can be considered as
quantitative. Capillary tube viscometers are preferred when the data are to be
used for pipe flow problems, and rotational viscometers, which subject the
material under the test to a precise and uniform rate of shear, have definite
advantages in the analysis of complex systems [58].
Simple mathematical models cannot at present describe the general flow
behavior of fluidized materials. The apparent viscosity of the fluidized material
is a multiparametric function and is dependent on the physical and chemical
properties of both the solids and the aerating fluid. Fluidized materials are
particularly complex and, if numerical simulations of the behavior are to be
reliable, it is critical that the measured values be consistent regardless of the
measurement techniques applied.
Bakhtiyarov, Overfelt and Reddy [58] presented the results of the experimental
study of the "apparent viscosity" of fluidized silica sand utilizing both
Poiseuille flows (capillary viscometer) and Couette flows (rotational
viscometer). Viscometric measurements by capillary tube viscometers were run
in the system shown schematically in Figure 8.
The system consists of a clear acrylic cylindrical chamber 457.2 mm long and
69.85 mm inside diameter which is sealed at the top and bottom. A funnel with
an angle of approach 57030 , was attached at the bottom of the cylinder.
Precision-bore copper capillary tubes of 4.7625 mm and 7.9375 mm inside
diameters and five different lengths (L/D = 15.328, 30.656, 61.333, 122.667 and
245.334) were screwed into the funnel exit. Air at a carefully controlled
constant pressure was admitted through the top plug of the chamber. The rate
of flow of solid particles is typically measured by collecting a sample over a
measured time and determining its mass. The experiments were run using both
non-coated capillary robes and capillary tubes coated (to prevent wall slippage)
by sand of the same particle diameter of sample to be tested. Water and 50/50
glycerol/water mixture were used as calibrating liquids.
1411

One of the main difficulties in capillary tube viscometry is in accurately


determining the appropriate pressure drops. The corrections for the head of
sample over the tube, for kinetic energy effects, and for entrance losses are
required. The first correction is straightforward. The other two corrections can
be respectively estimated by (1) repeating experiments in capillary tubes of
different lengths and extrapolating the overall pressure drop to zero length and
(2) experimental calibration with Newtonian fluids of known viscosity and
density.

Figure 8. Capillary tube viscometer used in [58].

A computer controlled Brookfield HADV-II+ rotational viscometer was also


used in the experimental program [58]. This device measures the torque
required to rotate a spindle immersed in a fluid. For a given viscosity, the
viscous drag, or resistance to flow, is proportional to the spindle's speed of
rotation and is related to the spindle's geometry. Measurements made using the
same spindle at different speeds are used to detect and evaluate the rheological
properties of the test material. Viscosity measurements were made at spindle
angular velocities in the range of 1 to 100 rpm. The viscometer was calibrated
by using 50/50 and 99/1 glycerol/water mixtures and 3.5% polyacrylamide
solution. A round acrylic tube with internal diameter of 58 mm and 227 mm
length was used as a fluidization chamber (Figure 9). Pressure taps were
provided along the column. The fluidization chamber was arranged for use with
air as the operating fluid. Compressed, dried and pre-filtered air under the
pressure up to 840 kPa was supplied to the bottom of the fluidization chamber.
1412

Air pressure and flowrate were measured by computer controlled pressure


transducer and rotameter, respectively. A polypropylene porous plate with
0.250 mm pore size and 3.175 mm thick was used as an air distributor. All
measurements were made with silica sand of average particle density 2.593
g/cm 3. The distributions of particle diameters used were as follows: less than
0.212 mm, from 0.212 mm to 0.425 mm, and from 0.425 mm to 0.710 mm. To
prevent wall slippage effects, the surface of the spindle was coated by sand of
the same particle diameter of sample to be tested.

Figure 9. Apparatus for fluidized sand viscosity measurements by rotational


viscometer [58].
The apparent shear strain rate and the apparent shear stress at the capillary wall
are defined by

Ta=4 Q / p r ~ R 3 = 4 t ~ / R , (7)

Xa= AP R / 2 L . (8)

The apparent viscosity is defined by the ratio of Xa over ~/a"

Tla - 'l;a / ~/a 9 (9)

In the case of sand flow, the capillary viscometer technique determines the
apparent viscosity as the averaged value of all inner local viscosity. The value
of the apparent viscosity is obviously a function of the shear strain rate q(a-
1413

Figure 10 shows the results obtained using the Brookfield viscometer for silica
sand (0.212 mm < dp < 0.425 mm) at different bed void factors. As either the
shear rate or the bed void factor (air flow rate) increases, the apparent viscosity
decreases. The apparent viscosity is most sensitive to shear rate and voidage
effects near incipient fluidization. Thereafter, the rate of decrease lessens.
However, the measured viscosity was only slightly affected by changes in
particle diameter for the diameters investigated.
The apparent viscosity can be correlated with the apparent shear strain rate in
the usual way as [58]

'rla = K~~an-1 . (10)

where ~ is a measure of consistency of the fluidized bed. Lower void factor of


the fluidized material yields larger values of consistency, n is a measurement
of the fluidization behavior of the fluidized bed and is a function of the design
and operating characteristics of the fluidized bed. Fluidized beds can exhibit
behavior that is almost Newtonian (n = 1) or can display several anomalous
behaviors. The greater the divergence of n from unity, the more anomalous is
the fluidized bed.
Figure 11 relates the apparent viscosity of fluidized silica sand measured by
SchOgerl, Merz and Fetting [54] and Bakhtiyarov, Overfelt and Reddy [58].
From these data one can recognize that at low air flow rates (or low void factor)
the apparent viscosity increases.
Data obtained with the capillary viscometer [58] show that the mass flow rate
of solids does not depend on the overall pressure drop (column height) at
certain drained angles of repose. Gregory [59] also observed that in stick-and
slip flow of solid particles no external application of pressure is required to
push the solids through the pipe, the force exerted by gravity being all that is
demanded. For this type of flow he proposed the following formula

Qp = kl d z'5 , (11)

where kl is a coefficient which depends on a tube length, particle diameter,


shape factor, etc.
Rausch investigated gravity flow of a wide variety of solids in 76.2 mm and
203.2 mm diameter tubes [6]. He correlated all data by the following equation

Qp = f c Cw (d]dp) 2"7 g0.5 Pb dp 2"5 [ (tan 13)0.5 , (12)


1414

Figure 10. Apparent viscosity of fluidized sand bed as a function of the shear
rate and void factor [58].

10000 0
W
o 50 rpm [58]
a.
E [] 100 rpm [58]
[] 0
~, [ 5 4 ]
o
0 oo
o 1000 []
(/) /1
[] /1
[]
e-
G)

12,
<
100 ' ' ' ' ''"~ ' ' . . . . . . . . . . . . . . .
~o ~oo ~ooo ~oooo

Air Flowrate, cm31s


Figure 11. Apparent viscosity of fluidized sand as a function of air flowrate.
1415

where c is a correction factor pertaining to the angle of approach in the hopper


bottom; Cw is a wall-effect correction factor; dp = (d~ d2)~ , d~ and d2 are
adjacent sieve openings; f is a correction factor for tube length to diameter ratio,
introduced in [58]. Variation of the Gregory coefficient (k~) and the factor (f)
versus the capillary tube length to diameter ratio (L/d) are shown in Figure 12
[58]. Both correction factors appear to be approaching a limiting values as L/d
increases.
1.2

~ 0
.=
0.8
ll_
.=_o 0.6

0.4
...__.o...-.-=-.---
0
0.2

0 50 100 150 200 250


Lid

Figure 12. Correction factors as functions of the capillary tube length to


diameter ratio [58].

The apparent viscosity of sand particles (0.425 mm < dp < 0.710 mm) obtained
on both the rotational and the capillary viscometers as function of the apparent
shear rate is shown for comparison in Figure 13. Regardless of the method of
measurement, there is a general consistency in the data obtained with the
Brookfield viscometer (low shear rates) and capillary viscometer (high shear
rates). As seen from this figure, the apparent viscosity can be satisfactorily
correlated with the apparent shear rate by the following empirical power law
equation:

'qa = K~ ~r . (13)

A non-continuum approach to describe particle-gas suspension flow was


presented by Cundall and Strack [60]. The method was based on the use of a
numerical scheme in which the interactions of the particles are monitored
contact-by-contact and the motion of the particles simulated particle-by-particle.
Aizawa, Tamura and Kihara [61] calculated the particle-to-particle interaction
1416

forces by assuming that the interaction force between two solid particles is
equal to the overlapped volume force from the elastic contact between the
particles. To stabilize the numerical solution, they introduced a friction-like
energy-dissipation mechanism in the contact algorithm.

1.00E+07 O Brookfield [58]


w [] L/D=245.33 [58]
m o L/D=122.67 [58]
a,,
E 1.00E+06 ",o,, [] L/D=61.33 [58]
:r ',o,,
9 L/D=30.67 [58]
w
o 9 L/D=15.33 [58]
o 1.00E+05
w
,,,,= ~,,,,,,, -'-- - - E q n . (13)[58]
r.,
-9t~ 1.00E+04
o,,
D,
J a , , |||,=! 9 | J **1.,| | 9 9 , ~ , 1
1.00E+03 m

0.1 1 10 100 1000

Apparent Shear Rate, 1Is

Figure 13. Apparent viscosity of sand bed as a function of the apparent shear
rate [58].

Although these rigorous many-body algorithms have been demonstrated for


simple geometries, excessive computational times preclude their widespread
application to complex industrial problems. Continuum algorithms must be
modified to enable design calculations. Central to the successful exploitation of
such an approach is the measurement of appropriate transport properties (i.e.,
viscosity) that represent the complex rheological behavior of these multi-phase
fluids. Therefore, continuum theories have been developed to describe the
dynamics of particle-gas suspension flow. These theories develop
hydrodynamic models based on the Navier-Stokes equations for multiphase
flow. Using fluid mechanics principles, Anderson and Jackson [62] proposed a
two-phase flow model. Pritchett, Black and Grag [63] applied a two-fluid
model to solve the problem numerically. Gidaspow [64] presented the
hydrodynamic model describing gas and solid phases in terms of separated
conservation equations. Tsuo and Gidaspow [65] generalized the Navier-Stokes
equations for two fluids to predict flow regimes in circulating fluidized beds.
Using a two-phase 2-D computational fluid dynamics model, Samuelsberg and
1417

Hjertager [66] computed axial solid velocity, void factor and solid shear
viscosity in the riser of a circulating fluidized bed. The radial profile of solid
shear viscosity computed by the turbulent kinetic energy model was lower in the
core than that found experimentally, but with a linear function of solid volume
fraction in the measurement, the computed profile agrees well with experiments.
Delassade [67] considered the fluidized bed as a mixture of a micro-stretch fluid
representing the particulate phase, and a Newtonian fluid. The effect of particle
inertia and other consequences of non-negligible spatial gradients of velocities,
concentration and void factor on vertically propagating concentration waves of
small amplitude has been analyzed. Scht~gefl [68] carried out the flow equation
of fluidized systems from the shear diagrams using the method of Pawlowski
[69]. According to his results, the apparent viscosity of the particle-gas
suspension is independent of the flow rate at all particle diameters. However,
although the viscosity is independent of the particle diameter at particle
diameters dp > 0.100 mm, the viscosity decreases with decreasing particle
diameter at particle diameters dp < 0.100 mm. Reviewing a number of
expressions proposed for correlating fluidized bed viscosity with overall bed
void factor, Johnson [70] proposed that the fluidized bed viscosity is related to
bed voidage by

Tla "-'Fig [1 + 0.5 (1 - e)][(1 - ~) / ~]9 / ~4. (14)

Simulations made by equation (14) show a dramatic decrease in bed viscosity as


the void factor increases.
A kinetic theory based on a moment expansion was developed by Tsao and
Koch [71], who numerically predicted the behavior of a dilute fluidized beds
subjected to a simple shear flow. The results of simulations show that as the
Stokes number is increased, the velocity distribution approaches the
Maxwellian and the rheology becomes more Newtonian. At small Reynolds
numbers the non-linearity of the drag increases the non-Newtonian behavior of
the bed.
Hydrodynamics and erosion modeling of fluidized beds have been introduced
by Li and Zakkay [72]. Particle velocity fields and fluidized bed dynamics
(bubble formation and motion, bed expansion and collapse) were analyzed.
Utilizing the Finnie erosion model, these authors show that the distance from
the air distributor to the tubes, tube size, tube orientation and operating pressure
have significant effect on bed dynamics and tube erosion.
However, in order to accurately analyze the kinetic and transport processes in
these complex gas-particle systems, it is critical to determine a rheological
1418

model to describe their flow behavior. Since the dynamics of gas-solid


suspension flow is a multiparametric process as noted above, it is convenient to
develop the theological model on the basis of a semi-empirical correlation for
fluidized beds.
Bakhtiyarov and Overfelt [73] developed a rheological model of gas-solid
suspension flow based on Ergun's semi-empirical correlation [74]. Introducing
two components of the apparent viscosity, the model predicts that the flow
curves of gas-solid suspensions depend on the gas properties, particle diameter
and sphericity, void factor and tube diameter. The following form of the
theological model has been developed

q;a = [Tla.sh. + Tla.k.(~a) ] ~ a , (15)

where two components of the apparent viscosity as apparent shear viscosity


~a.sh. and apparent kinetic viscosity ~a.k.(~a), defined as

1]a.sh. = 75 (1 - I~) 2 ~l.gd 2 / 16 e 3 (0 dp)2 , (16)

Tla.k"---- 7 (1 - e) Og d 3 "Ya[ 1024 1~3 ~ dp, (17)

where a sphericity of particles r is defined as a ratio of the surface areas of a


sphere and the particle having identical volumes: t~ = S~ / Sp.
According to formula (15), the net apparent viscosity is the sum of the two
components: the first is due to shear viscous pressure drop and the second due
to the kinetic pressure losses. Therefore, it is of interest to establish the
contribution of each viscosity component to the total apparent viscosity. The
viscosity ratio ~ = 1"la.k./Tla.sh.as a function of the apparent shear rate and particle
diameter is shown in Figure 14. As seen from this figure, the viscosity ratio
increases as both the shear rate and the particle diameter increase.
Figures 15 and 16 show the variation of the apparent shear stress and the
apparent viscosity, respectively, with the apparent shear rate obtained
experimentally for silica sand of different particle diameters and over-all
voidage fraction measured by procedures described in [73]. The predictions of
the rheological model using equation (15) [73] are also shown in Figures 15 and
16. As seen from these figures, there is a good agreement between the model
and the experimental data.
Figures 15 and 16 show the variation of the apparent shear stress and the
apparent viscosity, respectively, with the apparent shear rate obtained
experimentally for silica sand of different particle diameters and over-all
1419

voidage fraction measured by procedures described in [73]. The predictions of


the rheological model using equation (15) [73] are also shown in Figures 15 and
16. As seen from these figures, there is a good agreement between the model
and the experimental data.

Figure 14. Viscosity ratio ~ = Tla.k.]Tla.sh.as a function of the apparent shear rate
and particle diameter [73].
An apparent viscosity model based on the integral characteristics of the radial
segregated fluidized beds has been developed by Bakhtiyarov and Overfelt [75].
Assuming that (i) the flow is laminar and steady on either side of the interface
between core and annular space, (ii) the interface is stable and smooth, (iii) the
pressure gradient is constant, (iv) end effects and a mass transfer between core
and annular flows are negligible, it has been determined that the apparent
viscosity of the segregated fluidized bed is a function of the viscosity and
density ratios of the core and annular flows. This embodies the effects of the
grain size and shape factor of particles, rheological characteristics and relative
velocity of the fluidizing fluid giving

l q s e g ~- [Xo3 q- p / 11 - Xo 3 (3 - p) / 211 - 3 Xo (p - 1) / 2111 / 11, (18)


1420

where 1"1 and p are viscosity and density ratios of the core and annular flows,
respectively; and Xo is a dimensionless radius of the core.

50

m
Ix. 40

W
G)
L_
30
L_
m
0
I
I
... 20 I
G}
t_
m
Q.
a. lO

0
500 1000 1500 2000 2500 3000
Apparent Shear Rate, 1/s

Figure 15. Apparent shear stress of air-sand suspension as a function of the


apparent shear rate [73]" r-! _ dp = 0.653 mm, ~ = 0.55; + - dp = 0.505 mm, ~ =
0.64; O - dp = 0.357 mm, ~ = 0.74; A - dp = 0.252 mm, e = 0.80; 9 - dp = 0.212
mm, e = 0.86.

Figure 17 shows the variation of the dimensionless apparent viscosity (1/Tlseg)


with the parameter rl, a ratio of the viscosities in core and annulus regions, and
with the dimensionless radius of the core (Xo). The radius of the core plays a
prevailing part in determining the apparent viscosity of the system. That is
especially conspicuous at high values of the parameter 1"1. The apparent
viscosity of segregated bed (1/rl~g) also shows strong subordination on the
parameter 1"1 which has an increasingly larger effect on 1/l"lseg with increasing
values of Xo.
Theoretical predictions for the dimensionless apparent viscosity (1/Tlseg) are
shown in Figure 18 together with experimental results obtained in [24] for silica
sand with grain size 0.212 mm to 0.425 mm and the average particle density
2.593 g/cm 3 fluidized by air. The apparent viscosities and average densities of
the segregated fluidized beds in the core and annulus regions have been
determined by using the procedure described in [24]. As seen from the figure,
1421

the apparent viscosity of the segregated fluidized bed is reasonably well


predicted at all values of the viscosity ratio rl.

16

w
== 12
o.
E

u 8 ,,~

-_ 9
13... " ~ -'~ .6 ~'' .'~;~ "

<
F.
I:_,. .- - , ' - " " "
ii ii ~ l " II ..... =_.,. . . . . . . . .
/
/

0 * ' * I * ' * I I I I I I I I I I ' i . , , ,

500 1000 1500 2000 2500 3000


Apparent Shear Rate, 1Is

Figure 16. Apparent viscosity of air-sand suspension as a function of the


apparent shear rate [73]: I'1 - d p = 0.653 mm, e = 0.55; + - d p - 0 . 5 0 5 m m , E =
0.64; O - d p = 0.357 mm, e = 0.74; A - d p = 0.252 mm, e = 0.80; @ - d p - 0.212
mm, ~ = 0.86.

The hydrodynamic behavior of fluidized beds in the presence of an external


physical fields such as magnetic, electric, acoustic or gravity are of both
methodological and practical interests. Sergeev and Dobritsyn [76] considered
the propagation of solid concentration disturbances in fluidized beds in an
external magnetic field. Using the Korteweg-de Vries-Burgers equation for the
departure of solid concentration from the uniform state, authors predicted the
structure of the wavefront and the thickness of the front as a function of
magnetic and other physical parameters of the bed.
A general approach for evaluating the sedimentation velocity and drag force of
charged particles in bounded system for small particle Peclet numbers and small
particle-surface potentials has been described by Pujar and Zydney [77]. The
viscous force that arises from the alteration in the velocity profiles associated
with the interaction between the electric field and the fluid was evaluated using
a generalized form of the Lorenz reciprocal theorem.
1422

Figure 17. Variation of the dimensionless apparent viscosity (1/~t) with the
parameter 1"!, a ratio of the viscosities in core and annulus regions, and with the
dimensionless radius of the core (x0) [75].

The effect of the microgravity on the "apparent viscosity" and onset of


fluidization of the fluidized solid particles, which is a two-phase flow system
free from surface tension, has not received much attention in the past mainly
due to the complex rheological behavior of this system. Bakhfiyarov and
Overfelt [78] first reported the results of the low gravity experiments developed
and conducted on board the NASA/KC-135 aircraft which were specifically
aimed to understand the effect of the gravitational forces on the "apparent
viscosity" of the fluidized materials.
Short (-- 20 s) duration of microgravity (-- 0.1 m/s 2) are created aboard
NASA's Zero-G KC-135 aircraft by a series of parabolic trajectories. Each
microgravity period followed a-- 50 s period of 1.8g gravity during the flight
maneuvers. The aircraft flew -- 40 parabolas on each flight which allowed for
repeating the tests several times and for changing the flow regimes. The KC-
135 aircraft climbs and descends between altitudes of 7.6 and 10.6 km. The
gravitational acceleration was measured by a computer operated capacitive
three axis accelerometer which enabled recording typical gravitational
1423

acceleration cycles in all three directions for each parabola during the flight
maneuvers of the KC- 135.

250

200

150
=I.

100

50

0
0 20 40 60 80 100

Figure 18. Comparison of the experimental data [24] and theory [75] for the
dimensionless apparent viscosity as a function of the parameter 11" x0 = 0.5, !"! -
u0= 1.25 m/s, A - u0= 1.70 m/s, O - u0= 2.25 m/s.

The experimental set up, presented in Figure 19, was specifically designed and
built with the safety and portability requirements of the flight program. A
computer controlled Brookfield HADV-II+ rotational viscometer was used in
both the ground based proof and the flight experimental programs.
The advantage of the measurements in reduced or increased gravity conditions
is the possibility to change the buoyant weight of the system at the same fixed
bed height and voidage. Figures 20 and 21 show the variations of the
expanded bed height to the fixed bed height ratio and the overall voidage,
respectively, with the air velocity at three different values of the gravitational
acceleration during the fluidization of the silica sand particles tested (dp < 0.212
mm and 0.600 mm < dp < 0.710 mm) where data pertained to the g = 9.81 m/s 2
were obtained in the ground based experiments [78]. As seen from these
figures, the particle diameter plays a predominant role in expansion of the bed
and the void factor. The bed expansion and the voidage also show strong
dependence on the gravitational acceleration which has an increasingly larger
effect on bed expansion and the void factor with increasing values of the air
velocity. Figure 22 shows the gravity effects on the air velocity required for
1424

the incipient fluidization for two solid particle sizes [78]. The experimental data
show that increasing the gravitational acceleration causes suppression of the
bed and increases the amount of gas necessary to initiate fluidization.

Figure 19. Schematic diagram of experimental apparatus designed for flight


experiments [78].

As we mentioned earlier, the "apparent viscosity" of fluidized systems strongly


depends on the overall void factor of the bed. Since the incipient fluidization
and the voidage are functions of the gravitational forces one would expect an
effect of the gravitational forces on the "apparent viscosity" of fluidized beds.
The results of the "apparent viscosity" measurements [78] for the fine (dp <
0.212 mm) and for the coarse (0.600 mm < dp < 0.710 mm) silica sand particles
at different spindle angular velocities and gravity levels are shown in Figures
23 and 24, respectively. As seen from these figures, a predominant influence of
the gravity level on the "apparent viscosity" is particularly noticeable at small
shear rates (angular velocities of the spindle). One is drawn to the inescapable
conclusion that at high shear rates the influence of the centrifugal forces and the
related phenomena (radial segregation, turbulence, etc.) on the measured
"apparent viscosity" become more significant. This conclusion can be
succinctly demonstrated in Figures 25 and 26 [78], which show the variation of
the "apparent viscosity" of the fluidized silica sand particles (dp < 0.212 mm
and 0.600 mm < d p < 0.710 mm, respectively) with imposed gravitational
1425

acceleration at a wide range of the spindle angular velocity changes (10 to 100
rpm).
Recently, experimental results obtained by Hunt, Hsiau and Hong [79] show
that the expansion of the bed increases significantly beyond a critical frequency
(10 Hz), and that the expansion does not depend on the amplitude of the
vibration. The effect of the vibrational amplitude may be significant if the ratio
of vibrational amplitude to particle diameter is significantly larger or smaller
than unity. Therefore, a more detailed study of the fluidized beds in
microgravity should include the effects of small oscillations in acceleration
under the reduced gravity conditions, called g-jitter.

0.4
"o

m
._x 0.3
u.
O.o

rfl t m

~
"o 0.I
/,.;" I
m

I A I

-031 0.1 1 10
Air Velocity, m l s

Figure 20. Variation of the expanded bed height to the fixed bed height ratio
with imposed air velocity (dp< 0.212 mm: 9 - 9.81 m/s 2, I - 11.772 m/s 2, A -
17.658 m/s2; 0.600 mm < dp< 0.710 ram: -~- - 9.81 m/s 2, r-1 _ 11.772 m/s 2, A -
17.658 m/s 2) [781.

4. CONCLUSIONS

Hydrodynamics of gas-solid fluidized beds cuts across many disciplines and is


of significance to many industrial processes. During the last decade substantial
progress has been made in this area. Appealing fundamental and complex
practical problems have challenged scientists and engineers to create new
theoretical models and unique experimental techniques for better understanding
1426

0.56

9 /

0.52 I,,l I:t


0
01
0.48
o
>

0.44
..;"
0.4 . 9 9 9 I'''l I l 9 ..,,.l . . . 9 ,.,.

0.01 0.1 1 10

Air Velocity, m/s

Figure 21 Variation of the void factor with imposed air velocity [78] (dp<
0.212 mm: 9 - 9.81 m/s 2, i - 11.772 m/s 2, A - 17.658 m/s2; 0.600 m m < dp<
0.710 mm: + - 9.81 rn/s 2, r-I _ 11.772 m/s 2, A - 17.658 m/s2).

0.8
o

0.6
E

0
~ D4
O
_c 0.4 -~
m ! i

8 12 16 20

Gravitational Acceleration, m/s 2

Figure 22. Variation of the incipient fluidization air velocity with imposed
gravitational acceleration ( i - dp<0.212 mm; r'i _ 0.60 mm< dp< 0.71 mm) [78].
1427

250

m 200

., 0o
"ILZ~:"
~ 100 I~ XT ~ ~"'~o " ~ ' " - I

< 50
E 9I , . ' ' ~ - . . " ~-.

"

0
0 20 40 60 80 100
Spindle Angular Velocity, RPM

Figure 23. Variation of the "apparent viscosity" of fluidized silica sand


particles with imposed spindle angular velocity (dp< 0.212 mm: -~ - 0.981 rn/s 2,
II - 9.810 m/s 2, A - 11.772 m/s 2, O - 17.658 m/s 2) [78].

600
w
te 500
Ix
>;
ul 400
o
u
.w_ 300
> Vi/x i
r 200 i II

D. ~Q
o. 100 "lOm D~
< _ -~" :." ~.-.:.-..~
0 20 40 60 80 100
Spindle Angular Velocity, RPM

Figure 24. Variation of the "apparent viscosity" of fluidized silica sand


particles with imposed spindle angular velocity (0.600 mm < dp< 0.710 mm: -~ -
0.981 m/s 2, II - 9.810 m/s 2, A - 11.772 m/s 2, O - 17.658 m/s 2) [78].
1428

250

ca 200
a.

100

L ~ ~ ~ li.~- - "~" " " ""


,~ so "'''2"'-'''-

0
0 5 10 15 20

G r a v i t a t i o n a l A c c e l e r a t i o n , nl/S 2

Figure 25. V a r i a t i o n o f the " a p p a r e n t v i s c o s i t y " o f f l u i d i z e d silica sand


p a r t i c l e s w i t h i m p o s e d g r a v i t a t i o n a l a c c e l e r a t i o n (dp< 0 . 2 1 2 m m : + - co = 10
r p m , r-I _ ~ = 30 r p m , A - o~ = 6 0 r p m , O - co = 100 r p m ) [78].

600

ca 500
(g
a.
_~ 400
(n
o
o
300

r~
9 2oo
D,,
,< 100
---I ..... x .....

0 5 10 15 20

Gravitational Acceleration, m/s 2

Figure 26. V a r i a t i o n o f the " a p p a r e n t v i s c o s i t y " o f f l u i d i z e d silica sand


particles with imposed gravitational acceleration (0.600 mm < dp< 0 . 7 1 0 m m : +
- CO= 10 r p m , r-I _ co = 12 r p m , A - o~ = 20 r p m , I - co = 30 r p m , 9 - o~ = 5 0 r p m ,
O - co = 6 0 r p m , & - t~ = 100 r p m ) [78].
1429

of fluidization phenomena. The heterogeneity of the fluidization still remains a


daunting feature of gas-solid systems. The present challenge is to understand
the physical mechanisms of heterogeneity and to find effective ways to prevent
this undesirable phenomenon. It is hoped that in the next years efforts will be
made to advance the computer-based hydrodynamic and rheological models of
fluidization. Systematic study is needed to establish the influence of external
physical fields (e. g. magnetic, electrical, acoustic, gravity, centrifugal) applied
individually or in combination on behavior and rheological parameters of
fluidization.

REFERENCES

1. J. F. Davidson, R. Clift and D. Harrison, Fluidization, 2 nd edn., Academic


Press, London, 1985.
2. D. Geldart, Gas Fluidization Technology, Wiley, Chichester, UK, 1986.
3. J. R. Grace, Can. J. Chem. Eng., 64 (1986) 353.
4. H. T. Bi and J. R. Grace, Int. J. Multiphase Flow, 21 (1995) 1229.
5. S. I. Bakhtiyarov and R. A. Overfelt, 31 st ASME National Heat Transfer
Conference Proceedings, 1 (1996) 239.
6. M. Leva, Fluidization, Mc. Graw-Hill Book Co., New York, 1959.
7. K. S. Lim, J. X. Zhu and J. R. Grace, Int. J. Multiphase Flow, 21 (1995) 141.
8. R. Jackson, A. I. Ch.E. Symposium Series, 90 (1994) 301.
9. M. C~rsky, M. Hartman, B. Ilyenko and K. E. Makhorin, Powder
Technology, 61 (1990) 251.
10. H. W. Piepers and K. Rietema, in Fluidization VI, Edited by Grace, J. R.,
Shemilt, L. W. and Bergougnou, M. A., Engineering Foundation, New York,
1989, 203.
11. M. Gautam, J. T. Jurewicz and S. R. Kale, J. Fluids Eng., 116 (1994) 605.
12. D. Gera and M. Gautam, J. Fluids Eng., 117 (1995) 319.
13. J. S. Halow, G. E. Fasching, P. Nicoletti and J. L. Spenik, Chem. Eng. Sci.,
48 (1993) 643.
14. J. G. Yates, D. J. Cheesman and Y. A. Sergeev, Chem. Eng. Sci., 49 (1994)
1885.
15. S. I. Bakhtiyarov and R. A. Overfelt, IEEE 31 st Intersociety Energy
Conversion Engineering Conference Proceedings, 2 (1996) 793.
16. S. Hovmand and J. F. Davidson, Fluidization, Edited by Davidson J. F. and
Harrison, D., Academic Press Publishing Corp., New York, 1971, 193.
17. A. W. Nienow and T. Chiba, Fluidization, Edited by J. F. Davidson, R. Clift
and D. Harrison, 2nd edn., Academic Press, London, 1985, 357.
1430

18. A. W. Nienow, N. S. Naimer and T. Chiba, Chem. Eng. Commun., 62


(1987) 53.
19. A. B. Delebarre, A. Pavinato and J. C. Leroy, Powder Technology, 80
(1994) 227.
20. J. Hong and Y. Tomita, Int. J. Multiphase Flow, 21 (1995) 649.
21. J. A. Pita and S. Sundaresan, A. I. Ch. E. Joumal, 39 (1993) 541.
22. S. Dasgupta, R. Jackson and S. Sundaresan, A. I. Ch. E. Joumal, 40 (1994)
215.
23. M. Rhodes, S. Zhou and H. Benkreira, A. I. Ch. E. Journal, 38 (1992) 1913.
24. S. I. Bakhtiyarov and R. A. Overfelt, Powder Technology, submitted (1997).
25. A. Berker and T. J. Tulig, Chem. Eng. Sci., 41 (1986) 821.
26. E. U. Hartge, D. Rensner and J. Werther, Circulating Fluidized Bed
Technology (II), Basu and Large, eds., Pergamon Press, 1988, 165.
27. B. Herb, K. Tuzla and J. C. Chen, Distribution of Solid Concentration in
CFB, in Fluidization VI, Edited by Grace, J. R., Shemilt, L. W. and
Bergougnou, M. A., Engineering Foundation, New York, 1989, 65.
28. H. Mineo, High Velocity Circulating Fluidized Beds, Ph. D. Thesis,
University of Tokyo, 1989.
29. Y. Tung, Z. Zhang, Z. Wang, X. Lui and Z. Qin, Eng. Chem. Met. (China),
10 (1989) 17.
30. S. I. Bakhtiyarov and R. A. Overfelt, Study of Spouted Flow in Fluidized
Bed, Journal of Powder Technology, submitted.
31. D. Gidaspow, R. Bezdaruah, A. Miller and U. Jayaswal, Dense Transport
and Fluidization of Solids in Gas or Liquid Using Kinetic Theory, Joint
DOE/NSF Workshop on Flow of Particulates and Fluids, Worcester, MA.
32. A. Miller and D. Gidaspow, A. I. Ch.E. J., 38 (1992) 1801.
33. G. Ahmadi and D. Ma, Int. J. Multiphase Flow, 16 (1990) 323.
34. J. Cao and G. Ahmadi, Int. J. Multiphase Flow, 21 (1995) 1203.
35. Y. Tsuji, Y. Morikawa and H. Shiomi, J. Fluid Mech., 139 (1984) 417.
36. P. Singh and D. D. Joseph, Int. J. Multiphase Flow, 21 (1995) 1.
37. J. L. Sinclair and R. Jackson, A. I. Ch. E. Journal, 35 (1989) 1473.
38. J. A. Pita and S. Sundaresan, A. I. Ch.E. Joumal, 37 (1991) 1009.
39. R. Ocone, S. Sundaresan and R. Jackson, A. I. Ch. E. Joumal, 39 (1993)
1261.
40. J. Y. Tu and C. A. J. Fletcher, A. I. Ch.E. Joumal, 41 (1995) 2187.
41. G. L. Matheson, W. A. Herbst and P. H. Holt, Ind. and Eng. Chem., 41
(1949) 1099.
42. H. Trawinski, Chemie. Ing. Techn., 25 (1953) 201.
43. H. Kramers, Chem. Eng. Sci., 1 (1951) 35
1431

44. I. Furukawa and T. Ohmae, Ind. and Eng. Chem., 50 (1958) 821.
45. W. W. Shuster and F. C. Haas, J. Chem. Eng. Data, 5 (1960) 525.
46. J. R. Grace, Can. Joum. Chem. Eng., 48 (1970) 30.
47. K. Peters and A. Schmidt, (3st. Chem. Ztg., 54 (1953) 253.
48. W. Siemes, Chem. Ing. Techn., 24 (1959) 82.
49. W. Siemes and L. Hellmer, Chem. Eng. Sci., 17 (1962) 555.
50. X. L. Wu, K. J. Maloy, A. Hansen, M. Ammi and D. B ideau, Phys. Rev.
Lett., 71 (1993) 1363.
51. E. Manger, T. Solberg, B. H. Hjertager and D. Vareide, Int. J. Multiphase
Flow, 21 (1995) 561.
52. R. M. Davies and Sir G. I. Taylor, Proc. Roy. Soc., A 200 (1950) 375.
53. H. D. Mendelson, A. I. Ch. E. J., 13 (1967) 250.
54. K. Scht~gefl, M. Merz and F. Fetting, Chem. Eng. Sci., 15 (1961) 1.
55. J. D. Murray, Rheologica Acta, 6 (1967) 27.
56. P. S. B. Stewart, Trans. Inst. Chem. Eng., London, 46 (1968) 80.
57. P. N. Rowe and B. A. Partridge, J. Fluid Mech., 23 (1965) 583.
58. S. I. Bakhtiyarov, R. A. Overfelt and S. Reddy, ASME International
Mechanical Engineering Congress, Proceedings, AMD-217 (1996) 243.
59. S. A. Gregory, J. Applied Chemistry (London), 2 (1952) 1.
60. P. A. Cundall and O. D. L. Strack, G6otechnique, 29 (1979) 47.
61. T. Aizawa, S. Tamura and J. Kihara, ASME Proceedings, PED-61 (1992)
31.
62. T. B. Anderson and R. Jackson, Ind. Eng. Chem. Fund., 6 (1967) 527.
63. J. W. Pritchett, T. P. Black and S. K. Grag, A. I. Ch.E. Joumal, 74 (1978)
134.
64. D. Gidaspow, App. Mech. Rev., 39 (1986) 1.
65. Y. P. Tsuo and D. Gidaspow, A. I. Ch. E. Joumal, 36 (1990) 885.
66. A. Samuelsberg and B. H. Hjertager, A. I. Ch. E. Joumal, 42 (1996) 1536.
67. X. A. Delassade, ASME Proceedings, AMD, 217 (1996) 251.
68. K. Sch~gefl, in "Fluidization", J. F. Davidson and D. Harrison, ed.,
Academic Press Publishing Co., London-New York (1971) 261.
69. J. Pawlowski, Kolloidzschr, 30 (1953) 129.
70. E. Johnson, Institute of Gas Engineers (London), 378 (1950) 179.
71. H.-K. Tsao and D. L. Koch, J. Fluid Mech., 296 (1995) 211.
72. C. Li and V. Zakkay, Trans. of ASME, 116 (1884) 746.
73. S. I. Bakhtiyarov and R. A. Overfelt, ASME FED SM, Proceedings,
Vancouver, Canada (1997) $245.
74. S. Ergun, Chem. Eng. Progress, 48 (1952) 89.
1432

75. S. I. Bakhtiyarov and R. A. Overfelt, Proceedings of Third International


Conference on Multiphase Flow, Lyon, France (submitted).
76. Yu. A. Sergeev and D. A. Dobritsyn, Int. J. Multiphase Flow, 21 (1995) 75.
77. N. S. Pujar and A. L. Zydney, A. I. Ch.E. Journal, 42 (1996) 2101.
78. S. I. Bakhtiyarov and R. A. Overfelt, IEEE 32 ~d Intersociety Energy
Conversion Engineering Conference Proceedings, 2 (1997) 1439.
79. M. L. Hunt, S. S. Hsiau and K. T. Hong, J. Fluids Engineering, 116 (1994)
785.

NOMENCLATURE

Ar Archimedes number
d diameter
g gravity
K expansion coefficient
L length
P pressure (AP is a pressure drop)
Q mass flow rate
R radius of fluidization chamber or tube
Re Reynolds number
S specific surface area (surface area per unit volume of medium)
U superficial velocity
13 angle of repose
e porosity
t~ sphericity of solid particles
), shear strain rate
consistency
~, friction factor
dynamic viscosity
1"1 structural viscosity
0 included angle
p density
x shear stress
velocity
viscosity ratio

Subscripts
a apparent
1433

av average value
b bulk
B bubble
g gas
k kinetic
mf minimum fluidization
P particle
S sphere
seg segregated fluidized bed
sh shear
W wall
1435

H E A T AND M A S S T R A N S F E R IN R H E O L O G I C A L L Y
COMPLEX SYSTEMS

R.P. Chhabra

Department of Chemical Engineering, hTdian hTstJtute q/Technology


Kanpur, h~dia 208016

1. INTRODUCTION

During the last four to five decades, considerable attention has been accorded to
the fluid mechanics of theologically complex materials and therefore significant
advances have been made in developing better insights into the underlying
physical processes. In contrast to this, the transport of heat and mass in non-
Newtonian materials have received much less attention. It is readily agreed that
most unit and processing operations encountered in the handling and processing
of non-Newtonian materials entail temperature and/or concentration gradients
within the fluid medium, thereby resulting in the net transport of heat (or mass)
from one region to another. For instance, in many industrial applications, process
streams need to be heated or cooled and a wide range of equipment may be
utilized for this purpose ,e.g., double pipe or tubular heat excl~angers or stirred
tanks fitted with cooling coils or steam jackets. Similarly, inter-phase mass
transfer between a non-Newtonian medium and a particulate please is frequently
encountered in fixed, fluidized bed and three phase reactors used to carry out a
range of polymerisation and biochemical reactions. Further examples are found
in devolatilization of thin films, de-gassing of molten polymers, drying and
aseptic processing of liquid food stuffs, concentration of fruit juices, oxygenation
of blood, etc. Solnetimes, heat is generated in the process itself, sucl~ as in
extrusion of polymers and foodstuffs. It may too be necessary to reduce the rate
of heat loss from a vessel or ensure that heat is removed at a sufficient rate in
equipment such as screw conveyors. In most applications, it is the rate of heat
transfer within the process equiplnent which is of plincipal interest, though with
thennally sensitive materials (such as food stuffs, fermentation broths, etc.), the
telnperature profiles must be known and maximum permissible temperatures must
1436

not be exceeded. Obviously, the rate of heat (or mass) transfer and the
temperature distribution in a given application are strongly dependent upon the
geometry of flow, kinematic conditions and the physical properties of the fluid.
Indeed this interplay between these factors is fi~rther accentuated in the case of
non-Newtonian fluids due to their non-linear flow behaviour. This chapter aims
to provide a state of the art review of the currently available body of infonnation
in this field.

2. SCOPE

In this chapler, consideration is given to lhe phenomena of heat and mass


transport in processes involving rheologically complex materials, especially with
regard to the implications of shear dependent viscosity and viscoelasticity of
materials. As suggested earlier, the highly non-linear and strongly coupled nature
of the field equations preclude the possibility of analytical results and therefore
usually numerical solutions are sought, even for as simple a situation as that of
laminar flow of a purely viscous fluid in a circular tube. In this chapter, the main
emphasis is on the presentation of the results and the underlying physical
processes rather lhan on the numerical techniques p e r se. Clearly, it is also not
feasible to include here all possible geometries encountered in engineering
applications. Instead this review mainly addresses the phenomena of heat and
mass transfer in internal (conduit) flows and in boundary layer flows, followed by
a short section on heat transfer in geometries of practical interest. First of all,
however, the thenno-physical properties of the commonly used non-Newtonian
materials will be described.

3. THERMO-PItYSICAL PROPERTIES

In addition to tile flow characteristics (viscous and viscoelastic), the other


important physical properties of non-Newtonian substances are thermal
conductivity, density, specific heat, surface tension, coefficient of thermal
expansion, solubility and molecular diffusion coefficients. While the first three
enter into virtually all heat transfer calculalions, surface tension exe~ls a strong
influence on boiling heat transfer and bubble dynamics in non-Newtonian fluids.
Likewise, the coefficient of thennal expansion is impo~lant in heat transfer by
free convection. Finally, the last two properties, namely, solubility and molecular
diffusivity play central roles in mass transfer processes.
Admittedly, only very limited measurements of physical properties have been
made, but for dilute and moderately concentrated aqueous solutions of commonly
used polymers including carboxymethyl cellulose (Hercules), polyethylene oxide
1437

(Dow), carbopol (Hercules), polyacrylamide (Dow), etc., density, specific Ileal,


thermal conductivity, coefficient of thermal expansion and surface tension differ
from tile values for water by no more than 5-10% [1-8]. The thennal conductivily
and molecular diffusivity may be expected to be shear rate dependen! for both of
these and the viscosity are dependent on structure. Although recent
measurements [9] on aqueous carbopol solutions confinn this expectatir :l for
thennal conductivity, the effect, however, is small [10]. For engineering design
calculations, there will be little enor in assuming that all the above physical
properties, except molecular diffusivity, of aqueous polymer solutions are equal
to the values for water at the same temperature. The available literature on the
values of the molecular diffusivity into non-Newtonian polymer solutions is much
more controversial and inconclusive, as revealed by the available reviews
[11,12]. For instance, from the data reported for gases and neutral solutes
diffusing in non-Newtonian polymer solutions, all one can conclude is that there
is some effect of polymer concentration on diffusivity. More authors report that
diffusivity increases with increasing polymer concentration (increasing
consistency), eg., diffusivity values as large as 2.5 times the value in the virgin
solvent (zero polymer concentration) have been reported for CO= diffusing in
solutions of cellulose polymers. The effect is not as great for solutions of
polymers having simpler structure with no branching, etc. All in all, the values
ranging from --- 10 -~= to ~ 10~ m=/s have been reported for CO2, SO=, C2H2 and
O= diffusing in aqueous solutions of poly ethlyene glycol (PEG), carbopol,
polyethylene oxide (PEO), hydroxyethyl cellulose (HEC), etc. and of polystyrene
in toluene. Similarly, the corresponding values for solid solutes inferred from the
rates of dissolution from inclined surfaces and rotating disks range fiom --- 3.4 x
10-12 to --- 2.5 x 10-11m2/s for 13-naphthol, benzoic acid and oxalic acid in aqueous
solutions of PEO and CMC. Little is known about the role of viscoelasticity on
diffusion coefficient [13]. Much confusion, however, exists regardi~.~ the
influence of shear rate on molecular diffusion. For instance, Wasan et al. [14]
used a wetted wall column for measuring the diffusivity of oxygen in a series of
polymer solutions and reported a rather steep rise in the value of diffusivity with
increasing shear rate. On the other hand, many other workers [15,16] reported
very weak dependence on shear rate. Likewise, most of the cunently available
data pertain to room or near room temperatures and hence little is known about
the temperature dependence of the diffusivity in polymer solutions.
Some data of these properties for polymer melts are also available [17-19].
Admittedly, some attempts have been made at developing semi-theoretical
expressions for the prediction of thennal conductivity [20], diffusion [21,22],
solubility [23], etc., these are not yet refined to the extent of being completely
predictive. Besides, the values of thenno-physical properties in these systems
1438

seem to be strongly influenced by the detailed molecular structure (molecular


weight distribution, method of preparation, etc.) and therefore extrapolation from
one system to another, even under nominally identical conditions, can lead to
siD~ificant errors.
For industrially important particulate slurries and pastes exhibiting strong non-
Newtonian behaviour, the thenno-physical properties (density, specific heat and
thennal conductivity) can deviate significantly from those of its constituents.
Early measurements [24] on the aqueous suspensions of powdered copper,
graphite, aluminium and glass beads suggest that both the density, p, and the heat
capacity, Cp, can be approximated by the weighted average of tl~e individual
constituents, i.e.,

Psus = ~Ps + (l - ~)PL (1)

-- + (] - )CpL (2)

where ~ is the volume fraction of the solids, and the subscripts L, s and sus refer
to the values for the liquid, the solid and the suspension respectively.
The thennal conductivity, k, of these systems, on the other hand, seems
generally to be well correlated by the following expression [24-26]:

1 + 0.5(ks/kL)-~b(l-(ks/kL) )
k sus = k L 1 + 0.5 (k S / k L ) ~+b ( 1 - (k S /kL) ) (3)

Thennal conductivities of suspensions up to 60% (by weight) in water and other


suspending media are well approximated by equation (3). It can readily be seen
that even for a suspension of highly conducting particles (ks/kL ~ ~), the thermal
conductivity of a suspension can be increased only by a fewfolds. Furtherlnore,
the corresponding increase in viscosity from such addition would more than offset
the effects of increase in thermal conductivity on the rate of heat transfer.
For suspensions of mixed size particles, the following expression due to
Bmggemann [27] is found to be satisfactory:

(ksus/ks)- 1 = (k~usl"~ (i- r (4)


(k /ks)-1 Lk J
1439

The scant experimental data [28] for suspensions (~ _<0.3) of alumina (0.5 - 0.8
l.tm) particles in a paraffin hydrocarbon are in line with the predictions of
equation (4). An exhaustive review on the thennal conductivity of structured
media including polymer solutions, filled and unfilled polymer melts, suspensions
and food stuffs has been published by Durra and Mashelkar [29].
Of all the physico-chemical properties, it is the rheology which shows the
strongest temperature dependence. For instance, the decrease is apparent
viscosity at a constant shear rate is well represented by the usual Arrhel~ius type
exponential expression; both the pre-exponential factor and the activation energy
are generally shear rate dependent. It is thus customary to denote the temperature
dependence using rheological constants such as the power law consistency index
and the flow behaviour index. It is now reasonably established that the power
law flow behaviour index, n, of suspensions, polymer melts and solutions is
nearly independent of temperature, at least over a 40-50~ temperature interval
whereas the consistency index follows the exponential dependence on
temperature, i.e.,

m = mo exp (E/RT) (5)

where mo and E are evaluated using experimental results in the temperature range
of interest. Similarly, for Bingham plastic fluids, both the plastic viscosity and
the yield stress decrease with temperature in a similar fashion but with different
values of the pre-exponential factors and the activation energies. Temperature
dependencies of the other rheological characteristics such as the first and second
nonnal stress differences, extensional viscosity, storage modulus, compliance,
etc., though studied less extensively, have been discussed by Ferry [30].

4. HEAT TRANSFER IN DUCT FLOWS

The study of flow and heat transfer to non-Newtonian fluids in ducts is of both
theoretical and practical importance, owing to its wide ranging applications in a
spectrum of industries including food, chemical and polymer processing. Though
process streams are heated or cooled in a wide variety of geometries, most
studies to date have employed circular tubes and occasionally triangular, square,
and elliptic tubes have also been used. Furthennore, owing to generally high
consistencies of polymeric and other non-Newtonian systems, laminar flow is
encountered much more frequently than in the case of Newtonian fluids. Hence,
the rates of heat transfer are inherently lower under these conditions than those
achievable under turbulent conditions. The rate of heat transfer is further
influenced by the type of boundary conditions imposed, e.g., constant
1440

temperature or constant heal flux. Additional complications arise from the


viscous dissipation effects which intensified the coupling between the momentum
and energy balance equations.Thus, many possibilities exist in studying heat
transfer with non-Newtonian fluids in duct flows. This section is primarily
concerned with the current state of the art on l~eat transfer to non-Newtonian
(purely viscous) fluids flowing in circular tubes and ducts of regular cross-
sections, though frequently reference is made to other important studies and
reviews available in the literature on this subject [1,2,31-34].
In addition to hydrodynamic boundary layers, lhermal boundary layers are also
present when a fluid entering a duct is at a temperature different fiom that of the
duct walls. Thus, a thennal entrance region exists, and the duct can be divided
into two regions: the thennal entrance region and the fully developed thermal
region. While the notion of fully developed flow is easy to visualise at points
remote from the entrance of tube, the fi~lly developed temperature profile needs
fi~rther elaboration. Following Kays [35], the tenn implies that there exists, under
certain conditions, a generalised temperature profile which is independent of the
axial coordinate, i.e.

0z T,,, - T b

where T = T(r, z) is the local fluid temperature, T,,. is tile duct wall temperature
and Tb is the bulk fluid or mixing cup temperature. Needless to say that the fully
developed temperature distribution implies fidly developed flow.
From engineering applications standpoint, the central problem in duct heat
transfer is to predict either the rate of heat transfer, q, fiom the duct walls to the
fluid for a known temperature difference, or the fluid-to-wall temperature
difference for a given rate of heat flow. In either case, it is convenient to
introduce a heat transfer coefficient, h:

q = h(Tw- Tb) (7)

From the scaling of the field equations and the relevant boundary conditions, it
can readily be shown that for flow in a tube, the dilnensionless heat transfer
coefficient is a fimction of the Reynolds and Prandtl numbers, i.e.,

hD
Nu = = f(Re, Pr) (8)
k
1441

where Nu is the Nusselt number, Re, the Reynolds number, and Pr is the Prandtl
number. Thus, the central objective of all theoretical and experimental
investigations is to establish this fimctional relationship.lt is convenienl to begin
with the discussion of heat transfer in circular tubes, followed by a similar
treatment for non-circular ducts.

5. H E A T T R A N S F E R IN C I R C U L A R T U B E S

5.1 L a m i n a r Flow
5. l. l Fully Developed and Constant Physical Properties Case
The first generation of analyses dealt with the thennally fiflly developed
conditions and for constant physical properties, with no viscous dissipation. For
the constant heat flux conditions at the tube wall, the asymptotic Nusselt number
for power law fluids is simply given by [36]:

Nu = 8(5n + l)(3n + 1) (9)


(31n 2 + 12n + 1)

It is not difficult to show that another form of tllis expression is possible as [37]"

3n + 1~ ~/3
Nu = 4.36 ~ , ~ j (1o)

In another analytical study, Skelland [38] extended the Newtonian solution


developed by Rohsenow and Choi [39] to power law liquids as

Nu = n+l / 1/8)+ -2 (11)


3n + 1 15n: T 8n + 1

Yet another version was presented by hunan [40]. All these expressions reduce
to file generally accepted value of Nu = 48/11 for n = 1 and for the plug flow
condition at n = 0, Nu = 8. Analogous results for the Bingham plastic fluids have
been presented by Grigull [37] and by Matsuhisa and Bird [41].

For tile case of constant wall temperature, the Nusselt number in the therlnally
fiflly developed region is given by:

2
Nu = /31 (12)
1442

where 13~ is the lowest eigen value for the boundary value problem governed by
the equation:

1 d dZ, l 3n+ 1 1
~- d---~ F,-~j+[3~ n + 1 -~,"JZ, : 0

with Zi (1)= 0 and Zi (0) = 0


where ~, = r/R. The resulting values for the Nusselt number for n = 1, 0.5,
0.333 as calculated by Lyche and Bird [42] are 3.657, 3.95 and 4.175
respectively.
The inclusion of viscous dissipation term in the energy equation produces a
dramatic effect on the heat transfer coefficient. Using an elegant analytical
scheme, Sestak and Charles [43] obtained a closed fonn expression for the
asymptotic Nusselt number for the constant wall flux condition:

-1
Nu = [ 3 InS + 43n2 + 13n + 1 n(6n + 2)" Br (14)
L8(5n + l)(3n + l)(n+ 1) 8n,,I

When the Brinkman number, Br = 0, equation (14) almost coincides with


equation (11). An inspection of equation (14) suggests that the negative values of
Brinkman number (i.e. heating) drive the value of Nu towards zero. Under these
conditions, the temperature difference, AT, between the fluid in the wall region
and the wall is reduced. For a specified value of n, there will be a value of the
Brinkman number for which AT = 0.

5.1.2 Effect of Temperature Dependent Viscosity


In contrast to these constant properties solutions, Joshi and Bergles [44]
employed a numerical scheme to account for the temperature dependence of the
power law viscosity (m = mo exp(-bY)) on the value of Nusselt number under
constant wall heat flux condition. Their results indicate that the Nusselt number is
nearly independent of the ratio (lUb/la,,.) but strongly dependent on the parameter
(Rb q,,./k) = ~. Their numerical results are well represented by the expression:
NU~p _ 1 + (0.124 - 0.0542n) W - (0.01013 - 0.0068n) ~2 (15)
Nu

where the subscript "vp" refers to the temperature dependent viscosity value.
Equation (15) is applicable for tit < 6. In qualitative tenns, the overall effect of
temperature-dependent power law consistency index is to enhance the rate of
heat transfer for heating and to decrease it in the case of cooling of a fluid.
1443

5.1.3 Thermally Developing Heat Tran,~fer


The next generation of solutions dealt with the analysis of heat transfer in the so
called thermal entrance region where the velocity profile is fiflly developed. The
physical properties are assumed to be temperature independent. For the constant
wall heat flux condition, the Nusselt number is given by:

Nuz = 1.41 5]/3Gz 1/3 Gz > 25 :t (16)

The conesponding expression for the case of constant wall temperature is given
by:

Nuz = 1.16 61/3 Gzl/3 G,. > 33 rt (17)

where 8 = ((3n + 1) / 4n).


It is interesting to note that the shear thinning effects are completely accounted
for via the factor 8. All the aforementioned treatments also neglect the axial
conduction term in the energy equation. A recent study [45] suggests this to be a
reasonable approximation provided Pe > 1000 for the constant wall temperature
condition and Pe > 100 for the constant heat flux condition.
The role of the non-unifonn inlet temperature profile for power law fluids has
been investigated by Tonini and Lemcoff [46]. They found that at high values of
the Graetz number, Gz, the lowest temperature gradient occurred at the wall and
hence the lowest Nusselt number corresponds to the unifonn initial temperature
profile while at small values of Gz, the reverse is true. Likewise, Faghri and
Welty [47] investigated the influence of circumferentially varying heat flux.
The thennally developing flow and heat transfer with the Bingham plastic model
has received much less attention. The earliest analysis for a constant wall
temperature and constant physical properties is due to Pigford [48]. Under these
conditions, the mean Nusselt number is given by:

Nu = 1.75 51/3 Gz 1/3 (18)

where 8
1-1"1
4 111z
3 rl+-g.~

and q = "toB / z,,,

These results have been substantiated subsequently by numerical solutions


[49,50] whereas the effect of axial conduction has been shown to be negligible
1444

for Pe > 1000 [51]. Moudachirou et al. [52] have studied the flow of Herschel-
Bulkley fluids in tubes with constant heat flux at the tube wall and presented
correlations for pressure drop and Nusselt number.
As mentioned previously, for most non-Newtonian fluids, the consistency index
is strongly temperature dependent. Therefore, any serious analysis must take into
account at least the temperature dependence of consistency index. A common
practice in Newtonian flow studies is to account for the temperature dependent
viscosity via the Sieder-Tate type empirical correlation, namely, (~tb/~t,,,)~ In
the initial studies, this approach was assumed to be applicable for non-Newtonian
fluids also [53]. Subsequent extensive numerical computations, however, clearly
bring out the deficiencies of this empirical approach [54]. Many forms of
telnperature dependence of the non-Newtonian characteristics have been used in
the literature. For instance, in the early numerical studies [54-56], the shear
stress itself was written as a fimction of temperature as:

Zrz=ln - exp (19)

where E is the activation energy of viscous flow. With this fonn of temperature
dependence, Christiansen and Craig [54] solved the coupled field equations
numerically for different values of n and tlle parameter ~(E)= (E/R) (1/To -I/T,,.)
where To is the unifonn temperature of the fluid at the inlet. These calculations
show that for n = I and ~(E) = 2, the Sieder-Tate equation overpredicts the
Nusselt number by 10% at Gz = 50 while it underpredicts by 11% at Gz = 5000.
This divergence rises with increasing values of ~I~(E). Positive values of ~I~(E)
implies that the fluid is being heated, i.e., T,,, > To. Subsequently, these authors
have obtained similar results for the cooling of fluids, i.e., ~II(E) < 0 [57]. Figure
1 shows representative results illustrating the interplay between various factors.
On the other hand, Forrest and Wilkinson [58,59] denoted the telnperature
dependence of the apparent viscosity as:

(20)
la= [1 + flw (T - To)]"

where jao is the apparent viscosity at the fluid inlet temperature To and [3w is the
temperature-viscosity coefficient to be evaluated using experimental data. Figures
2a and 2b show their representative results for tile constant wall temperature and
the constant heat flux boundary conditions respectively when the viscous
dissipation effects are negligible. Broadly, the telnperature dependent
1445

consistency index facilitates the heating of fluids whereas the cooling is impeded.
The effect is, however, very weak for the constant wall flux conditions.

10 3

10
v(E)= -3 x - 2 . -1 0 1 2
L. 10 2

F
101

10 ~ ...... I ,, I , ~ I ,

10 ~ lO l 10 2 10 3 10 4 lO s

Graetz Number, Gz

Figure 1 .Nusselt number as a fimction of Graetz number and ~I~(E)for n = 0.3 and
(E/RTw) = 10.

By far the most popular form of the viscosity-temperature relationship is the


exponential form, i.e., g =mo exp (- bT). This form has been found convenient
from a numerical standpoint as well as to approximate the experimental
rheological data. Consequently, it has been used widely to mimic the role of
temperature dependent viscosity on heat transfer [44,60-62]. Kwant et al. [60]
presented extensive numerical predictions of local and average Nusselt number
distribution and of the pressure loss trader non-isothermal conditions. The ratio
of the true pressure drop evaluated at the tube wall temperature to that under
isothennal conditions is given by:

(- Ap/L)w _ [ _ ~ 1 - ~'~b(T' v~
(- ap / - (21)

Equation (21) is applicable in the range 0.2 <_ n <


_ 1 and 0.001 _<X * < 0.4.

For heating of fluids, i.e., b(Tw- To) >_ 0, O~'k = 20tk n/(n + ]) with Otg values
given in Table 1 and for cooling, Ot'k takes on constant values as:
1446

" ' !
i
10 3 I
13w=0 13w= 1 0 , ~ = 1.2 _

n = 0.5
101
13w = 1 0 , 4 ) = 0.91

10 ~
10 2

13w = 0 .,.

Nu
13w= 1 0 , V = ' 0 " 1

10 ~ I .... I I 1
10 ~ 101 10 2 10 3 10 4

Gz

Figure 2. Mean and local Nusselt number for heating and cooling of power
law fluids with temperature dependent consistency index (11 = 0.5).

(/,'k =0.125 - 2.5 < b ( T , , - To) < 0 (22a)

=0.135 b (T,,. - To) < 2.5 (22b)

Table 1
Values of a'k for heating [60].

b(Tw- To) n~ 1 0.75 0.5 0.33 0.2


1 0.105 0.115 0.129 0.140 0.141
2 0.1 0.107 0.105 0.094 0.090
3 0.95 0.101 0.086 0.066 0.062
1447

In addition to the extensive numerical results, Kwant et al.[60] argued that when
the consistency shows strong temperature dependence, the only significant
parameter governing the rate of heat transfer is the shear rate at the wall. Based
on this premise, Kwant et al. [60] reconciled their results for the constant wall
temperature condition as:

Nu vp
- 1 + 0.271 In ~o + 0.023 (In ~o)= (23)
Nu

where r = [1 " or' k b ( T w - T o ) ] l/n (X*/6) - ot 'k b(Tw " T o ) / n (24)

for 0.001 < X* < 0.4 and 0.2 ___n < 1.

Similar expressions, though somewhat more involved, have also been presented
by Joshi and Bergles [44].
The effect of viscous dissipation on entrance heat transfer has been examined by
numerous investigators [58,59,63-73]. Since detailed discussions are available
elsewhere [33,36,69,74], only the salient features are recapitulated here. In
practice, a fully developed entrance flow condition is accomplished by preceding
the healed section with another long section. Under the conditions of significant
viscous dissipation (large Brinkman numbers), Gill [64] argued that the
establishment of fully developed flow is not feasible under the constant wall
temperature condition and therefore a plausible boundary condition can be
provided by the solution of energy equation with viscous dissipation in an
infinitely long isothennal section. On the other hand, Forrest and Wilkinson
[58,59] solved the fifll energy equation with llle viscous dissipation as well as
internal source tenns. Figure 3 shows typical results elucidating the effect of
viscous dissipation on the mean Nusselt number. Qualitatively, the study of
Fo~Test and Wilkinson [58,59] shows that even though the mean temperature of
the fluid may be lower than that of the wall, for certain values of Brinkman
number there exists the possibility of a fluid layer in the wall region with an
average temperature higher than that at the wall thereby resulting in heat transfer
from the fluid to the wall. Under these conditions, the local Nusselt number
would obviously be negative.
Some idea about the role of viscous dissipation can also be gauged from
approximate solutions [74]. For instance, for the Poiseuille flow of constant
properties power law fluids, in the so called equilibrium regime, the temperature
of the fluid becomes independent of the axial coordinate z and is a function of r
1448

alone. Under these conditions, the maximum temperature rise occurs at the
center of the tube which is given by:

10 3 !
i ia
! !
i i
!
Br=l,~=l.2

~
B r = 1 0 , ~ = 1.2
B r = 10, ~ =0.91
Br=0
Nu h
I
/
/
101 /
B r = 1 , ~ = 0.91 t
I l

10 ~
10 ~ 101 10 3 lO s

Gz

Figure 3. Mean Nusselt number as a fimction of Graetz number with si/:,mificant


viscous dissipation effects (n - 0.5).

nt n (25)
,ST ] ma -- k 3n + i)'-" v"+'
R"-!

Figure 4 shows the value of ATmax as a function of n for a polymer (m = 104


Pa.sn; k = 0.2 W/In~ flowing at an average velocity of 0.2 m/s in a capillary of
5 ~run radius. Obviously, as n decreases, the apparent viscosity of polymer
decreases, and ATm,~,,drops. The rather large values of ATma~ shown here for n =
1 are not realistic.
In the so-called transition region, the fluid temperature depends on both the
radial as well as the axial positions, and under these conditions, an approximate
expression for the temperature rise is given by:

Tb(z)-Tlr-o
411+1
\5n + I t 2 /Sn+,ll3n+,// l
ATm,~ 1 - exp -
Gzz n 4,1 + 1
(26)

where ATmaxis given by equation (25).


The effect of free convection deserves particular attention for two reasons first,
enhancements of up to 200-300% in heat transfer have been documented in the
1449

2000

g
al
l::
1000
.<3

0
0 0.5 1.0

n u162

Figure 4. Maximum temperature rise due to viscous heating for a power law fluid
in thennal equilibrium regime.

literature with natural convection being present in experimental studies [75];


Secondly and more significantly, velocity profile distortions caused by free
convection are known to induce flow instability even at low Reynolds numbers
thereby resulting in underestimation of the rate of heat transfer. Scheele and
Greene [76] suggested that flow transition and/or reversal will occur in vertical
upward flow at Gr'/Re' > ~ 52 while the corresponding value is 33 for downward
flow. Transition may occur even at lower values of Gr'/Re' in shearthinning
fluids due to greater distortions of the velocity profiles [77]. Figure 5 shows tlle
critical values of Gr'/Re' at which the maximum in velocity occurs at off-center
for upflow heating and at which the wall velocity gradient becomes zero in
downflow heating. Both Gori [78,79] and Mamer and McMillen [80] have
carried out detailed theoretical analyses to highlight the role of fiee convection on
heat transfer in power law fluids and the resulting correlations for the Nusselt
number tend to be rather cumbersome.
Analogous treatments for thermally developing flows for the other generalised
Newtonian fluid models are also available. For instance, heat transfer to the Ellis
model fluids has been investigated by Matsuhisa and Bird [41] and by Gee and
Lyon [81]. Similarly, heat transfer to Bingham and yield-pseudoplastic model
fluids has been analyzed among others by Wissler and Schechter [49], Vradis et
al. [73], Hirai [82], Schechter and Wissler [83], Henning and Yang [84] and
1450

Dakshina Murty [85]. Schenk and van Laar [86] investigated tl~e behaviour of
Prandtl-Eyring model fluids.
60 120 -
i i

9
i

1
i i|

40 80

Gr' / Re'
20 40

[
oi 0
' ,I
1
~, I,,
2
~ O!
0
, l
1
, . .I
2

Figure 5. Dependence of Gr'/Re' on power law index at which maximum


velocity moves off-centre for upflow heating (left figure) and at which
wall velocity becomes zero for downflow heating (right figure).

5. I. 4 Simultaneously Developing Flows"


When the Prandtl number is smaller than unity, the temperature profile develops
more rapidly than the velocity profile. This regime of heat transfer has received
much less atlention. Besides, owing to generally high consistency of non-
Newtonian materials, the corresponding Prandtl numbers are high and therefore
many authors have questioned the relevance of this regime to the processing of
non-Newtonian materials [33]. McKillop [87] extended the method of Atkinson
and Goldstein [88] to analyze the heat transfer to power law fluids in the entrance
region of a tube. Remote from the entrance, a perturbation solution was sought
and matched with that near the entrance. Table 2 gives a summary of their results
for three values of the Prandtl number and n = 0.5. Subsequently, McKillop et al
[89] have elucidated the effect of temperature dependent viscosity. The entrance
region heat transfer to Bingham plastics has been analyzed by Samant and
Marner [90] while Lin and Shah [91] and Victor and Shah [92] have performed
similar analysis for Herschel-Bulkley model fluids, though all are based on the
assumption of the constant physical properties.
1451

5.1.5 Experimental Studies


Laminar Regime
Numerous experimental studies oll heat transfer to power law fluids in circular
tubes have been reported in the literature [48,53,87,93-101 ].Detailed discussions
regarding their reliability and range of applicability have been provided by Cho
and Hartnell [2] and by Lawal and Mujunldar [33]. However, the salient featllres
of lhe experimental studies are re-capitulated here.
In the filly developed flow regime, the maximum or minimum temperature,
depending upon healing or cooling, occurs at the centre of the tube and the
magnitude of this temperature is important while handling temperature sensitive
materials (e.g., food, fennenlaiton broths etc.). Chann [102] measured the
centre-line temperatures of banana pure6, apple sauce and ammonium alignate
(all power law fluids) during heating or cooling in a straight tube of constant wall
temperature. Except for the ammonium alignate solution, fully developed
velocity and temperature profiles were not realised owing to the finite length (3.8
m) of the experimental section. However, the resulting values of the Nusselt
number were found to be substantially gneater than the predicted asymptotic
values of the Nusselt number thereby suggesting the presence of fiee convection.
Matthys and Sabersky [103] studied the flow and heat transfer characteristics of
tomato pure6 in circular tubes but no con'elations were presented.
In the thermal entrance region, scores of conelations are available with some of
these built in corrections for natural convection effects. Table 3 gives a selection
of widely used conelations available in the literature. While some of these
[53,62,93,94,98] account for the temperature dependence of power law
consistency, only three of them take into consideration the fiee convection
effects; the latter tend to be more important for relatively less viscous fluids.
For relatively small fluid bulk-to-wall temperature difference, the available
experimental results [104] are in good correspondence with the predictions of
Bird [105] for constant wall flux condition, as shown in Figure 6 for a carpobol
solution (n = 0.73). Note that the lilniting Nusselt number, equation (9), is also
seen to be approached for diminishing values of the Graetz number. Similar
comparisons are obtained with the data of Mizushina et a1.[94]; Mahalingam el
al. [62]; Bassett and Welty [96]; Scirocco et al. [97] and Deshpande and Bishop
[98]. Finally, the scant experimental results for viscoelastic fluids [95,104]
suggest that the value of the Nusselt number is not altered appreciably by the
viscoelasticity of the fluid.
1452

Table 2
Local and mean Nusselt numbers for simultaneously developing flow of a power-
law fluid ofn = 0.5 in a circular pipe

Constant wall temperature Constant heat flux


x* Nu Nti Nu Nu
Pr = 1
0.008 10.05 18.27 15.43 27.93
0.019 7.95 12.98 10.77 19.29
0.059 5.48 8.57 7.15 12.02
0.099 4.61 7.13 5.97 9.79
0.15 4.18 6.20 5.31 8.39
0.20 4.04 5.67 5.02 7.57
0.30 3.96 5.11 4.82 6.68
Pr = 10
0.005 11.32 16.55 14.55 24.52
0.010 8.68 13.31 11.22 18.83
0.020 6.64 10.45 8.52 14.32
0.060 4.78 7.13 6.06 9.43
0.10 4.30 6.08 5.40 7.93
0.15 4.08 5.45 5.06 7.03
0.30 3.95 4.73 4.79 5.97
Pr = 100
0.001 8.01 11.08 10.26 15.59
0.0195 6.41 9.03 8.16 12.23
0.0595 4.75 6.56 6.00 8.57
0.0995 4.29 5.73 5.38 7.39
0.1495 4.08 5.20 5.05 6.66
0.1995 3.99 4.91 4.90 6.23
0.2995 3.95 4.59 4.79 5.76

From the foregoing brief account, it can thus be concluded that for tile constant
wall flux condition, the analytical predictions of Bird [105] provide good
estimates of the Nusselt number for viscous and viscoelastic systems for small
values of AT, the correlation of Mahalingam et al. [62] and Deshpande and
Bishop [98] might be the best ones to use in design calculations.
1453

Turbulent Regime
Despite the fact that turbulent flow is encountered much less frequently with
non-Newtonian systems (except with the so called drag reducing dilute polymer

Table 3
Experimental correlations for laminar heat transfer

Constant wall heat flux

Mizushina et al. [94] Nuz = 1.4 ~51/3Gz~/3 (~-~w):' '~176

Bassett and Welty [96] Nuz = 1.85 Gz 1/3- 0.03


6

Mahalingam et al. [62] Nuz = 1.46 81/3

ln b
[Gz + 0.0083 (GrPr)~] ''~

Mehta et al. [ 106] Nuz = 1.873 Gz 1/3 + 0.87

Deshpande & Bishop [98] Nuz = 1.41 81/~

o3,:,,,o
IYIb

Constant wall temperature (mean Nusselt number)

Metzner et al. [53] Nu = 1.75 81/3 Gz1/3 I'~mb


w-w10.14

Oliver & Jenson [93]


I'~.1014
Nu = 1.75 [Gz + 0.0083 (GrPr)3(.4]1,3 mb
1454

60 1 ! , ! I I I I ' i I I "1
43 < Re < 1780 , 105 < Pr < 231 B i r d 11051
t

! n=0.73
Nuz . , _
10 I-

I
Equation(9)
2_ I I I I I ! I ~ II I I I l
101 102 103 104

Graetz Number ,Gz

Figure 6. Typical comparison between predictions and experiments for laminar


heat transfer in a circular tube (n = 0.73).

solutions and low concentration particulate suspensions), considerable effort has


been expended in investigating heat transfer in turbulent regime [2]. It is readily
acknowledged that much longer thennal entrance lengths (up to 400-500 D) are
needed under turbulent conditions as compared to the corresponding 10-15 D
required in the laminar regime. This fact alone raises questions about the utility
of some of the early data on heat transfer in the turbulent region.
Metzner and Friend [107] extended the Reichardt's framework of analogy
between heat and lnomentum transfer to power law fluids as:

St = f /2 (27)
1.2 + l l . 8 4 f / 2 ( P r - 1) Pr ''~
where the fi-iction factor, f, fimction of the generalised Reynolds number (pV 2-"
D"~/8"-~m 6n) and the power law index, n, is given by the following
equation[ 108]
(l/f) ~ = 4 n ~ log (Re f(2-,)/2) _ 0.4 n "12 (28)

Equation (27) is restricted to the condition (Pr Re2)f > 5 x 105. Preliminary
comparisons showed equation (27) to be adequate. Since this pioneering study,
many workers have reported new experimental data and/or analysis for turbulent
heat transfer to non-Newtonian systems, see [2] for an exhaustive compilation.
Based oll a critical evaluation of most previous data oll heat transfer for purely
viscous fluids, Yoo [109] put forward the following empirical correlation:
1455

St Pr,~2/3 = 0.0152 Re~"~ (29)

Equation (29) is based on data extending over the ranges 0.2 < n < 0.9 and 3000
_<Re,, < 90,000.
More recently, Kawase et al. [110] have examined the effect of pipe roughness
on heat transfer to power law fluids. Kawase and Moo-Young [111] also re-
analysed the data of Thomas [112] on turbulent heat transfer to thorium oxide
slurries and concluded that the yield stress had an adverse effect on heat transfer.
Analogous data on TiO2 suspensions [113] shows considerable scatter, though
they attempted to develop new correlations.
Considerable attention has also been accorded to heat transfer in drag reducing
systems [2,114]. Most developlnents in this field invoke the Reynolds analogy
without any justification. This coupled with the poor reproducibility of
experimental data has impeded the development of universally valid correlations
for calculating heat transfer in these systems. Therefore, all available correlations
are too tentative to be included here. Also, most mass transfer studies relate to
dilute drag reducing systems and therefore pose similar difficulties in devising
predictive correlations [2]. Kostic [115] has provided a critical review of the
pertinent literature in this field.
Thus, in conclusion though the turbulent heat transfer in non-Newtonian systems
is of considerable theoretical significance, it is perhaps of limited practical utility,
for turbulent conditions are encountered much less frequently than for Newtonian
fluids in industrial applications.

6. HEAT TRANSFER IN NON-CIRCULAR DUCTS

In contrast to the volulninous literature on heat transfer in circular tubes, the


corresponding literature for non-circular conduits is much less extensive. Some
of the comlnonly studied configurations include parallel plates or slits, square and
rectangular ducts, concentric cylindrical annuli, triangular and trapezoidal ducts
[33,34]. While some of these shapes are used in extrusion, others find
applications in heat exchangers. One distinguishing feature of heat transfer in
non-circular ducts is the variety of boundary conditions in addition to those of
isothennal walls or constant wall flux. Since extensive compilations and
discussions thereof have been provided recently by Lawal and Mujulndar [33]
and Hartnett and Kostic [34], only the key points are su~mnarised here.

6.1 Parallel Plates


For the fiflly developed flow, constant properties and in lhe absence of natural
convection effects, the asymptotic value of the Nusselt number for the constant
1456

wall flux condition is given by [116]:

2(n+ 1) -
~( it 5n +

,In +
n n -!

NH = (30)
2n+l

The corresponding expression with viscous dissipation is:

[~ (32n'- + 17n + 2)
(4n + 1)(5n + 2) (31)
Nu = (4n + l)(5n + 2)

Payvar [116] also presented an analogous expression for Binghaln plastic fluids.
On the other hand, in the hydrodynamically and thennally developing flow
regime, when the classical separation of variables method is applied to the Graetz
problem for power law fluids, a series solution is possible only for certain values
of n. Tien [117] obviated fills difficulty by using the approximate velocity
distributions deduced from file variational principle. Richardson [66] investigated
fl~e similar problem wifl~ constant thennophysical properties but included viscous
dissipation effects. His solution, however, converges only for large values of Gz
and is fires restricted to very close to fl~e entrance. Numerous other studies of
laminar heat transfer in slit geometries have been reported but all have presented
fl~eir results in gn'aphical form and therefore reference must be made to original
papers [118-123]. However, Kwant and Van Ravenstein [123] extended flleir
previous study to channel flows and presented fl~e following expressions for
pressure loss and mean Nusselt number:
!

(- Ap / L)isothemlaI

for 0 . 2 <_ n <_1 ; 0.001_< X* _<0.9 (32)

For heating, i.e., b(Tw - To) _> 0, C~'k = 2 ak n/(n + 1) where ak is tabulated in
Table 4 while for cooling

C~ ~ k -~ 0.11 - 2 _< b ( T , , , - To) < 0


1457

= 0.116 b (T,,,- To) < - 2

The mean Nusselt number with temperature dependent viscosity is given by:

Ntl] vp
= 1 + 0.238 In~b, + 0.0224 (ln ~bo)-" (33)
Nu

where ~, = 1 - Ct'k b (Tw- To) TM (X*/1.8) - Ct'k b(Tw- To)/n

Table 4
~k as a fimction of b(Tw- To) and n in a parallel plate geometry

b(Tw- To ) n --~ 1
.....
0.75 0.50 0.33 0.2
1 .
0.105
. . .
0.109
. .
0.123 0.134 0.132
2 0.10 ,,
0.102 0.100 0.88 0.070
3 0.095 0.096 0.082 0.061 0.049

Note the increasing effect of b(T,,,- To) with an increasing degree of shear
thinning behaviour.
Aside from these studies, Lin and Shah [91] and Shulman and Zaltsgendler
[124] have studied heat transfer to viscoplastic fluids whereas Yau and Tien
[125] studied developing flow heat transfer in power law fluids.

6.2 Rectangular and Square Ducts


The geometries considered so far, namely, circular pipes and parallel plates are
not only characterized by a high degree of symmetry but there is only one non-
zero component of velocity. This attribute results in appreciable simplification in
both the field equations and the non-Newlonian constitutive equations.
In turn, the shapes like square and rectangular, though still possess symmetry,
are much more difficult to analyze owing to the 2-D flow. Furtherlnore,
additional possibilities also exist in terms of boundary conditions. For instance,
one also needs to specify the condition at the duct wall in a circumferential
direction, a common variant being constant heat flux in both flow and
circumferential directions (H2) or constant heat flux in the flow direction and
constant temperature in the circumferential direction (HI) or constant temperature
in both directions (T).
For fully developed flow and constant physical properties, the results of
Chandraputla and Sastri [126] for a square duct are shown in Table 5.
1458

Table 5
Fully developed laminar flow in a square duct

n NLI1- NLIIII Nu112


1 2.975 3.612 3.O95
0.9 2.997 , ,
3.648
....
3.106
0.8 3.030 3.689 3.135
0.7 3.070 3.741 3.171
0.6 3.120 3.804 3.216
0.5 3.184 3.889 3.274

Note that shear thinning behaviour only marginally enhances the value of the
Nusselt number. Chandraputla and Sastri [126] also computed the values of the
Nusselt number as a fimction of the Graetz number and power law index for the
thennally developing flow in square ducts. Finally, they [127,128] also tackled
the problem of heat transfer in simultaneously developing flows. Representative
results are shown in Figure 7 for the isothennal wall boundary condition. Lawal
and Mujumdar [129] have also studied heat transfer to power law fluids in square
ducts under silnultaneously developing conditions.
Analogous results for rectangular ducts are fi~rther compounded due to the
additional geometric factor, namely the aspect ratio [34]. Few numerical studies
[130-134] and experimental data [135-137] on laminar heat transfer to purely
viscous and viscoelastic polymer solutions are available. Most theoretical
analyses assume constant physical properties; however, a notable exception being
a recent study of Shin and Cho [138] for a 21 rectangular duct. The effect of
viscoelasticity on heat transfer has been found to be small in rectangular ducts
[139]. Limited data on turbulent heat transfer suggest the j factor to be smaller
for viscoelastic fluids in rectangular ducts [140].

6.3 Triangular Ducts


The fidly developed laminar heat transfer to power law fluids in isosceles
triangular ducts has been numerically solved by Cheng [141] for the boundary
conditions H1 and H2; the resulting mean values of the Nusselt number are
summarised in Table 6 for both shear thinning and shear thickening materials.
Here also, both shearthinning and shear-thickening behaviours seem to have little
influence on the value of the Nusselt number.
The effect of temperature dependent viscosity on the laminar heat transfer to
power law fluids in equi-lateral triangular ducts has been investigated by Lawal
[142] in the thennal entrance region. He reported detailed temperature profiles at
1459

various axial locations thereby showing the gradual development of temperature


distribution. The role of free convection was also discussed, especially in
relation to the distortion of velocity profiles thereby yielding higher rates of heat
transfer.

12
n = 0.5 Pr'= 0

Nu 8

O0

10 20 50 100 200

Graetz Number , Gz

Figure 7. Mean Nusselt number as a fimction of Prandtl and Graetz numbers for
simultaneously developing profiles for isothermal condition.

Subsequently, Lawal and Mujumdar [143] and Etemad et al. [144] respectively
have elucidated the role of viscous dissipation and simultaneously developing
flow on the laminar heat transfer to power law fluids in equi-lateral triangular
ducts. Lawal [142] also paralleled a similar study for trapezoidal and pentagonal
ducts as encountered in the extrusion of polymers and food stuffs.

6.40iher Geomeiries
In addition to the aforementioned ideal flow geometries, industrial processes
entail a great variety of complex shapes of flow passages including helical,
double-sine [145], tapered tubes and these are not readily amenable to rigorous
analysis. For instance, Jarzebski and Wilkinson [146] solved the laminar 11o11-
isothennal developing flow, lemperature dependent consistency power law fluid
flow in a slightly tapered tube to simulate the flow in extrusion dies and they
developed a predictive expression for pressure drop. Similarly, sudden
contraction also represents a geometry of considerable theoretical and pragmatic
significance in polymer processing. Christiansen and Kelsey [147] analyzed the
1460

Table 6
Fully developed laminar flow Nusselt numbers for isosceles triangular ducts

Apex angle n NuHj Null2


(deg) ,
10 1.2 2.374 0.0756
1.0 2.391 0.0792
0.6 2.434 0.0846
. . . .

0" 1.2" 3106 1.882


1.0 3.101 1.891
0.6 3.250
,.
1.933
..

90 1.2 2.946 1.350


1.0 2.974 1.351
0.6 3.098 _
1.356

flow and heat transfer characteristics of Powell-Eyring model fluids with


temperature dependent viscosity. Likewise, Halrnnad and Vradis [72] have
studied the flow and thennal behaviour of Bingham plastic fluids. Rao [148] has,
on the other hand, reported scant data or turbulent heat transfer to viscoelastic
fluids in helical tubes and reported 10-15% enhancement in heat transfer.
Turbulent heat transfer data with drag reducing separan solutions in sudden
expansions has been reported by Pak et al. [149]. Mass transfer to pseudoplastic
fluids in spiral flow has been studied by Wronski and Jastrzebski [150] and they
showed that with a suitable choice of effective viscosity, it was possible to use
the Newtonian correlation for power law fluids also. It is unlikely that theoretical
and/or experimental results of heat transfer would ever become available for all
possible geometries and conditions. In this context, the following correction
factor, due to Cheng [141 ], is of considerable significance:

NUpo,,o~,~,, = [ a + b n ] ''~ (34)


Nu~,,.,..... (a + b)n

where a and b are purely geolnetric factors introduced by Kozicki and Tiu [151]
and are reproduced here in Table 7. The advantage of this simple approach
which is restricted to laminar flow is obvious.
Dunwoody and Hamill [152,153] predicted enhancement in heat transfer for 3rd
grade Rivlin-Ericksen fluids in rectilinear flow in rectangular channels. Similarly,
1461

Table 7
Duct Flows: Geometric Constants

Geometry (x* a b
0.1 0.4455 0.9510
0.2 0.4693 0.9739
0.3 0.4817 0.9847
CONCENTRIC ANNULI 0.4 0.4890 0.9911
0.5 0.4935 0.9946
0.6 0.4965 0.9972
0t* = inner radius/outer radius 0.7 0.4983 0.9987
0.8 0.4992 0.9994
0.9 0.4997 1.0000
1.0 0.5000 1.0000
0.0 0.5000 1.0000
RECTANGULAR DUCT 0.25 0.3212 0.8482
0.50 0.2440 0.7276
a * = height / width 0.75 0.2178 0.6866
1.00 0.2121 0.8766
0.00 0.3084 0.9253
0.10 0.3018 0.9053
0.20 0.2907 0.8720
0.30 0.2796 0.8389
ELLIPTICAL DUCT 0.40 0.2702 0.8107
0.50 0.2629 0.7886
0.60 0.2575 0.7725
0t* = minor axis / major axis 0.70 0.2538 0.7614
0.80 0.2515 0.7546
0.90 0.2504 0.7510
1.00 0.2500 0.7500
2~(deg)
I0 0.1547 0.6278
ISOSCELES TRIANGULAR 20 0.1693 0.6332
40 0.1840 0.6422
Apex angle = 2 60 0.1875 0.6462
80 0.1849 0.6438
90 0.1830 0.6395
N
REGULAR POLYGON 4 0.2121 0.6771
5 0.2245 0.6966
with N sides 6 0.2316 0.7092
8 0.2391 0.7241

, secondary flow patterns in flattened tubes and coils have been shown to yield
higher heat transfer to non-Newtonian fluids [154-156]. Such enhancements ill
1462

heal transfer have been observed experimentally [157]. The scant results point to
fi~rlher increases in heat transfer due to free convection in non-circular conduits
for purely viscous and viscoelastic fluids [158-160].
This section is concluded by noting that the highest value of the Nusselt number
occurs in circular tubes and then progressively decreasing as the cross-section of
flow changes to pentagon, square, trapezoidal and triangular. This trend is seen
for both shear thinning and shear thickening materials. Furthermore, if the
various geometries are ranked on the basis of the ratio of heat transfer to pressure
drop, a large value of this ratio is desired. In this ranking, for n < 1, the circular
duct is evidently superior to other shapes while short square tubes are to be
preferred for shear thickening materials [33].

7. HEAT AND MASS TRANSFER IN BOUNDARY LAYERS

Considerable research effort has been expended in elucidating the role of non-
linear flow characteristics on the rates of heat and mass lransfer in boundary
layers over submerged objects. Though theoretical results are available for
highly idealised shapes such as plates, cylinders and spheres, often this treatment
serves as a useful starting point for more complex situations encotmtered in
engineering applications. The available voluminous literature on both forced and
flee convection has been reviewed by several authors [161-165], only key
findings are summarised here.
At the outset, it is instructive to point out l~re that the flow of shear lhinning
(usually modelled as power law fluids) and viscoplastic media have been studied
most extensively, followed by the flow of viscoelastic fluids. Furthennore,
though it is realised that the momentum boundary-layer analysis for non-
Newtonian fluid is gennane to the study of the convective heat and mass transfer
in these systems, these are not included here for the sake of brevity. However,
attention is drawn to the pioneering efforts of Schowalter (1960); Bizzell and
Slattery (1962); Na and Hansen (1966) ; Lee and Ames (1966); Chen and
Radulovic (1973), etc. for power law liquids and of others [171-173]for
viscoelastic boundary layers. A cursory examination of the available reviews
reveals that thermal convection has been investigated more extensively than the
analogous mass transfer process. The three geometries that have been accorded
the greatest amount of attention are spheres, cylinders and plates, although some
authors [174-178] have attempted to develop general framework for 2-D
boundary layer flows. The ensuing discussion is presented in tlu-ee sub-sections,
namely, spheres, cylinders and plates, with fiwther classification depending upon
the mechanism of transport, e.g., free, forced or mixed convection.
1463

7.1 Spheres
7.1.1 Free Convection
The general field equations describing fiee convective heat/mass transport in
external flows and the associated difficulties, due to non-linear viscosity models,
in seeking rigorous solutions have been alluded to by some authors[31,161],
Similarity solutions are not feasible even for the simplest non-Newtonian
viscosity model, namely, power law model [166,179]. Acrivos [179] employed a
Mangler-type transformation to derive the following expression for the mean
Nusselt number for the constant temperature condition in the laminar boundary
layer region:

| B

Nu = ~hR = C1(ll) G r o ~ Pro3n+, (35)


k

where cl(n) is a function of n.


In another study, Stewart [180] also obtained approximate solutions for laminar
boundary layer equations for a range of shapes including a sphere and the
resulting expression for lnean Nusselt number is:
1

Nu = c, (n) (Gr, Pr,) ~~ (36)

where c2(n) is another fimction of n. Tile values of both cl(n) and c2(n) are
available in respective papers. Equation (36) was stated to be particularly
appropriate for high Prandtl numbers, as is generally the case with non-
Newtonian fluids. The theoretical developments in this field have been matched
by excellent experimental studies. Both Liew and Adehnan [181] and Amato and
Tien [182] measured the temperature distributions and the mean Nusselt numbers
for electrically heated copper spheres immersed in polymer solutions. In the
range 0.66 _<n _< 1, Liew and Adelman [181 ] reported that a single value of cj(n)
= 0.56 in equation (35) correlated their data with an average deviation of 6%.
Likewise, the experimental [182] temperature distributions around the sphere are
in line with the predictions [179]. However, they put forward the following
correlations for Nusselt number:

Nu = ,QO6~ f o r f ' 2 < 1 0 (37a)

and

Nu = 0.49 f2 (37b)
1464
| !1

for 10 < s < 40 where s = G r ~ Pro~~ .


Equation (37) was stated to correlate their results with a mean error of 8%.
Funhennore, the numerical coefficient of 0.49 in equation (37b) is remarkably
close to the theoretical value of 0.468 predicted by Acrivos [179]. Further
examination of the data of Amato and Tien [182] suggests that the ratio of the
thermal and momentum boundary layer thicknesses is of similar magnitude for
water and polymer solutions, except near the stagnation points. Alhamdan and
Sastry [183] used a musluoom shaped aluminium panicle to measure the heat
transfer coefficients under natural convection regime in the aqueous
carboxymethyl cellulose solutions and developed empirical correlations in terms
of the Rayleigh number. They estimated the apparent viscosity based on the
actual fiee convection velocity measured by tracer particles. This severely limits
the utility of their correlations because neither such velocities can be measured
with great accuracy nor can these be estimated independently. Awuah et al.
[184] while attempting to measure force convective heat transfer from a sphere to
aqueous CMC solutions reported that the free convection was important in their
study, and successfiflly correlated their data in tenns of the Rayleigh number with
an index o f - 0.1 - 0.12 which is rather small compared with the usual 1/4 power
for Newtonian systems.
The analogous mass transfer data for flee convection fiom spheres has been
reported by Lee and Donatelli [185] in the range (0.58 < n < 1). By invoking the
usual analogy between heat and mass transfer, they adapted equations (35) and
(36) by noting equality of the Nusselt and Shel~,ood numbers, Prandtl and
Schmidt numbers. But they found the following correlation [186] to represent
their data better than the mass transfer analogues of equations (35) and (36),
especially with decreasing values of Gr,n:
i

k d c,' (Qr,,,Sc,)s"-'
Shl = = 2 + - 4
(38)
1 + (0.43/ Sr ~

Witll the diminishing value of Grin, one would intuitively expect the Sherwood
number to approach the pure diffi~sion limiting value of 2. However, due to the
thin boundary layer assumption inherent in the aforementioned dleories, neither
equation (35) nor (36) approaches fills limit whereas equation (38) does. Widl
fl~e generally large values of die Schmidt number for non-Newtonian fluids, Lee
and Donatelli [185] simplified equation (38) as:

Shi = 2 + c2'(n) (GrlmSc,)~"*' (39)


1465

and they found tl~at c2'(l) = 0.59 provided satisfactory conelation of their mass
transfer data in the range 0.58 _<n _< 1.

7.1.2 Forced Convection


Westerberg and Finlayson [187] numerically solved the governing equations for
heat transfer from a sphere to a typical Nylon-6 melt, and elucidated tlle effect of
various kinematic and physical parameters. Qualitatively, the Nusselt number for
a single sphere is influenced, in decreasing order, by the value of the Peclet
number, viscous dissipation, extent of shear thinning, temperature dependence of
viscosity, viscoelasticity, and the telnperature dependent thermal conductivity.
The observation that the fluid viscoelasticity plays little role in determining tl~e
value of the Nusselt number is consistent with other works for heat transfer [188]
and for mass transfer [189,190]. Morris [191 ] studied the relative importance of
the temperature and shear rate dependence of viscosity on the flow field
produced by an isothermal sphere moving in Newtonian and power law fluids. In
a series of papers, Nakayama and co-workers [176-178] have developed an
integral method for analyzing the forced thennal convection in laminar boundary
layers over axisymmetric bodies immersed in power law liquids. By a suitable
choice of shape factors, the results can be obtained for spheres, cylinders,
ellipsoidal shapes, though the validity of this approach has been questioned for
highly shear thinning fluids (n < --- 0.3) [192]. Mixed convection heat transfer
from spheres to power law fluids has also been analyzed [193,194].
In the creeping regime, the mass transfer between a sphere and power law fluids
has been investigated by Kawase and Ulbrecht [195]. Owing to the large values
of Schmidt numbers, the fonnalism introduced by Lochiel and Calderbank [196]
is applicable for non-Newtonian systems. Thus, the Sherwood number, Sh, is
given by:

where V0* is the tangential velocity on the surface of the sphere. Kawase and
Ulbrecht [195] used the approximate stream fimctions available in the literature to
deduce the value of the Sherwood number. The predicted 50-70% enhancement
in mass transfer due to shear thinning viscosity (n --- 0.7) is in line with the scant
experimental results [197,198].Beyond the creeping flow regime, little theoretical
information is available on heat and mass transfer beween a non-Newtonian fluid
and spherical particles. Based on the penetration model [199,200], Kawase and
Ulbrecht [201 ]) derived the following formula for a sphere:
1466

1,-A n-*2 1
Shl = "~" = A(n) Re,,3~.,,> Sc ~ (41)
DAB

where the values of A(n) are available in their paper; A(n=l) = 0.85. The scant
experimental results on mass transfer from spheres [202] appear to be within
+30% of the predictions of equation (41) [165].Likewise, admittedly numerous
experimental studies on forced convection heat transfer between a sphere and
inelastic power law type polymer solutions have been reported but sufficient
details are not available to recalculate the experimental results in the form needed
by theoretical analyses mentioned in the preceding section. For instance in some
studies [203,204] particles were held stationary in moving polymer solutions
while in olher stLidies [205] particles were being conveyed in moving liquid
streams. Similarly, significant uncertainty also exists with regard to the effect of
particle-to-tube diameter ratio in exerimental results. Clearly there are some
variables which are not being duly accounted for. Similarly, based on their
experimental data, Yamanaka and Mitsuishi [206] put forward the following
semi-empirical correlation for mixed convection from a sphere to power law
fluids:
2 ! 3
hd
Nu = = 2 + { (0.866o- Pe~ - 0.553o-- 0.341) -~
k
!
I n 3 2 [ l l l b ] - -3n-,-!
+ (0.44 Gr~ <~ pr~"")2 }~ (42)
\ 111.

where cr = -2.475n 3+6.74n 2-7.67n+4.74.

Similar empirical conelations have been put fo~vard by many other


investigators, e.g., see [204,207,208] but all these are too tentative to be included
here but these have been reviewed recently [209,210].However, Ghosh et al.
[165,211] have developed a unified correlation for heat and mass transfer fiom
spheres to power law fluids by invoking the usual heat-mass transfer analogy,
i.e., Pr ~ Sc and indeed this approach does reconcile most of the data on heat and
mass transfer available in the literature. Their correlation is given by:

Y = 1.428 Rep !/3 Rep<4 (43a)

Rep 1/2 Rep < 4 (43b)


1467
1

where for heat transfer, Y = (Nul - 2) m, P r ? and Y = (Shl - 2) Scp-~/3 for


1-n b ,,/ P

mass transfer. Figure 8 shows the overall correlation. Furthermore, equation


(43) also seems to correlate the data for short (L/D--- 1) cylindrical pellets
provided the equal volume sphere diameter is used.

10 2

Equation (43a)
lO~

Equation (43b) i

10 -2

10. 3 L I I | I I I _ ! !
10 -7 10 -5 10 -3 I 0 -1 101 10 3

Reynolds Number, Rep

Figure 8. Overall correlation for heat and mass transfer from a sphere
(0.32 _< n <_ 0.93).

Mass transfer fiom spheres suspended in pseudoplastic solutions in agitated


vessels has been investigated [212,213] and the corresponding Newtonian
fonnulae provide satisfactory representation when the effective viscosity is
evaluated as m (4nN/n) nl.
7.2 C y l i n d e r s
7.2.1 Free Convection
The analysis of Acrivos [179] can easily be adapted for free convection from a
horizontal cylinder and the resulting expression is of a form similar to that for a
sphere, i.e. equation (35). However, Ng and Hartnett [214] have re-cast equation
(35) in tenns of the Rayleigh number as follows:
1

Nu = c(n) Ra 3~ (44)
1468

The thin boundary-layer assumption inherent in the analysis of Acrivos [179]


!

now becomes Ra ~3~ > 1. Since c(n) is a weak fimction of n, the Nusselt
number rises with the decreasing value of n at a fixed value of Ra. Analogous
results for constant properties Ellis model fluid are also available [215]. The
available limited experimental results for shear thinning polymer solutions [216]
are in fair agreement with theoretical predictions over the range 0.64 _< n _< 0.93.
It is also somewhat puzzling that their results of the surface averaged Nusselt
number in carbopol solutions are virtually indistinguishable from the Newtonian
correlations provided the cylinder diameter is used instead of cylinder radius! Ng
and co-workers [214,217-219] have reported extensive results on the free thermal
convection from horizontal (thin) wires to pseudoplastic and viscoelastic fluids
for the constant heat flux condition on the wire surface. These experimental
studies differ from the analysis of Acrivos [179] and the work of Gentry and
Wollersheim [216] in an important way that the wire diameters were of the same
order as the botmdary layer thickness. For pseudoplastic carbopol solutions, their
final correlation is:
i

Nu = (0.761 + 0.413n) Ra :r (45)

in the range 0.001 _<Ra < 1.

Furthermore, Ng and Hartnett [217,219] reported that the fluid was virtually
stagnant as ascertained by the stationary air bubbles trapped in the liquid. Under
these conditions, pure conduction is the main mode of heat transfer, and this is
borne out by their experiments. Subsequent experiments suggested viscoelastic
effects to be minor [214,218].

In drag-reducing dilute polymer solutions of PEO, Lyons et al. [220] reported


up to 40% reduction in laminar fiee convection heat transfer, which is
qualitatively consistent with the approximate analysis based on the second-order
fluid model[221]. Recent experimental results[222] on fiee convective mass
transfer from cylinders in shear thinning fluids also conform to equation (42).

7.2.2 Forced Convection


In a series of papers, Acrivos and co-workers [174,175,179,223] developed a
theoretical fiamework for the laminar boundary layer heat and mass transfer from
two dimensional objects including a horizontal cylinder submerged in power law
fluids. Their expression for the surface-averaged Nusselt number is of the form:
I i

Nu = c(n) Re~ ?g5 Pr- (46)


1469

where c(n) accounts for both varying degree of shear thinning behaviour, and
different shapes. Subsequently, many other investigators [224,225] have
employed different solution procedures to arrive at substantially similar
conclusions [165].Mizushina and Usui [226,227]argued that the aforementioned
analysis are flawed due to the so called zero defect at the separation and
stagnlation points. They obviated this difficulty by combining the approaches of
Kannan-Pohlhousen [228] for momentum boundary layer and of Dienemann
[229] for thermal boundary layer to obtain approximate results for the power law
and Powell-Eyring model fluids. Mixed convection from horizontal cylinders
submerged in power law fluids has been investigated amongst others by Shenoy
[230] and Wang and Kleinstreuer [231].
The first experimental study in this field is that of Shah et al. [223] who
measured the local values of heat transfer coefficient and reported good
agreement with theoretical predictions, especially for large Prandtl numbers.
Mizushina et al. [232,233] measured both the rate of heat and mass transfer from
horizontal cylinders oriented nonnal to flow. Their data show good
correspondence with equation (44) and they correlated the constant c(n) as:

c(n) = 0.72 n~ (47)

Similarly, Luikov et al. [234,235] reported tile values of mass transfer


coefficients from cylinders oriented normal to the flow of polymer solutions in
laminar region. For power law fluids, the point of separation moved downslream
ca. 0 = 155 ~ Also, their data do not seem to conform to the generic foma of
equation (44). Thus, the value of c(n) increased from 0.31 to 0.5 as the value of
the flow behaviour index dropped from unity to 0.88. The conesponding change
in the value of the exponent of Reynolds number ranged from 0.52 to 0.39.
Surprisingly, similar trends are also present in the heat transfer data of Takahashi
et al. [236] who reported the exponent of the Reynolds number to be 0.35 in the
range I > n > 0.78 as opposed to the predicted values of 0.5 and 0.55 in this
range. Kumar et al. [237] also reported extensively on mass transfer fiom
cylinders placed in cross-flow configuration in aqueous CMC solutions. Indeed,
this study elucidates the importance of the wall effects and the aspect ratio of the
cylinder. Subsequently, Ghosh et al. [238]have reviewed the literature and put
forward the following correlation for heat and mass transfer fi'om a cylinder in
cross-flow to power law media:

Nu~ (or Shl) = A Rep'" (Prp o1" Scp)1/3 (48)


1470

where A = 2.26 and m = 1/3 for Rep < 10 and A = 0.785 and m = 1/2 for Rep >
10. Equalion (48) was fotmd to be successfid not only in bringing together the
heat and mass transfer data but also the resulls for Newtonian and power law
liquids up to Re--- 25000, as can be seen in Figure 9.
Considerable interest has also been shown in studying the forced convection
from horizonlral cylinders immersed in drag reducing polymer solutions, though
results of different investigators often do not agree even qualitatively. For
inslance, James and Acosta [239] and James and Gupta [240] found that at low
flow rates, the drag coefficient and heat transfer coefficient for cylinders placed
nonnal to the flow of PEO solutions virlually coincided with the values for water
but beyond a critical flowrate (presumably Deborah or Weissenberg number),
both levelled off to constant values whereas the more recent data[241] show no
such behaviour with a Separan AP-302 solution. Numerous plausible
mechanisms including the viscoelasticity, boundary-layer thickening due Io the
solid-like behaviour at high defonnation rates, etc. have been postulated;
however, while none of these has proved to be completely satisfactory,
Ruckenstein and Ramgopal [242] developed an empirical correlation for such
anomalous data. Likewise, a few investigators have studied heathnass transfer
from horizontal cylinders rotating in drag-reducing polymer solutions [243,244]
and immersed in boiling non-Newtonian polymer solutions [245].

7.3 Plates
7.3.1 Free Convection
The theoretical treatment of Acrivos [179] results in a qualitatively similar
expression for the mean Nusselt number as that for a cylinder, equation (44).
Shenoy and Mashelkar [161] have treated in great detail free convection fi'om
vertical and horizontal plates (with isolhennal as well as constant heat flux
boundary conditions) in non-Newtonian fluids whence only the key findings are
summarised here. For instance, the scant data [216,246,247] for isothermal
surface are in line with the predictions[179,248].Shenoy and Ulbrecht [249]
presented an inlegral solution for laminar thermal convection from a vertical
surface to power law fluids under a variety of boundary conditions on the plate
surface, i.e., constant heat flux, constant temperature and temperature varying
with distance. For the constant temperature condition, their expression for the
mean Nusselt number (averaged over the plate length) is given by:
(2n + I)

3n + 1 ,-3. ,~ GrL2~"~'~ Pr,~.*, (49)


Nul = fl(n) 2n +
1471

10 2
I

Ur~
_ Equa,, 8, 785 J
I0 ~

A = 2.26
,m m
1,1
m = 1/3

1 0 -2

10"3 I. I I _ I ..... i i

10 "3 10 4 I0 x I0 3

Reynolds Number, Rep

Figure 9. Alternate correlation for heat and mass transfer from a sphere to power
law fluids.

where
n

(50a)
2 A. (10/3) ~j"

1
with An = (50b)
3n
6

f2(An) = '~;-'cz,AJ (50C)


J=i

cx~ = 1/15; or2 = -5/42;0t3=3/28


ot4 = - 1/18; (x5 = 1/63; 3/1540 13-6 -- " (50d)

Shenoy and Ulbrecht [249] reported good agreement between their data and
predictions. No analogous mass transfer results are available in the literature.
This sub-section is concluded by noting that most theoretical studies in this field
may be regarded as direct descendants of the pioneering work of Acrivos [179],
except for the fact that a variety of teclmiques have been employed to seek
approximate solutions.
1472

7.3.2 F o r c e d Convection
Little work has been reported on the forced-convective transport fonn a
horizontal plate/plane surface to non-Newtonian media since the study of Acrivos
et al. [174,175]. For the conditions when the thermal boundary layer is thinner
than the momentum boundary layer (a condition which is fulfilled over a
substantial part of the plate except near the leading edge). In this region (x --+ 0),
Acrivos et al. derived the following expression for the local Nusselt number:

":

Nux = x Re"~n*,~ Pr"2 (51 )

More recently, a complete numerical solution for laminar heat transfer from a
plane surface (aligned) immersed in Newtonian and power-law fluids has been
carried out by Pittman et ai. [250]. The predictions and experiments show good
match for detailed temperature profiles and the Nusselt number for the constant
heat flux condition on the plate in the range (30 < Re < 2000; 50 < Pr < 400).
The corresponding mass transfer problem has been studied by Mishra et al.
[251], who employed the von Kannan integral method to solve the approximate
boundary-layer equation for power law liquids. The average Sherwood number is
given by:

Sh = Co(n) Re~_''~n+'~ Sc,, ''~ (52)


( 3 :3 (
where Co(n) = 1.5 (n + 1) 2n + (4.64).:,~. ~ ,~ 2 i

(53)
l n +

Both Ghosh et al. [252] and Luikov et al. [234,235] have re-cast the results of
Acrivos et al. [174] into the equivalent mass transfer problem. While the
expression of Acrivos et al. [174] is similar to equation (52), the one due to
Luikov et al. is given below:

ShL = (0.474 + 0.436n - 0.12112) Re~ ("~ '~ Sc 1/3 (54)

Luikov et al. [234,235] reported good agreement between equation (52) and
limited data.
Subsequently, Ghosh et al. [252] have reconciled the limited data on mass
transfer from plane surfaces to a single correlation as:
1473

SIlL = 3.23 Re~ Sc~ L ,,3 0.2 < ReL < 100

= 1.12 Rel '': Scl ''~ 100 < ReL < 5000 (55)

10 3 -- i I
~ I I"l I I I I ~ ~ I I ~ 'I I

4m
10 ~

~ Slope = 1/3

10-1 I J ~_ I I ~ I J ~ ] t t .I l ~ ! t t
1 0 .2 10 ~ 10 2 10 4

Reynolds Number, Re*L

Figure 10. Correlation for mass transfer from a plate to power law fluids (plate
length 10 - 90 ram).

Figure l0 shows the overall conelation for different lengths of plates immersed
in water as well as in a series of CMC solutions. The co~xesponding range of
Schmidt number is 840 _< Sc~, _<2 x ] 0 6 while the power law index varies fiom 1
to 0.89 only.
Analogous boundary layer-analyses from continuously moving surfaces and/or
to the other type of generalised Newtonian fluids have been reported by Sharma
and Adelman [253], Lin and Shih [254], Gorla [255,256]. On the other hand,
Skelland [257] and Mishra and Mishra [258] have used the Chilton-Colburn
analogy to obtain approximate expressions for heat and mass transfer in turbulent
boundary layers. Forced convective heat transfer to power law fluids flow over a
rotating non-isothermal sphere has been analyzed [259] and as expected the
rotation yields higher values of the Nusselt number, as also noted previously by
other investigators [231]. More recently, Ruckenstein [260] has re-examined
1474

transport in laminar boundary layers and the viscoelasticity is shown to moderate


the role of the Reynolds numebr which is qualitatively consistent wilh the
observations of James and Acosta [239].
In concluding this sub-section, it is perhaps appropriate to add that although this
review has mainly focussed on heat and mass transfer from three simple
configurations, namely, a sphere, a cylinder and a plate, considerable information
is also available on mass transfer to non-Newtonian fluids from rotating spheres,
cylinders, disks, cones, inclined plates and other shapes. All these studies,
however, have been motivated largely with a view to measure molecular diffitsion
coefficients [261 ].

8. H E A T AND MASS TRANSFER IN P A C K E D AND FLUIDIZED BEDS

Packed and fluidized beds are widely used in chemical and polymer processing
applications to affect a variety of operations including mixing, filtration and
chemical reactions. Despite their overwhelming pragmatic significance in the
handling of non-Newtonian materials, very little information is available on
liquid-solid as well as bed-to-wall heat and mass transport with non-Newtonian
systems. For instance, Kawase and Ulbrecht [201] developed a semi-theoretical
expression for Sherwood number describing the mass transfer between power
law liquids and a packed bed of spheres and they reported satisfactory
correspondence between their predictions and limited data. On the other hand,
Kumar and Upadhyay [262] and Wronski and Szembek-St6eger [263] carried out
experimental works on liquid-solid mass transfer in packed beds and put forward
the following empirical correlations.

Kumar-Upadhyay

0.765 0.365
ej = + (56)
Rel '~: Re, ~"

Wronski-Szembek-St0eger:

~j= 1
0.097 Re ~ + 0.75 Re~ 6' (57)

Figure I I shows the overall CO~Telation in terms of the j-factor, correspondence


between the two independent sets of data as well as between equations (56) and
(57) is seen to be satisfacto~7 over the ranges 10-4 _< Re~ _< 100; 800 _< Sc~ _<
1475

72000. Note that equations (56)and (57) are also valid for short cylindrical
pellets (L/D <--- 1) provided an equal volume sphere diameter is used in lieu of
sphere diameter. Hilal et al. [264], on the other hand, proposed the use of (6/a,.)
where av is the specific surface area. Coppola and BNam [265] have studied
mass transfer from a bed of screens to power law fluids.

500
Equation (56)
100

"~ 10

Equation (57)
1.0

0.1
I 1 I .... l I , r J

10"4 10 "2 10 ~ 102 500

Reynolds Number, Re1

Figure 11. Power law liquid-solid mass transfer in packed beds.

The effect of drag reducing additives on mass transfer in packed bed reactors
has been studied by Sedahmed et al. [266]. Depending upon the type and dose of
the additive, they reported up to 50% reduction in mass transfer. Similar
deterioration in gas-liquid mass transfer has also been reported by others [267].
Likewise, convective heat and mass transport in fluidised beds have received
only very scant attention. For instance, Kumar and Upadhyay [262] asserted
equation (56) to be applicable to fluidised beds also. More recently, Hwang et
al. [268] reported new experimental data on liquid-solid mass transfer in fluidised
beds and found equation (56) to over-estimate the value of j factor at low
Reynolds numbers (see Figure 12). They found it necessary to develop the
following new correlation (0.01 _<Re~ _<600; 0.915 >_n >__0.63):

log (~j) = 0.169- 0.455 log (Re~)- 0.0661 (log(Re~))2 (58)


1476

No such results are available on heat transfer and for viscoelastic media.

50
Equation (58)
10
O

Equation (56)
0.1
|

0.01 [- I , ! t ~ .

0.001 0.10 10 500

Reynolds Number, Re I

Figure 12. Overall correlation for liquid-solid mass transfer in fluidized beds.

9. ]VIISCELLANEOUS GEOMETRIES

Beside the simplified geometries dealt with in the preceding sections, numerous
other configurations have been studied in the literature. While, it is not possible
to include all such studies in this review; some key references are provided here
for the benefit of the interested reader. The subject of heat transfer in
mechanically agitated vessels used for the heating/cooling of process streams has
been reviewed by several authors [74,269,270]. The CO~Tesponding literature on
gas-liquid mass transfer with non-Newtonian liquids in bubble columns and
packed towers has been summarized in many papers [271-274]. Heat transfer in
boiling non-Newtonian fluids has only begun to receive systelnatic attention [275-
280] and thus far only preliminary results are available on different flow patterns
and critical heat flux, etc. Similarly, the thennal perfonnance of compact heat
exchangers and their fouling characteristics with non-Newtonian polymer
solutions have been studied only in a cursory manner [145,281,282]. Thermal
effects in the extrusion of food and polymer extrusion have been reviewed
recently by Jaluria [283].
1477

10. CONCLUDING REMARKS AND FUTURE NEEDS

From the aforementioned discussion, it is clear that in the laminar region tlle
shear thinning enhances heat transfer and reduces pressure drop as compared
with a Newtonian fluid. Shear thickening produces the opposite effect. The
temperature dependent viscosity facilitates the heating of fluids while it impedes
the efficacy of cooling of fluids for shear thinning materials. Likewise, a
significant degree of viscous dissipation aids in the cooling of fluids with the
opposite effect in the heating. Furthermore, in situations with Br > 0, inespeclive
of the heat transfer mechanism at the wall, the local Nusselt number attains an
asymptotic value, about three times the value predicted for forced convection. In
horizontal flow, the natural convection increases both heat transfer as well as
pressure loss above those expected from the pure forced convection analyses.
While from a heat transfer standpoint, the circular geometry is the most efficient
one but when pressure loss is also of concern, short pipe lengths are to be
preferred. Besides, under certain conditions, square pipes offer a good
compromise between high heat transfer rates and low pressure drops, particularly
for shear thickening fluids. Experimental work in this area of duct flow is trailing
behind the great strides made in theoretical developments.
Heat and mass transfer in laminar boundary layer flows have also witnessed
remarkable years of progress. Consequently, sound theoretical frameworks are
now available for the prediction of heat and mass transfer from simple geometries
such as spheres, cylinders and plates, at least to purely viscous type non-
Newtonian materials. Available experimental results, while essentially in
agreelnent with the predictions, do not seem to cover anywhere near as wide
ranges of kinematic conditions as encompassed by the corresponding expressions
for Newtonian media.
Considerable scope exists fi~rther developments in this rugged terrain of non-
Newtonian fluid mechanics. For instance, the role of fluid elasticity is not at all
clear both in internal flows in ducts as well as in boundary layer flows. For
instance, secondary flows are ta~own to exist in non-circular ducts even in
creeping regime but it is not yet fully known how these impact on heat/mass
transfer. The question of stability has hardly evoked any interest. Furthermore,
most heat transfer studies employ either the constant wall temperature or constant
heat flux condition, whereas in practice the actual conditions may be somewhere
in between these two idealizations. Also, the flow patterns and heat/mass
transfer phenomena in systems of practical importance including packed and
fluidized beds, heat exchangers, boiling, etc. merit much more attention than it
has received in the past.
1478

11. NOMENCLATURE
a geometric shape factor, Table 7
B width of plate, m
Br Brinkman number, = (D/2) x,,,("+l~"~/q,,,j.tl/",.
b geometric shape factor, Table 7, - or coefficient in temperature
dependent viscosity equation, I/K
Cp heat capacity, J/kg K
d particle diameter, m
D tube diameter, m
DAB molecular diffusivity of A in B, m2/s
DH hydraulic diameter, m
E activation energy for viscous flow, J/kg K, equation (5)
f fanning friction factor, D (- Ap)/2L 9V 2
Gr Grashof number, = I3AT D 3p2g/~teff2
Grh Grashof number, = (p2d3gl3AT/m2)(d2/002"2
Gro modified Grashof number for sphere/cylinder, =
p2Rn+2(13gAT)2-n/m2
Grl modified Grashof number, p2d3g[ST/m2(d2/ot)2'~-2
=

Grim Grashof number for mass transfer, = p2d3gAp/ln2(d2/ot)2n2


Gr' modified Grashof number, = p213gqw(Dh/2)2"+2V2"2n/k~t2
or p213g AT (Dh/2) 2n+l V22n/la2, )
Gz Graetz number, = m Cp/kL
g acceleration due to gravity, m/s 2
h heat transfer coefficient, Whn2K
J j-factor, = (kdV)Sc "2/3
k thermal conductivity of fluid, W/InK
kc mass transfer coefficient, m/s
L length in flow direction, m
1"11 mass flowrate of fluid, kg/s
In power law consistency index, Pa.s"
lIl o value of m at temperature To, Pa.s"
Nu Nusselt number, = lff)/k for pipes and hR/k for spheres and
cylinders
Nul = 2Nu Nusselt number for spheres and cylinders
n power law index
q heat flux, W/in 2
Pe Peclet number, p Cp VD/2k
Pem Peclet number for mass transfer (dV/DAB)
1479

Pr Prandtl number, Cp
Pr' Prandtl number = Cp m (V/Dh)"-l/k
Pra Prandtlnumber'-Cpm
l k[ 3n4,,+ 8V]""D

Prh Prandtl number = (pC~r (1-'')/(l+,,)


Pro Prandtl number = (pCp/k)(ln/p)2/(n+l)R r 1-,,)/(l+n)(Rg[3AT)S(n-1)/2(,+I)
Pr~ Prandtl number = (mot/d2)n-l/pot
R Universal gas constant, Pa.mS/mole.K
R tube/sphere/cylinder radius, m
Ra Rayleigh number, = Gro Pro
Rea Reynolds number, oV 2-n D"/8n-'m( a n + 1.]"
4n
R ep Reynolds number, pV2-n d"/m
Rel modified Reynolds number, Rep (3,, 4nn +1) (12 (1 - e) / e")""

r radial coordinate, m
SClm Schimdt number, = (mDAB/d2)"-J/pDAB
Scp Schimdt number, = (m/pDAB)(d/V)I-"
St Stanton number, h/pV Cv
T temperature, K
Tb bulk temperature of fluid, K
T~ outlet fluid temperature, K
To inlet fluid temperature, K
Zw wall temperature, K
V mean velocity in pipe, m/s
Vz point velocity in z-direction, m/s
Vo angular velocity in 0-direction, m/s
X dimensionless axial distance, x/D
X* dimensionless axial coordinate X/Pe
Z axial coordinate, m
Greek Symbols
ot thennal diffilsivity, m2/s
Ot.'k constant in equation (22)
coefficient of thennal expansion, K l
13W coefficient in viscosity equation (20), K -~
5 ((3n + 1)/4n)
Ap pressure drop, Pa
1480

porosity of packed or fluidised beds


rl "roB/Tw
viscosity, Pa.s
~tcff effective viscosity, = re(V/d) nl, (Pa. s)
~-to viscosity at reference temperature, Pa.s
dimensionless radial coordinate, r/R
P density, kg/m 3
17rz shear stress, Pa
B
"Co Bingham yield stress, Pa
wall shear stress, Pa
volume fraction
Subscripts
b bulk
H1 pertains to thermal condition H l
H2 pertains to thermal condition H~_
iso pertains to isothennal case
L liquid
S solid
SIIS suspension
T constant temperature
vp variable property case
w wall condition
X local value
Z local value
average value

REFERENCES
1. Porter,J.E.,Trans.Inst.Chem.Engrs.,49(1971 ) 1.
2. Cho,Y.I.,Hartnett,J.P.,Adv.Heat Transf.,15(1982)59.
3. Paul,D.D.,Abdel-Khalik,S.I.,J.Rheol.,27( 1983)59.
4. Pmd'h~176176176 Interfacial Sci.,93(1983)274.
5. Irvine,Jr.,T.F.,Kim,I.,Cho,K.,Gori,F.,Exp.Heat Transl.,1 (1987) 155.
6. Hu,R.Y.Z.,Wang,A.T.A.,Hartnett,J.P.,Exp.Thennal Fluid Sci.,4(1991 )723.
7. Ishiguro,S.,Hartnett,J.P.,Int.Colmn.Heat Mass Transf.,l 9(1992)285.
8. Rodrigue,D.,DeKee,D.,ChanManFong,C.F.,J.Non-Newt.Fluid
Mech.,66(1996)213.
9. Loulou,T.,Peerhossaini,H.,Bardon,j.p., Int.J.Heat Mass Transf.,
35(1992)2557.
10. Shin,S., IntC~ 23(1996)665.
1481

11. Hubbard,D.W.,Encyclopedia of Fluid Mech.,6(1987)chapter 3.


12. Ghosh,U.K., Kumar,S.,Upadlayay,S.N.,J.Chem.Eng.Data,36(1991 )413.
13. Wickramasinghe,S.R.,Boger,D.V.,Pratt,H.R.C.,Stevens,G.W., Chem.
Engng. Sci.,46(1991)641.
14. Wasan,D.T.,Lynch,M.A.,Claad,K.J.,Srinivasan,N.,AlChE J.,] 8(1972)928.
15. Clough,S.B.,Read,H.E.,Melzner,A.B.,Behn,V.C.,AIChE .1.,8(1962)346.
16. Perez,J.F.,Sandall,O.C.,AIChE J.,l 9(1973)1073.
17. Tadmor,Z.,Gogos,G.,Principles of Polymer Processing,Wiley,New
York(1979).
18. Brandrup,J.,lmmergut,E.H. (eds.),Polymer Handbook,3rd ed.,
Wiley,NewYork(] 989).
19. Domininghaus,H.,Plastics for Engineers:Materials,Properties and
Applications,Hanser,Munich(1993).
20. Kulkami,M.G.,Mashelkar,R.A.,Polymer, 22(1981 )867.
21. Duda,J.L. Vrentas,J.S.,Ju,S.T.,Liu,H.T.,A1ChE J.,28(1982)279.
22. Kulkami,M.G.,Mashelkar,R.A., Chem.Eng.Sci.,38(1983)925.
23. ]wai,Y.,Arai,Y.,J.Chem.Eng.Jpn.,22(1989)155.
24. On" Jr.,C.,Dallavalle,J.M.,Chem.Eng.Prog.Symp.Ser.,50,No.9(1954)29.
25. Yareef,B.M.,Colloid J.(USSR),6(1940)545.
26. Skelland,A.H.P., Non-Newtonian Flow and Heat Transfer,Wiley,New
York(1967).
27. Bruggemann,D.A.G.,Ann.Phys.(keipzig),24(1935)636.
28. Raiaiah,J.,Andrews,G.,Ruckenstein,E.,Gupta,R.K.,
Chem.Eng. Sci.,47(1992)3863.
29. Dutta,A.,Mashelkar,R.A.,Adv.Heat Transf., 18(1987)16 I.
30. Feny,J.D.,Viscoelastic Properties of Polymers,Wiley,New York,3rd
ed.(1980).
31. Astarita,G.,Mashelkar,R.A.,The Chem.Eng.,Feb(1977)100.
32. Kakac,S.,Shah,R.K.,Aung,W.(editors),Handbook of Single Phase
Convective Heat Transfer,Wiley,New York (1987).
33. Lawal,A.,Mujumdar,A.S., Adv.Transp.Processes,5(1987)352.
34. Hartnett,J.P., Kostic,M., Adv.Heat Transf., 19(1989)247.
35. Kays,W.M.,Convective Heat and Mass Transfer,Mc-Graw Hill,New York
(1966).
36. Bird'R'B"Annstr~ of Polymeric Liquids,Vol
I,Wiley,New York(1977). Ilnd ed. (1987).
37. Grigull,V.,Chem.lng.Tech.,28(1956)553.
38. Skelland,A.H.P., lnd.Eng.Chem.Fundam.,6(1967)148.
39. Rohsenow,W.M.,Choi,H.,Heat,Mass and Momentum Transfer,Prentice
Hall,Englewood Cliffs,NJ(1961 ).
1482

40. Inman,R.M.,NASA Tech. Note D-2674 (1965).


41. Matsuhisa,S.,Bird,R.B.,AIChE J.,l 1(1965)588.
42. Lyche ,B. C. ,Bird,R .B.,Chem.Eng.Sci.,6( 1956)35.
43. Sestak,J.L.,Charles,M.E.,Chem.Eng.Prog.Symp.Ser.,64,No.82(1968)212.
44. Joshi,S.D.,Bergles,A.E.,AIChE J.,27(1981 )872.
45. Johnston,P.R., Mathl.Comput.Modelling,19(1994)1.
46. Tonini,R.D.,Lemcoff,N.O., lnt.Comm.Heat MassTransf.,8(1981)425.
47. Faghri,M.,Welty,J.R.,AIChE J.,23(1977)288.
48. Pigford,R.L., Chem.Eng.Prog.Symp.Ser.,51(1955)79.
49. Wissler,E.H.,Schechter,R.S.,Chem.Eng.Prog.Symp.Ser.,55 (1959)203.
50. BlackwelI,B.F.,J.Heat Transfer (ASME),107(1985)466.
51. Johnston,P.R.,Int.J.Heat Mass Transf.,34(1991 ) 1209.
52. Moudachirou,l.,Lebouche,M.,Devienne,R.,Wamle-und
Stoffubertragung,29( 1994)243.
53. Metzner,A.B., Vauglm,R.D., Houghton,G.L.,AIChE J.,3(1957)92.
54. Christiansen,E.B.,Craig,S.E.,AIChE J.,8(1962)154.
55. Hanks,R.W.,Christiansen,E.B.,AIChE J.,7(1961 )519.
56. Chrisliansen,E.B., Jensen,G.E., AIChE J., 15(1969)504.
57. Christiansen,E.B., Jensen,G.E., Tao,F.S., AIChE J.,l 2(1966)1196.
58. Fonest,G.,Wilkinson,W.L.,Trans.lnst.Chem.Eng.,51 (1973)331.
59. Forrest,G.,Wilkinson,W.L.,Trans.lnst.Chmn.Eng.,52(1974)10.
60. Kwant,P.B., Zwaneveld,A.,Dijkstra,F.C.,Chem.Eng.Sci.,28(1973)1303.
61. Mahalingam,R.,Coulson,J.M.,Can.J.Chem.Eng.,53(1975)589.
62. Mahalingam,R., Tilton,L.O., Coulson,J.M., Chem.Eng.Sci.,30(1975)921.
63. Toor,H.L., AIChE J.,4(1958)319.
64. Gill,W.N.,Appl.Sci.Res.,l 1A(1962)10. Also see AIChE J.,8(1962)137.
65. Tyabin,N.V.,Shishlyannikov,V.V.,Dachin,O.H.,Torner, R.V.,Sth Int. Heal
Transf. Conf.,2( 1974)213.
66. Richardson,S.M., Int.J.Heat Mass Transf.,22(1979) 1417.
67. Shih,Y.P.,Tsou,J.D., Chem.Engmg.Sci.,33(1978)55.
68. Glyglaszewsky,P.,Nowak,Z.,Stacharska-Targosz,J.,Wanne-und
Stoffilbertragung, l 4(1980)81.
69. Dinh,S.M.,Annstrong,R.C.,AIChE J.,28(1982)294.
70. Manglik,R.M., Prusa,J., J.Thennophysics Heat Transf.,9(1995)733.
71. Prusa'J"Manglik'RM"NumHeal Transf.,26A(1994) 199.
72. Hammad,K.J.,Vradis,G.C.,Inl.J.Heal Mass Transf.,39(1996)1555.
73. Vradis,G.C.,Dougher,J.,Kumar,S., lnl.J.Heal MassTransf.,36(1993)543.
74. Caweau,P.J.,DeKee,D.,Chhabra,R.P.,Rheology of Polymeric Systems:
Principles and Applications,Hanser,Munich(1997).
75. Melzner,A.B.,Adv.Heat Transf.,2(1965)357.
1483

76. Scheele,G.F.,Green,H.L.,AIChE J.,l 2(1966)737.


77. DeYoung,S.H.,Scheele,G.F.,AIChE J., 16(1970)712.
78. Gori,F.,J.Heat Transf.(ASME),100(1978)220.
79. Gori,F.,Int.J.Heat Mass Transf.,21(1978)247.
80. Mamer,W.J.,McMillen,H.K., Chem.Eng.Sci.,27(1972)473.
81. Gee,R. E. ,Lyon,J.B.,Ind.Eng Chem.,4 9( 1957)959.
82. Hirai,E.,AIChE J.,5(1959)130.
83. Schechter,R.S.,Wissler,E.H., Nucl.Sci.Engng.,6(1959)371.
84. Henning,C.D.,Yang,W.J.,Appl. Sci.Res., 18(1968)336.
85. Dakshinamurty, V.,Chem.Engng.Commun., 126( 1993)127.
86. Schenk, J.,van Laar, J.,Appl.Sci.Res., 7(1959)449.
87. McKillop,A.A., Int.J.Heat Mass Transf.,7(1964)853.
88. Atkinson,B.,Goldstein,S.,Moden Developments in Fluid Dynamics,Oxford
Uni 9Press, Oxford(1938).
89. McKillop,A.A., Harper,J.C.,Bader,H.J.,Korayem,A.Y.,lnt.J.Heat Mass
Transf., 13(1970)901.
90. Samant,A.B.,Mamer,W.J.,Nucl.Sci.Engng.,43(1971 )241.
91. Lin,T.,Shah,V.L.,6th Int.Heat Transf.Conf.,5(1978)317.
92. Victor,S.A.,Shah,V.L., Int.J.Heat MassTransf.,l 9(1976)777.
93. Oliver,D.R.,Jenson, V.G., Chem.Eng.Sci.,l 9(1964)115.
94.Mizushina,T.,lto,R.,Kuriwake,Y.,Yahikazawa,K.,KagakuKogaku,31 (1967)
250.
95. Sylvester,N.D.,Rosen,S.L., AIChE J.,l 6(1970)697.
96. Bassett,C.E.,Welty,J.R.,AIChE J.,21 (1975)699.
97. Scirocco, V. ,Devienne,R.,Lebouche,M.,Int. J.HeatMa ssTrans.,28( 1985)91.
98. Deshpande,S.D.,Bishop,A.A.,Chem.Engqag.Commun.,52(1987)339.
99. Filkova,l.,Lawal,A.,Koziskova,B.,Mujumdar,A.S., J.Food Engng.,
6(1987)143.
100. N guyen, V.T.,Ouardighi,A.E.,Lebouche,M.,
C.R.Acad. Sci.Paris,t.319,SerieIl(1994)983.
101. Lin,C.-X.,Ko, S.-Y.,Tsou,F.K.,Int.J.Heat Mass Transf.,39(1996)503.
102. Chann,S.E.,lnd.Eng.Chem.Fundanl.,l (1962)79.
103. Matthys,E.F.,Sabersky,R.H., lnt.J.Heat Mass Transl.,31 (1988) 1913.
104. Ng,K.S.,Hartnett,J.P.,Tung,T.T.,17th Heat Transf.Conf. (1977)p.74.
105. Bird,R.B.,Chem.lng.Tech.,31 (1959)569.
106. Mehta,D.C., DOE Report # DOE/OR/03054-58(DE84013792),Int.Coal
Refining Co.,Allentown,PA(1984).
107. Metzner,A.B., Friend,P.S.,Ind.Eng.Chem.,51 (1959)879.
108. Dodge,D.W.,Metzner,A.B.,AIChE J.,5(1959)189.
109. Yoo,S.S.,PhD Dissertation,Uni.illinois at Chicago Circle,IL(1974).
1484

11 0 .Kawase, Y., Shen oy,A. V., Wak abay sh i,K., Can .J. Chem. Eng., 72 ( ! 994 ) 798.
II 1. Kawase,Y.,Moo-Young,M.,lnt. Comm.Heat Mass Transf., 19(1992)485.
112. Thomas,D.G., AIChE J.,6(1960)631.
ll3. Quader,A.K.M.A.,Wilkinson,W.L.,lnt.J.Multiphase Flow,7(198! )545.
114. Dimant,Y.,Poreh,M.,Adv.Heat Transf., 12(1976)77.
ll5. Kostic,M., Int.J.Heat Mass Transf.,37(suppl. 1) (1994)133.
116. Payvar,P.,Appl. Sci.Res.,27(1973)297.
117. Tien,C., Can.J.Chem.Engng.,40(1962) 130.
118. Vlachopoulous,J.,Keung,C.K.J., A1ChE J., 18(1972) 1272.
119. Suckow,W.H.,Hrycak,P.,Griskey,R.G.,Polym.Engng.Sci., 11 (1971 )40 I.
120. Ybarra,R.M.,Eckerl,R.E., AIChE J.,26(1980)751.
121. Dang, V.-D.,Can.J.Chem.Engng.,58( 1980)401.
122. Dang,V.-D., J.Heat Transf.(ASME),105(1983)542.
123. Kwant,P.B.,Van Ravenstein,T.N.M.,Chem.Eng. Sci.,28(1973) 1935.
124. Shulman,Z.P.,Zaltsgendler,E.A.,4th lnt.Heat Transf.Conf.,4(1970)Rh 7.
125. Yau,J.,Tien,C., Can.J.Chem.Engng.,41 (1963) 139.
126. Chandraputla,A.R.,Sastri,V.M.K.,lnt.j. Heat Mass Transf.,20(1977) 1315.
127. Chandraputla,A.R.,Sastri,V.M.K., 6th lnt.HeatTransf.Conf.,5(1978)323.
128. Chandraputla,A.R.,Sastri,V.M.K., Num.Heat Transl., 1(1978)243.
129. Lawal,A.,Mujumdar,A.S.,Num.Heat Transf.,8(1985)217.
130. Gao,S.X.,Hartnett,J.P.,Int. Commun .Heat Mass Tran sf., 19( 1992)673.
131. Gao,S.X.,Hartnett,J.P.,Int.Commun.Heat Mass Transf.,20(1993) 197.
132. Gao,S.X.,Hartnett,J.P.,Int.J.Heat Mass Transf.,39(1996) 1317.
133. Gingrich,W.K.,Cho,Y.l. Shyy,W.,Int.J.Heat Mass Transl.,35(1991 ) ] ] 95.
134. Shin,S.,Cho,Y.]., Gingrich,W.K.,Shyy,W., lnt.J.Heat
Ma ssTran sf. ,36( 1993 )4365.
135. Hartnett,J.P., Kostic,M.,Int.J.Heat Mass Transf.,28(1985)1147.
136. Hartnett,J.P., Xie,C.,Zhong,T.,Int.Conf.Heat Transf. & Energy
Conserv.,Shenyang,PRC,October (1988).
137. Xie,C.,Hartnett,J.P., Ind.Eng.Chem.Res.,31 (1992)727.
138. Shin,S.,Cho,Y.l., Int.J.Heat Mass Transf.,37(suppl. 1)(1994)19.
139. Kostic,M.,Hartnett,J.P.,ZAMM,66(1986)239.
140. Rao,B.K.,Exp.Heat Transf.,2(1989)201.
141. Cheng,J.A.,PhD Disserlation,SUNY,Stony Brook,NY(1984).
142. Lawal,A.,Ph D Dissertation,McGill Universily,Montreal,Canada(1985).
143. Lawal,A.,Mujumdar,A.S., Warme-und - Stoffubellragung,27( 1992)437.
144. Etemad,S.G.,Mujumdar,A. S.,Nassef, R.,Appl.Math.Modell.,20(1996)898.
145. Manglik,R.M.,Ding,J., lnt.J.Heat Mass Transf.,40(1997) 1379.
146. Jarzebski,A.B.,Wilkinson,W.L.,J.Non-Newt.Fluid.Mech.,8(1981 )239.
147. Christiansen,E.B., Kelsey,S.J.,Chem.Eng. Sci.,28(1973) 1099.
1485

148. Rao,B.K.,Exp.Heat Transf.,6(1993) 189.


149. Pak,B.,Cho,Y.I.,Choi,S.U.S., lnt.J.Heat Mass Transf.,34(1991 ) 1195.
150. Wronski,S.,Jastrzebski,M., Int.J.Heat MassTransf.,33(1990) 1.
151. Kozicki,W.,Tiu,C.,Can.J.Chem.Engng.,45(1967) 127.
152. Dunwoody,N.T.,Hamill,T.A.,ZAMP,30(1979)587.
153. Dunwoody,N.T.,Hamill,T.A., lnt.J.Heat Mass Transf.,23(1980)943.
154. Oliver,D.R.,Trans.Inst.Chem.Engns.,47(1969)T18.
155. Oliver,D.R.,Karim,R.B., Can.J.Chem.Engng.,49(1971 )236.
156. Oliver,D.R., Asghar, S.M., Trans.Inst.Chem.Engrs., 54(1976)218.
157. Mena,B.,Best,G.,Bautista,P.,Sanchez,T.,Rheol.Acta, 17(1978)454.
158. Lawal,A., J.Heat Transf.(ASME) 111(1989)399.
159. Naccache,M.,Souza Mendes,P.R.,lnt.J. Heat Fluid Flow, l 7(1996)613.
160. Naccache,M.,Souza Mendes,P.R., J.Thermophysics Heat
Transf., 11(1997)98.
161. Shenoy,A.V.,Mashelkar,R.A.,Adv.Heat Transf.,15(1982)143.
162. Chhabra,R.P.,Bubbles,Drops and Particles in non-Newtonian Fluids,CRC
Press, Boca Raton,FL(1993).
163. Chhabra, R.P.,Adv.Heat Transf.,23(1993) 187.
164. Chhabra,R.P.,Adv.Transport Processes,9(1993)501.
165. Ghosh,U.K., Chhabra,R.P.,Upadhyay,S.N.,Adv.HeatTrans.,25(1994)251.
166. Schowalter,W.R., AIChE J.,6(1960)24.
167. Bizzell,G.D.,Sl attery,J. C.,Chem.Eng. Sci., 17( 1962) 177.
168. Na,T.Y.,Hansen,A.G., lnt.J.Heat Mass Transf.,9(1966)261.
169. Lee,S.Y.,Ames,W.F.,AIChE J.,12(1966)700.
170. Chen,J.L.S.,Radulovic,P.T.,J.Heat Transf(ASME).,95( 1973)498.
171 Denn,M.M.,Chem.Eng. Sci.,22(1967)395.
172. Serth,R.W., AIChE J.,19(1973)1275.
173. Verlna,R .L.,Rheol.Acta, 16( 1977) 510.
174. Acrivos,A.,Shah,M.J.,Petersen,E.E.,AIChE J.,6(1960)312.
175. Acrivos,A., Shah,M. J. ,Pet ersen,E .E.,Chem .Eng. Sci. ,20( 1965) 101.
176. Nakayama,A.,Koyama,H., Warme-und-Stoffubertragung,20(1988)29.
177. Nakayama,A., Shenoy,A.V., Int.J.Heat Fluid Flow, 12(1991)336.
178. Nakayama,A., Shenoy,A. V., J.Heat Transf.(ASME), 114(1992) 127
179. Acrivos,A.,AIChE J.,6(1960)584.
180. Stewart,W.E., Int.J.HeatMassTransf., 14(1971 ) 1013.
181. Liew,K. S.,Adelman,M .,Can. J. Ch era. En gqlg.,53( 1975)494.
182. Amato,W.S.,Tien,C.,Int.J.Heat Mass Transl.,19(1976)1257.
183. Alhamdan,A.,Sastry,S.K.,J.Food Process Engng., 13(1990) 113.
184. Awuah'G'B"Ramaswamy'H'S"Simps~176176 Process Engng.,
16(1 993)39.
1486

185. Lee,T.-L., Donatelli,A.A.,lnd.Eng.Chem. Res.,28(1989) 1056.


186. Churchill,S.W.,Chem.Engng.Commun.,24(1983)339.
187. Westerberg,K.W.,Finlayson,B.A.,Num.Heat Transfer, l 7A(1990)329.
188. Shanna,O.P.,Bhatnagar,R.K.,ZAMM,55(1975)235.
189. Kawase,Y., Mashelkar,R.A., Uibrecht,J.J.,Int.J.Multiphase Flow,
8(1982)433.
190. Ogawa,K.,Kuroda,C.,lnoue,l.,J.Chem.Engng.Jpn., 17(1984)654.
191. Morris,S.,J.Fluid.Mech., 124(1982) 1.
192. Andersson,H.l.,lnt.J.Heat Fluid Flow.,9(1988)343.
193. Wang,T.-Y.,Kleinstreuer,C.,Int.J.Heat Fluid Flow,9(1988) 182.
194. Meissner,D.L.,Jeng,D.R.,DeWitt,K.J., Int.J.Heat Mass Transf.,
37(1994)1475.
195. Kawase,Y., Ulbrecht,J.J., Chem.En~nlg.Commun.,8(1981 )213.
196. Lochiel,A.C.,Calderbank,P.H.,Chem.Eng.Sci., 19(1964)471.
197. Moo-Young,M., Hirose,T.,Ali,S.,Proc.5th
hat.Con g.Rheol.,Kyoto, 1( 1968)233.
198. Hyde,M.A.,Donatelli,A.A.,Ind.Eng.Chem.Fundam.,22( 1983)502.
199. Carberry,J.J.,AIChE J.,6(1960)460.
200. Mixon,F.O.,Carberry,J.J., Chem.Eng.Sci., 13(1960)30.
201. Kawase,Y., Ulbrecht,J.J., Electrochimica Acta,28(1983)643.
202. Kumar, S.,Mall,B.K.,Upadhyay,S.N.,Lett.Heat Mass Transf.,7(1980)55.
203. Chandarana,D.l.,Gavin,A.,Wheaton,F.W.,J.Food Process En~nag.,
13(1990)191.
204. Zuritz,C.A.,McCoy,S.C.,Sastry,S.K.,J.Food Eng.,11(1990)159.
205. Bhamidipati'S"Singh'R'K"Trans ASAE,38(1995)857.
206. Yamanaka,A.,Mitsuishi,N.,Heat Transf.-Jpn.Res.,6(1977)85.
207. Zitoun,K.B.,Sastry,S.K.,J.Food Process Engug.,l 7(1994)209 & 229.
208. Balasubramaniam,V.M.,Sastry,S.K.,J.Food Engng.,23(1994) 169.
209. Gadonna,J.P.,Pain,J.P.,Barigou,M.,Trans.l.Chem.E.,C74(1996)27.
210. Baptista,P.N.,Oliveira,F.A.R.,Oliveira,J.C.,Sastry,S.K.,J.Food
Engng.,31 (1997)199. Also see ibid 125.
211 .Ghosh,U.K.,Kumar, S.,Upadhyay,S.N.,Polym.Plast.Tech.Eng.,31 (1992)271.
21 2. Keey,R.B.,Mandeno,P.,Tuoc,T.K.,l.Chem.E. Sym.Ser.,33(1971 )53.
21 3. Lal,P.,Upadhyay,S.N.,Chem.Eng. Sci.,36(1981) 1865.
21 4. Ng,M.L., Hartnett,J.P., lnt.Comm.Heat Mass Transf.,l 3(1986)115.
21 5. Gorla,R.S.R.,Polym.-Plast.Teclmol.Eng.,30(1991 )37.
21 6. Gentry'CC"W~ Transf.(ASME),96(1974)3.
21 7. Ng,M.L., Hartnett,J.P., lnt.J.Heat Mass Transf.,31 (1988)441.
21 8. Ng,M.L., Hartnett,J.P., J.Heat Transf.(ASME), 108(I 986)790.
21 9. Ng,M.L., Hartnett,J.P., Int.Comm.Heat Mass Transf.,l 5(1988)293.
1487

220. Lyons,D.W.,White,J.W.,Hatcher,J.D.,Ind.Eng.Chem.Fund.,l 1(1972)586.


221. Shenoy,A.V.,Mashelkar,R.A.,Chem.Engng.Sci.,33(1978)769.
222. Chhabra,R.P.,Proc. 14th Nat.Heat & Mass Transf. Conf.,Kanpur (1997)in
press.
223. Shah,M.J.,Petersen,E.E.,Acrivos,A., AIChE J.,8(1962)542.
224. Kim,B.K.,Lee,H.S.,Korean J.Chem.Eng.,6(1989)227.
225. Wolf,C.J.,Szewcyzk,A.A.,Proc.3rd Int.Heat
Tranf.Conf. ,Chicago,1( 1966)388.
226. Mizushina,T., Usui,H.,Heat Transf.-Jap.Res.,7(1978a)83.
227. Mizushina,T., Usui,H., Kagaku Kogaku Ronbunshu,4(1978b)166.
228. Schlichting,H.,BoundaryLayer Theory,McGraw Hill,New York(1968).
229. Dienemann,W.,ZAMM,33(1953)89.
230. Shenoy,A.V., AIChE J.,26(1980)505.
231. Wang,T.-Y.,Kleinstreuer,C., J.Heat Transf(ASME).,l 12(1990)939.
232. Mizushina,T., Usui,H.,Kato,T., Kagaku Kogaku Ron.,4(1978)173.
233. Mizushina,T., Usui,H.,Kato,T., Kagaku Kogaku Ron.,4(1978)214.
234. Luikov,A.V.,Shuhnan,Z.P.,Puris,B.l.,Heat Transf.Sov.Res., 1(1969) 121.
235. Luikov,A.V.,Shuhnan,Z.P.,Puris,B.I., Zhdanovich,N.V.,Prog.Heat Mass
Transf.,2(1969)262.
236. Takahashi'K"Maeda'M"lkai'S"Preprint 14th Heat Transf.Sym.Jpn.
(1977)A-306.
237. Kumar,S., Tripathi,P.K., Upadhyay,S.N.,Lett.Heat MassTransf.,
7(1980)43.
238. Ghosh,U.K.,Gupta,S.N.,Kumar,S.,Upadhyay,S.N., Int.J.Heat Mass
Transf.,29(1986b)955.
239. James,D.F.,Acosta,A.J.,J.Fluid Mech.,42(1970)269.
240. James,D.F.,Gupta,O.P.,Chem.Eng.Prog.Sym.Ser.,67,No.111 (1971 )62.
241. Hoyt,J.W.,Sellin,R.H.J.,Exp.Heat Transf.,2(1989)113.
242. Ruckenstein,E., Ramgopal,A., J.Non-Newt.Fluid Mech.,l 7(1985) 145.
243. A1Taweel,A.M.,Sedahmed,H.G.,Abdel-Khalik,A.,Farag,H.A.,
Chem.Eng.J., 15(1978)81.
244. Kawase,Y., Ulbrecht,J.J., J.Appl.Electrochem.,13(1983)289.
245. Garg,N. S.,Tripathi,G.,Ind.J.Technol., 19(1981) 131.
246. Fuj ii,T.,Bull.JSME.,2(1959)365.
247. Reilly,I.G.,Tien,C.,Adehnan,M.,Can. J. Chenl.Engqlg.,43(1965)157.
248. Huang,M._j.,Chen,C.K.,int.j.Heat Mass Transf.,33(1990)119.
249. Shenoy,A.V.,Ulbrecht,J.J.,Chem. Engng.Commun.,3( 1979)303.
250. Pittman'J'F'T"Richards~ Int.J.Heat Mass
Tran sf.,37(suppl. 1)( 1994)333.
251. Mishra,l.M., Singh,B.,Mishra,P.,lndian J.Technol.,14(1976)322.
1489

A ONE-DIMENSIONAL M O D E L FOR VISCOELASTIC


DIFFUSION IN POLYMERS

C.J. Durning 1, P.H. Tang I and R.A. Cairncross z

1Department of Chemical Engineering and Applied Chemistry,


Columbia University, New YorkNY 10027

2Department of Chemical Engineering, Drexel University,


Philadelphia PA 19104

1. I N T R O D U C T I O N

Very complex motions can result when a viscoelastic fluid is put under external
forces. In general, for a given set of forces one has to solve continuity, momentum
and constitutive equations simultaneously to determine the details. The majority
of effort along these lines has focused on incompressible media, which execute
isochoric motions; viscoelastic flows with a strong dilational component have not
received as much attention, primarily because there are relatively few applications
demanding their study. One situation involving viscoelastic media where the resulting
flow is dominated by the dilational component is inter-diffusion. Here, flow occurs
in a mixture with relative motion between components, driven by a combination
of thermodynamic and mechanical forces. The simplest case involves just two
components with no applied mechanical tractions; the relative motion is mutual
diffusion, generally resulting in flow of each component dominated by dilation. We
discuss this situation for one-dimensional motions in the case of a binary mixture with
one viscoelatic component and the other an ordinary viscous fluid.
A quantitative understanding of mutual diffusion in concentrated polymer/viscous-
fluid mixtures is of great practical importance, in order to understand and control a
number of polymer processing operations and end-use applications. For example, it
1490

is a controlling physical process in devolatilization, fiber spinning, coating, drying,


microlithography and several control release schemes; accurately modeling the
process is essential to the design and control of these operations.
Experimental studies indicate that unsteady, one-dimensional diffusion in
concentrated polymer/viscous-fluid mixtures ranges from ordinary Fickian behavior
to strongly non-Fickian depending on the conditions (see, for example, references
[ 1] -[4] ). At temperatures far above the glass transition (greater than 50~ above
Ta), diffusion follows Fick's law on all accessible length scales, reminiscent of the
behavior in mixtures of simple viscous fluids. However, at lower temperatures and
high polymer densities, the dynamics begin to deviate from Fickian on small length
scales. Dramatic non-Fickian effects appear on relatively large scales throughout
the concentrated regime, near and below the glass transition. It is known that the
polymer's viscoelasticity, causes the deviations from Fickian response [5, 6], hence
the non-Fickian diffusion in such cases is often called viscoelastic diffusion. In
this work we demonstrate the capabilities of a relatively simple one-dimensional
model to account for the most obvious viscoelastic, non-Fickian effects observed
experimentally. This model allows a straitforward calculation of the diffusion-
induced flow of the polymer component during mutual diffusion.
Mutual diffusion in concentrated polymer/viscous-fluid mixtures is often studied
by a vapor sorption experiment. Since this technique is the main focus of our
modeling effort we describe it briefly, along with key experimental findings. The
mixture is confined in a thin film, initially in equilibrium with a large reservoir of
the vapor of the viscous fluid at activity a - . At time t = 0 the activity is suddenly
increased to a +, driving the system to a new equilibrium, with kinetics controlled
by diffusion into the mixture. This causes "swelling" or distention of the film in
the lateral direction, corresponding to longitudinal flow of polymer in the mixture.
The kinetics are usually tracked by measuring the evolution of concemration profiles
or by following the mass of the mixture as a function of time. For a "differential"
or "interval" sorption, a + - a - << 1 so that a very small disturbance from the
initial state occurs. For "integral" sorption, a - -- 0 and a + ~ O(1) so that a
relatively large disturbance from the initial state results. During an integral sorption,
the transport properties can vary considerably, since the local composition usually
changes significantly, so the response is typically non-linear. This may or may not be
true for a differential sorption; it depends on how much the composition changes as
1491

a result of the incremem a + - a - . In principle, if a + - a - is small enough, linear


response is observed.
Of all the results from sorption documented, two stand out as unequivocal evidence
for viscoelastic behavior during diffusion in polymer/viscous-fluid systems: Two-
stage sorption, observed in differential sorptions on thin films of concentrated
solutions, and Case II transport, observed during imegral sorption of strong swelling
agents and solvents in films of dry, glassy polymer.
Odani et al. [7-9], Billovits and Durning [10, 11] and Tang et al. [12] carried out
differential sorptions on thin films (2 - 10#m) of concentrated polystyrene solutions
over a range of compositions around the glass transition (T9) for the mixture. Two-
stage sorption was observed in all three studies just below Tg 9 The fluid mass
absorbed during sorption increased in two distinct steps. This effect seems to result
from an initial elastic longitudinal swelling of polymer in the mixture to the applied
thermodynamic driving force, followed by a protracted relaxation, controlled by the
same microscopic mechanisms involved in the viscoelastic mechanical response of
the mixture.
The other benchmark observation is of Case II transport. Here, a relatively thick
(~ O ( l m m ) ) , dry, glassy plate is suddenly exposed to an interactive fluid at near unit
activity. The most widely studied systems are alchohols in poly(methyl methacrylate)
and alkanes in polystyrene (see references [ 13] -[ 17] ). A sharp fluid concentration
front develops, separating highly distented, uniformly swollen polymer from nearly
dry, glassy polymer; the front propagates into the sample at constant speed. Hence,
Case II differs radically from ordinary diffusion, where disturbances in concentration
travel with t 1/2 and spread with time. It is clearly a non-linear effect, since it is
not seen for the same systems in linear perturbation experiments, such as differential
sorption. In principle, it should be described by the same theory adequate for linear
response, i.e. for two-stage sorption, but including key nonlinearities.
A number of years ago we derived from thermodynamic arguments a linearized,
one-dimensional model for mutual diffusion in concentrated polymer/fluid mixtures,
which includes the effect of viscoelasticity [18] . In this work, the model is
extended to nonlinear cases by introducing composition dependencies in the dominant
relaxation time of the mixture and in the mutual diffusion coefficient. There are
four material parameters in the model, which are evaluated for two systems where
careful sorption experiments have been carried out and sufficient auxiliary data
exists: Polystyrene -ethylbenzene (PS-EB) and poly(methylmethacrylate)-methanol
1492

(PMMA-MeOH). The model's predictions are compared with observations in both


systems, of two-stage sorption in the PS-EB system for differential sorption in thin
films just below Tg, and of Case II sorption in the PMMA-MeOH system for integral
sorption in relatively thick films at room temperature. Because of the practical
importance of the Case II process, in microlithograpy (see reference [19] , for
example) and in controlled release technologies (see reference [20], for example), we
also present a systematic numerical study of the effect of the four material parameters
in the model on the Case II process.

2. MATHEMATICAL MODEL

2.1 Diffusion Equation and Auxiliary Conditions for Sorption


We consider one-dimensional sorption in a thin film surrounded by an infinite fluid-
vapor reservoir. The film occupies 0 _< x _< l in the dry state along a lab-fixed
coordinate, x. During sorption, the fluid diffuses into the film along x, distending the
polymer outward along the same direction. The one-dimensional kinematics and an
assumption of no volume change on mixing [ 18] allow one to analyze the process in
terms of the fluid concentration in the film, governed by a diffusion equation derived
from the equation of continuity for the fluid and a constitutive law for the diffusion
flux; momentum balances and the polymer continuity equation are not needed. Using
"polymer material" coordinates [21 ] the fluid continuity equation is:
OC OJx
= (1)
ot ox
where X, C, and J x mean position, concentration and liquid flux in the material
system. The expression for J z derived in reference [ 18] is

0 t' OC' _,
Jx -- - D ( C ) ~ x - D'(C)-~-~g(C) r ;C')-~dt . (2)

Here, r characterizes the linear viscoelastic response of the mixture in shear; the
notation r t'; C') means that it is a functional of the concentration history between
t and t'. 9(C) - G(C)/Go with G(C) being the instantaneous shear modulus of
mixture and Go being G(C - 0), D(C) - D12(pzV2) 2 with D12 being the binary
mutual diffusion coefficient and

D'(C) - D ( C ) V I ( 0 f / 0 w 1)-lp~GO/(RTw2). (3)


1493

In equation (3), R T f is the e__quilibrium (i.e. thermostatic) fluid chemical potential,


relative to the pure fluid, Vi (V/) means the partial molar (specific) volume of
component i (i = i means fluid, i = 2 means polymer).
Substituting equation (2) into equation (1) gives the diffusion equation for C.
Note that D, D t, 9 and 05 all depend on C, which introduces nonlinearities,
making solution of the diffusion equation challenging. One should note that the
fluid concentration C is related to the local dilation of the polymer during sorption,
relative to its initial state" The Jacobian of the polymer deformation with the dry state
as reference is CV1 + 1. Hence, one can directly calculate the entire polymer flow
history from the concentration field.
It turns out that in order to use certain time-integrator software packages to solve
the diffusion equation numerically, one needs to eliminate the time derivative inside
the memory integral in equation (2), using an integration by parts, which permits
replacement of the integral by

C - q~(t, 0; C')Co -
fo t -~
0r (t, t' ; C' )C'dt.

where Co is the initial fluid concentration in the polymer film; for an initially dry film
(4)

Co equals zero.
Boundary and initial conditions need to be specified in order to solve for C. For
the boundary conditions, at the interfaces between the mixture and the reservoir, we
assume continuity of the fluid chemical potential, #1, which can be written in terms
of the fluid activity in the reservoir, tfl, as

In ~ - In Cgq + Q c/)(t,t'; C') dCb dt'


oo (5)
where Cb means the boundary concentrations, C(X = 0, 1). The second term on the
right in equation (5) corresponds to a relaxing (nonequilibrium) contribution to the
fluid's chemical potential in the mixture. The first term on the right corresponds to
the choice f - ln(C/C~q) for the thermostatic potential, implying the approximation
of Henry's law for the equilibrium isotherm between the vapor reservior and the
condensed mixture. C~q is a constant corresponding to the fluid concentration in
the film at equilibrium with a reservoir at unit activity and Q - ( 1c~- ~ ) ( ~1 ) , where c~
is explained below.
Equation (5) proves awkward with certain integration schemes. Integration by
parts in equation (5) gives and alternative form
1494

Cb t ,t' ' ' '


In r - In C Oq+ Q [r t';C;)Cs o - Q r ; Cb)Cbdt (6)

which requires an initial value for Cb. Taking the limit for t ~ 0 gives

In Cb + a ( cb Co ) - - 0 (7)
c or 1 - ~ Cgq co
defining the initial value.
So far, the development can accommodate any viscoelastic model. Hereafter, we
assume single exponential relaxation with a concentration dependem relaxation time
t dr"
r c') - - ~ s,, . - ~ . (8)
In order to include the most important nonlinearities in the simplest way, we adopt the
following assumptions, which approximate experimental results in the concentrated
regime.
(i) The relaxation time decreases exponentially with concentration:

T(C) = TO exp - m C (9)


where m is a constant, and TO is the relaxation time for the dry polymer.
(ii) The diffusion coefficient increases exponentially with concentration:

D ( C ) = Do exp k C (10)
where k is a constant and Do is the diffusion coefficient of the dry polymer.
(iii) The instantaneous shear modulus of mixture, G, is independent of
concentration, which implies g(c) - 1.
Equation (3) then gives:

D ' ( C ) " D~o(C/C~ expkC (ll)


with

D~ -- DoGoV1 ~ p 2 V 2 C ~ (12)

2.2 Scaling
Introduce the dimensionless quantities

u - C X t(Do + D~o) kC~q M - mC ~ (13)


COeq ; x-- T; 8- 12 ; K- ; eq"
1495

where l is the dry film thickness. After substituting the expressions for
D(C), "r(C), and r C'), and using equation (4), the diffusion equation
becomes:
Ou 0
Os --
- Ox (1 - a + au) exp K u OuO--x aOaO OI
. x u exp K u Ox (14)

with
I -- Ouo exp (/0
- s0 exp~- M u ~)
+
Jo~ exp(Mu') exp
x: M~ u " / u_' d s '
( I- , ~~ 0 e~'' (15)

where
a -- D6 0 - T0(D0 + D6) (16)
D~ + Do' 12
are dimensionless constants. Equation (6) becomes

lnr
~ ( Ub--
o) (17)

where ~ -- C~/C;~
For numerical solutions, it is easier to work with differential equations. Equation
(15) is the solution of
OI I
0---~= exp M u ( u - ~). (18)

Here, 0I has an important physical meaning. It represents the dimensionless


nonequilibrium contribution to the local fluid chemical potential. Differentiation of
equation (17) gives

dub = _ Ub exp Mub In u~. (19)


d~ 0[~ + (1---z%)~] r
The initial conditions on u are

u--uo @ t--O forO<x < l. (20)


The initial conditions needed for I are

I(x,s-O)-O for O<x< l. (21)


1496

The initial conditions at the boundaries can be obtained from the scaled form of
equation (7). The experimentally measurable relative fluid uptake, W(s), can be
calculated from

W(s) - Jo.1(u(z, s) - uo)dz. (22)

Caimcross and Duming [22] derived an equivalent dimensionless representation,


based on equation (2), i.e. integration of the memory integral by parts (equation (4))
was not done. Somewhat different boundary conditions were employed in that study
(see subsequent discussion).

3. N U M E R I C A L M E T H O D S
We aimed to solve equations (14)-(21) for the broadest possible range of the
dimensionless parameters c~, 0, K, M, uo and ~b. In the Case II limit, steep moving
concentration fronts are expected. This demands a numerical scheme able to handle
"stiff" problems effectively. Consequently, finite element spatial discretization,
together with an ODE integrator designed to handle stiff initial value problems was
the approach chosen.
Two different finite-element schemes were used. The first employs orthogonal
collocation. The details are more or less the same as described by Fu and Duming
[23] and Tang et al. [ 12]. A solution is constructed by discretizing space into equal
subintervals and approximating the spatial dependence with piece-wise continuous
Hermite cubic polynomial basis functions. Orthogonal collocation is then applied to
each subinterval to convert the PDE system into a coupled system of ODEs for the
time dependent coefficients in the polynomial approximations to u(z, s) and I(z, s).
We used LSODI [24] for the ODE solver, which requires that the system be in linearly
implicit form; equations (14) and (18), employing equation (4), collapse to a linearly
implicit ODE system when discretized.
The second scheme is that discussed by Caimcross and Durning [22] . The
dimensionless form of the diffusion equation retaining the memory integral in
equation (2) is integrated directly by Galerkin's method, i.e. equation (4) is not
used, using quadratic basis functions and a differential/algebraic equations systems
solver, DASSL [25] . Actually, the boundary condition used with this scheme at
the reservoir/mixture interface is somewhat different from equation (19): A "surface
evaporation" condition is imposed where the external fluid flux is put proportional
1497

to the fluid activity drop between the film surface and the bulk phase and set equal to
the flux in the film at the surface, given by equation (2). The proportionality constant
appearing in the external flux at the surface is a mass transfer coefficient; for very
large values of the mass transfer coefficient, the surface evaporation condition gives
that the surface and bulk phase fluid activities nearly match, i.e. it reduces to our
surface condition, equation (19). In fact, calculations based on the two schemes agree
closely when the (dimensionless) external mass transfer coefficient in the second
scheme (a Biot number) is set to large values ( ~ O(103)) and all other model
parameters are set equal (see below). Values of the Biot number ~ O(103) were
used for all of the calculations by the Galerkin scheme.

4. N U M E R I C A L RESULTS
There are six dimensionless parameters in the model, a, 0, K, M, u0 and ~b. The
first four characterize the polymer-fluid system, while the last two define the driving
force for the sorption. For all the calculations, we fixed the number of subintervals
at 30. As a basic benchmark we solved the case a - 0, K - M - 0 and recovered
linear Fickian diffusion, as expected. By trial calculation, we found that the numerical
solutions were most suseptible to instabilities for integral sorption (u0 -- 0.0 and
= 1.0). In this case, when a is near 0, K and M must be less than 4 in order
to avoid instability of the collocation scheme. When a is away from 0, M can be
increased to 5, while K can go up to 12 before numerical instabilities occur with
collocation. The code based on Galerkin's method enjoyed a wider range of stability
in parameter space for integral sorptions than the one based on collocation, permitting
values of both M and K up to 12 for a wide range of a and 0.
A comparison of concentration profiles predicted by the collocation and Galerkin
schemes for the demanding case of integral sorption near the Case II limit (a =
0.9, 0 = 0.01, K = 5.0, M = 5.0, u0 = 0.0, and~b = 1.0)showed
excellent agreement between the two codes. Subsequent remarks refer to results from
collocation, except where noted.

4.1 Predictions of Two Stage Sorption and Case II Transport


Billovits and Durning [10, 11] reported a careful set of differential sorption
experiments in the poly(styrene)-ethylbenzene (PS-EB) system. They clearly
observed non-Fickian, viscoelastic effects, including two-stage sorption, just below
Tg. The first four model parameters ~, 0, M and K can be calculated for the PS-EB
1498

system from published data; the details are given in Appendix D of Tang [26]. For
a 5# thick PS film, at 40~ we find c~ -~ 0.97, 0 ~ 432, M ~ 3 5 , / ( ~ 17.
To model the conditions where two-stage weight uptake was observed, we picked
u0 - 0.03, and ~ - 0.1, which corresponds to run R4 in reference [10] . This
choice corresponds to the solvent mass fraction starting at w I = 0.018 and ending
at w~- - 0.059, somewhat too large of a change for purely linear response, i.e. the
response is weakly non-linear. Figure (la) shows u(x, s) while figure (lb) shows
the corresponding dimensionless weight uptake, W(s) vs. x/~, which is clearly a
two -stage weight uptake curve. The concentration profiles show that the two-stage
sorption process is controlled by the viscoelastic response in surface concentration,
which which was first recognized by Long and Richman [27].
Consider next liquid methanol (MeOH) in poly(methylmethacrylate) (PMMA). It
is known that Case II transport occurs when liquid methanol contacts a relatively
thick ( ~ lmm) PMMA film at room temperature (see reference [ 14], for example).
Under these circumstance, u0 -- 0, and ~b = 1, i.e. we are considering an integral
sorption. From the physical properties provided by Thomas and Windle [14] we
find c~ ~ 0.86, 0 -~ 0.0087. We adopt K = 6.2, as suggested by Thomas and
Windle. According to Thomas and Windle, M = 13.0; however, M = 5.0 was used
due to the limitation of the collocation scheme. Consider the concentration profiles
predicted for this system, shown in figure (2a). Note two key features, characteristic
of Case II [ 14-16]. First, the surface concentration relaxes with time. Second, a steep
concentration front develops and propagates into the film at nearly constant speed
after the surface equilibrates. Figure (2b) shows the corresponding (dimensionless)
weight uptake, W(s) vs. s. The main, increasing part of the plot is nearly linear,
and there is an induction time visible as illustrated by the construction on the figure;
both features are additional signatures of the Case II process. Clearly, the predictions
for the PMMA-liquid MeOH system capture all the characteristic features of Case
II transport noted in previous work. The forgoing calculations demonstrate that the
model considered here can predict the two most well documented manifestations of
viscoelasticity in sorption experiments: Two-stage uptake in differential sorption and
Case II transport in integral sorption. The remaining sections focus on the effect of
the material parameters a, 0, M and K on the predictions of the model in the Case
II limit, assuming integral sorption, i.e. u0 - 0 and ~b = 1.
1499

0.12

0.10

-~. 0.08
9.08

0.06

0.04

0.02
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.20

0.16

0.12

0.08

0.04

-, , , _,, l_., , .... L .... ,. | 9 ,_ , , I - J - ~ -~ i

0 3 6 9 12

S~r2

Figure 1. (a). Concentration profiles (u(z, s)) from the non-linear model with
parameters for the PS-EB system studied by Billovits and D u m i n g [ 10] 9a -- 0.97,
0 -- 432, K - 17, M - 35, uo = 0.03, ~b - 0.1; (b). Weight uptake kinetics ( W
vs. s) from the non-linear model with the same set of parameters.
1500

.0 _ _ . rM

0.8

0.6
g-
0.4 24

02 s=O.O 12

O0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

1.0

0.8

0.6

0.4

0.2

0.0
0.00 si 0.01 0.02 0.03 0.04 0.05 0.06

Figure 2. (a). Concentration profiles (u(x, s)) from the non-linear model with
parameters for the PMMA-MeOH system studied by Thomas and Windle [14] :
c~ = 0.86, 0 = 0.0087, K = 6.2, M = 5, Uo = 0.0, ~ = 1.0.; (b). Weight
uptake kinetics ( W vs. s) from the non-linear model with the same set of parameters.
1501

4.2 Surface Concentration in the Case II Limit


It is a well-known feature of Case II transport that the surface concentration
increases at a finite rate during the initial stages of sorption. This effect is predicted
by the model. A dimensionless induction time, si, characterizing the time scale for
surface equilibration, can be calculated from the numerical Ub vs. s plots using a
construction analogous to that on the weight gain plot in figure (2b). Hui et al. [ 15],
and later Duming et al. [28], developed asymptotic results for si for a model of Case
II transport by Thomas and Windle [ 14] (TW model hereafter) corresponding to a
small Deborah number limit of the model considered here (see reference [23] for a
discussion). The asymptotic result for si is
1
s~ ~ M >> 1. (23)
MlnM
Duming et al. [28] also show

si ~ 0 (24)
according to the TW model.
In the present model, three parameters a, 0, and M appear in the equation
governing surface concentration, equation (19). One sees that the effect of 0 is
only to rescale s, which implies that si ~ 0 as in the TW model (equation (24)).
The parameter M describes the dependence of the relaxation time on concentration.
Figure (3) shows its effect on surface swelling with a fixed at 0.99, and 0 set to 1.0;
lnM ranges from 2 to 3.25. When in M < 1.5 ( M < 4.48), there is no discemable
induction time; the Ub vs. s curves in this case do not show upward curvature. For
In M >_ 1.5 ( M >_ 4.48), autoacceleration of Ub occurs, and an induction time can be
calculated using a process analogous to that illustrated in figure (2b). Figure (4) plots
si vs. ( M in M ) -1 for a - 0.99 and shows a linear relationship. Taking differences
in 0 into account, the slope agrees in order of magnitude with the value obtained by
Fu and Duming [23] from the TW model.
In the current model the parameter a also affects the induction time. a shows up
as a measure of the instantaneous elasticity of the polymer in the mixture; there is
no analog for a in the TW model, since it presumes the polymer is purely viscous.
a -- 0 means a perfectly compliant polymer and corresponds to ordinary diffusion;
for finite a instantaneous elasticity and relaxation effects are switched on. Figure (5)
shows the effect of a on surface concentration with M fixed at 10.0. For finite a one
sees an initial jump, permitted by the instantaneous elasticity, which decreases with
1502

increasing c~. In the linearized version of the model, the surface concentration jumps
initially to 1 - a (Tang et al. [ 12] ). In the non-linear version, the value given by the
dimensionless version of equation (7) is near, but not exactly equal to 1 - c~. From
figure (5) one sees that when a is less than 0.1, the relaxation of UD is minimal and
there is no discernible induction time. As c~ increases from 0.1 to 0.9, the amplitude
of the relaxation in Ub increases, induction times si appear and increase with a. The si
increase monotonically with a, but the numerical data do not suggest an unambiguous
analytical representation for si(a).

1.0 ! I '

0.9

0.8

0.7

0.6
.Q

0.5

0.4

0.3

0.2 A M = 12.18
v M= 15.64
0.1

0.0 , ~ ~ I , ~ , I J , J I , J '

0 2 4 6 8

Figure 3. Effect of M on surface concentration (Ub vs. s) for a -- 0.99 and 0 -- 1.0.
1503

9 , , ! , , , !

0 i i I I = |

0.00 0.02 0.04 0.06 0.08

1/MIn(M)

Figure 4. Linear relation between induction time from surface concentration kinetics,
si, and ( M In M) -1.

1.0

0.9

0.8

0.7

0.6

0.5

0.4

0"3 I a a=0.3
0.2 ~ " a=?.5
[ - a=0.7
0.1 [ o a=0.9
/
0.0
0.000 0.005 0.010 0.015 0.020

Figure 5. Effect of a on surface concentration kinetics (Ub VS. 8) for M - 10.0 and
O - 1.0.
1504

4.3 Profiles and Weight Uptake in the Case II Limit


In what follows we systematically investigate the effects of a, 0, M and K
on the concentration profiles (u(s, x)), nonequilibrium chemical potential profiles
(OI/Os(x, s) ), weight uptake kinetics (W(s)) and several other characteristic
features, including the concentration front's speed and the induction time based on
weight gain kinetics. All of the calculations are on parameter ranges where Case II
transport is predicted, u0 and ~ are fixed at 0 and 1, respectively, which corresponds
to a dry polymer film suddenly exposed to unit fluid activity.
4. 3.1 Effect of a
Trial calculation showed that concentration profiles only begin to resemble Case-II
when a is near 1. To determine its effect on the process, we varied it over the range
0.9 to 0.999. Provided 0, M and K also have appropriate values (see later sections),
the weight uptake plots in this range are near linear with time, and show an induction
as in figure (2b). From this, we conclude that in order to observe Case-II phenomena
D~ >> Do (see equation (16)). Figure (6) shows the effect of a on weight uptake in
detail for 0 = 0.01, K - M -- 5.0. From the plot one sees that, when a is small,
there is practically no induction. The induction time shows up more clearly when a
increases; at the same time the weight uptake curves become more nearly linear with
time.
Induction times for the process were calculated from figure (6) as illustrated
in figure (2b); The results are qualitatively, consistent with those for induction
times determined from Ub, i.e. the induction times from weight gain kinetics and
surface concentration both increase monotonically with a, but they do not agree
quantitatively. The quantitative discrepancy warns that experimental induction times
determined from weight gain measurements not be interpreted in terms of surface
concentration kinetics.
Figures (7) show u and 0~ profiles for a - 0.9 with 0 - 0.01 and K and M
fixed at 5. Recall ~oI is the (dimensionless) nonequilibrium contribution to the local
chemical potential, and is part of the driving force for diffusion. Initially a large peak
in 0I develops at the surface. As time increases, the peak decays, broadens, and then
propagates into the film with fixed shape, at nearly constant speed. Eventually, the
two peaks from both sides of the film combine into one at the film center and slowly
relax while concentration reaches equilibrium.
The peak in 0I is responsible for the dominant feature of Case II transport, i.e.
1505

1.0

0.8

0.6

0.4 / / = o~= 0.92


/ / . -o.oo

0.0 , I , , , ,
i. i ,

0.0 0.2 0.4 0.6

Figure 6. Effect of a on W vs. s for 0 - 0.01, K - 5.0, M - 5.0, Uo - 0.0,


1.0.

the sharp concentration front which establishes near the surface after the external
activity is switched on and propagates into the polymer film at a constant speed. We
conducted systematic calculations to investigate the effect of a on the key features of
OI
the from: The front speed v and the values of u and 8-; at the front (uf and (o~):),
with the front's position being defined by the position of the maximum in N. oI Figure
(8) displays a typical trace of u: and (oi) i with time, for M - / f i - 5.0, a - 0.9
and 0 = 0.01. The plot shows that, after an initial induction at the surface both u /
and ( ~ ) I achieve nearly steady values and move at nearly constant speed into the
film. When the front reaches the center of the film, (oi N ) : stops and sinks to zero
gradually. At the same time, uf relaxes to the equilibrium concentration.
Two measures of the moving front speed were calculated. One, v:, is the slope of
the linear part of front position versus time, and the other, Vw, is the slope of the linear
portion of W(s) vs. s. Since the moving front invades the film from both sides, vw
corresponds to about twice Vv (the values ofvf and vw/2 agree within 8%).
1506

1.0

0.8

0.6

0.4

0.2

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

0.20

0.18

0.16
s = 0.0004
0.14

o.12 s= 0.011
o.lo s=oos s=o.o3 /~/I
0.08

0.06

0.04

0.02

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

Figure 7. (a) Concentration profiles (u(z, s)) determined by two different numerical
schemes. Hollow symbols correspond to collocation method while filled symbols
correspond to Galerkin method. The model parameters are a = 0.9, 0 -- 0.01,
K = 5.0, M = 5.0, Uo = 0.0, ~b = 1.0.; (b). c9I/c3s profiles from collocation for
the same set of parameters.
1507

Results for effects of a on both v~ and "of with K = M = 5.0 and 0 = 0.01 are
summarized in figure (9), which shows that the v decrease linearly with increasing
a. That plot suggests v f ~ (1 - a) since the speed gets vanishly small near a - 1.

0.3 ~ Uf

0.2

?
0.1

0.0
0.0 0.1 0.2 0.3 0.4 0.5

Fig 8. u f , ( O I / O s ) s vs. x ; a - 0.9, 0 -- 0.01, K - M - 5.0, Uo - 0.0, r - 1.0.

4.3.2 Effect of O
0 is a Deborah number for this model. It corresponds to the ratio of a characteristic
relaxation time of the polymer to the characteristic diffusion time, evaluated for
the dry film. For nonlinear diffusion, the behavior cannot be anticipated from 0
along, as in the linear limit, since the actual relaxation and diffusion times vary with
concentration making the instantaneous, local value of the Deborah number vary
significantly. However, for this model in the special case M = K the concentration
dependence of the relaxation and diffusion times cancel, and 0 does correspond to the
dimensionless relaxation time at all compositions. It happens that M ___ K is required
for Case II (see later), so we may view 0 as a key parameter for Case II.
1508

14[: .... I .... t .... I .... = ....


d
12O
~ -i

10
8 [- ~. Vw

4_ vf

-2
0.90 0.92 0.94 0.96 0.98

Gt

Figure 9. v f and v~ vs. a with 0 - 0.01, K - 5.0, M - 5.0, Uo - 0.0, r - 1.0.

We found that 0 must be _< O(0.01) in order to predict Case II, consistent with
Fu and Durning's [23] analysis which derived the TW model of Case II as a small
Deborah number limit of the model considered here. We picked the range 0.008 _< 0
_< 0.022 to study the effect of 0 on the process. From calculations of concentration
profiles and weight uptake with a -- 0.9, / ( = M = 5.0, one finds that an increase
of 0 causes a proportional increase in the induction time calculated from weight ~ain
kinetics, i.e. si ~ 0, exactly as anticipated from analysis of the surface concentration.
0 also decreases front speed, and therefore the slope of the linear portion of W ( s ) vs
s plots. The results for v f and vw are plotted in figure (10); we find v f ~ 0 -0.52 under
these conditions. Fu and Durning [23] studied the TW model numerically and found
v ~ 0 -~ identical with the dependence of in the current model in the Case II limit.
One can anticipate this result from dimensional analysis.
1509

1.2 9 9 l-- w 1 w i i w

1.1
1.0
0.9
c
0.8
b.O z
9
0.7
0.6 Iogvf
0.5 L
0.4 -

3 1 I I I I I

-5.0 -4.5 -4.0 -3.5


logo

Figure 10. vf and v~ vs. 0 with a - 0.9, .K - 5.0, M - 5.0, Uo - 0.0, ~b - 1.0.

4.3.3 Effects of M and K


The parameter M controls how fast the local relaxation time decreases with
concentration; it characterizes one of the nonlinearities in the model. The effect of M
on weight uptake kinetics was determined for a -- 0.9, 0 = 0.01 and K -- 5.0 for
2 < M < 10. The data showed that increasing M decreases the induction time, in
qualitative, but not quantitative agreement with the result from the analysis of surface
concentration (figure (4)). When M is large, the average relaxation time in the surface
layers is low, and it takes less time to establish equilibrium at the surface. Increasing
M also increases the slope of the linear portion of weight uptake plots, corresponding
to an increase in front speed. The effect of M on the front speed is summarized figure
(11) for a = 0.9, 0 = 0.01 and K = 5.0; the data for, M = 6, 8, 10, were generated
by Galerkin's method. The front speed increases monotonically with M, with the
data for vf being nearly linear with M. Duming et al.'s [28] asymptotic prediction
1510

from the TW model, that v ~ ( K + M ) a/2 is consistent with figure (11).

22 . . . . 1 ' ' '" I . . . . I . . . . I . . . . I ' ' ' r

20

18 II
16 ~ vw
14
El
12 r-I 9 -
lO ~ n
_ o 9 -
- 0
8. [3 9 -
6~ 0 0 O0

4
. ooOO v, :

2 , = j , I , , , I 1 . . . . I , , , , I , , = i I . . . .

0 2 4 6 8 10 12

Figure 11. Dependence of v / a n d vw on M with a - 0.9, 0 - 0.01, K - 5.0,


Uo - 0.0, ~b - 1.0. Filled symbols represent values calculated by Galerkin's method.

The numerical solutions give evidence that M has to be sufficiently large in order
to predict Case II 9 We calculated u and 0I profiles with 3/1 -- 2 and other parameters
as above. By comparing the results with figure (7), one finds that the decrease of
M from 5 to 2 suppresses the Case II features. The decrease results in considerable
broadening of the concentration front; at the same time, the peak in 0I continuously
decays at M = 2 and spreads progressively more along the spatial coordinate during
sorption.
The parameter K govems how fast the diffusion coefficient increases with
concentration and characterizes the other key nonlinearity in the model. Figure (12)
shows the dependence of induction time from weight gain on K with ce = 0.9, 0 =
0.01 and M = 5.0. The four data points in the high K range were generated
1511

by the Galerkin code. One expects that the induction time from weight gain be
independent of K, if it truly characterizes the surface relaxation process, i.e. if it
corresponds closely to the induction calculated from the surface relaxation kinetics
(e.g. figure (4)). However, we find that the induction time from weight gain increases
linearly with K until near the value of M (5.0 in figure (12)) after which it becomes
independent of K as expected. This indicates that diffusion in the surface layers can
limit the induction time determined from weight gain kinetics. We note, however, that
compared with the effect of M, which reduces the induction time even if the value
exceeds K, the influence of K on the induction from weight gain is much weaker, as
one expects intuitively.
Calculations show a roughly linear relationship between front speed and K, as
shown in figure (13) for M = 5, a - 0.9, 0 = 0.01. This finding is again consistent
with that by a perturbation analysis ofthe TW model, which predicts v ~ ( K + M ) 1/2.
Calculations showed that in order to have Case II, M needs to be at least 2 - 3
and K has to be at least equal to M. For example, it was found that if K is kept at
5.0, when M rises above 8, the predicted behavior deviates from Case II in that the
linear weight uptake kinetics were not predicted. This is because when M exceeds
K, the front speed becomes so fast that diffusion behind the front quickly becomes
rate limiting. Consequently, the front begins to show diffusive dynamics (v ~ sl/2).

5. C O N C L U S I O N
A one-dimensional nonlinear model for viscoelastic diffusion in concentrated
polymer-fluid mixtures was constructed by an ad-hoc generalization of the linear
response model by Durning and Tabor [18] . The nonlinearities were introduced
by retaining concentration dependencies of physical properties. In order to get a
numerical solution for sorption in films, the diffuson equation was cast as coupled
partial differential equations by introducing a new dependent variable. Finite element
methods were used to discretize the spatial domain, converting the PDEs into an ODE
system, which was solved by a time integrator package. In one scheme, orthogonal
collocation on Hermite cubic basis functions was used to discretize and LSODI was
adopted to do the time integration. An alternative integration technique employed
Galerkin's method on quadratic basis functions together with the DASSL integrator.
The two schemes were shown to agree.
There are six dimensionless parameters in the model, a, 0, M, K, u0 and ~b. The
1512

0.0092

0.0088
O
9 9 @
0.0084

0.0080

0.0076

0.0072

3 4 5 6 7 8

Figure 12. Dependence of the induction time from weight gain kinetics, si, on K with
a - 0.9, 0 - 0.01, K - 5.0, Uo - 0.0, ~b - 1.0. Filled symbols represent values
calculated by Galerkin's method.

first four characterize the mixture and the last two define the initial and final states for
sorption. We first investigated the predictions for two well-studied situations, where
all the parameters could be calculated apriori: Differential sorption in polystyrene-
ethylbenzene (PS-EB) solutions and integral sorption in poly(methylmethacrylate)-
liquid methanol (PMMA-MeOH). For PS-EB, the two-stage sorption process was
correctly predicted for differential sorptions in thin films at concentrations just below
Tg. For PMMA-MeOH, the Case II diffusion was correctly predicted for immersion
conditions in thick, dry plates at room temperature. The calculation shows that
the model can predict the most well documented and striking non-Fickian effects
observed in sorption w i t h o u t empiricism.
A systematic investigation was conducted of the effects of the materials parameters
1513

22 '"" ' ' I ' i , , I ' "' ' I ' ; ; ' I ' "' ' I ''w ' ' "

20 i

18
16
Vw
14
12
10
0 9
D 9 9 9
8
6 0 0
0 0 Vf
4

4 5 6 7 8 9

Figure 13. Dependence of vf and vw on K with a - 0.9, 0 - 0.01, M - 5.0,


Uo - 0.0, ~b -1.0. Filled symbols represent values calculated by Galerkin's method.

on Case II diffusion. The study should facilitate analytical asymptotic work, and
provide guidelines for design and control of systems relying on Case II. Case II
diffusion appears only when a, a measure of the instantaneous elasticity of the system,
is close to 1. This implies D~ > > Do and physically means that the osmotic modulus
is weak compared to the mixture's shear modulus. The situation occurs when the fluid
is a poor solvent or swelling agent. It was found that the front's speed decreases nearly
linearly with increasing a. 0, the diffusion Deborah number, has to be ~ O(0.01)
for Case II to appear, indicating that Case II is a slow-motion limit of the model. 0
affects the front speed according to v ~ 0 -1/2.
The numerical study shows that strong nonlinearities in the relaxation time and
diffusion coefficient, represented by large M and /4, are both essential for the
prediction of Case II transport. Importantly, the values of M and K should be about
the same, and at least 2 - 3. I f M exceeds K by too much, the process rapidly becomes
1514

controlled by diffusion behind the moving front.


Regarding the numerical methods used, collocation can handle parameters K an
M only up to about 5, beyond which the calculation becomes unstable. For modelin
more severe non-linearities, one should employ Galerkin's method, which seems t
have a wider range of stability in parameter space.
It was demonstrated that the model predictions qualitatively match the behavk
observed experimentally. A forthcoming publication [29] reports in detail o
practical procedures for evaluating the model parameters, and on the quantitatN
capabilities of the model.

6. Acknowledgment
1LAC acknowledges support from Sandia National Laboratory.

REFERENCES
1. Alfrey, T., E. E Gumee, and W. G. Lloyd, ~ Polym. Sr C, 12, 249 (1966).
2. Crank, J., and G. S. Park, Diffusion in Polymers, Academic Press, Londc
(1968).
3. Frisch, H. L., Poly Eng. & Sr 20, 2 (1980).
4. Vrentas, J.S., and J.L. Duda, Encyclopedia of Polymer Science ar,
Engineering, 5, 36 (1986).
5. Vrentas, J. S., C. M. Jarzebski, and J. L. Duda, AIChE J., 21, 94 (1975).
6. Vrentas, J.S., and J.L. Duda, J. Polym. Sci., Polym. Phys. Ed., 15, 441 (1977~
7. Odani, H., S. Kida, M. Kurata, and M. Tamura, Bull. Chem. Soc. Japan, 3,
571 (1961).
8. Odani, H., J. Hayashi and M. Tamura, Bull. Chem. Soc. Japan, 34, 817 (19611
9. Odani, H., S. Kida and M. Tamura, Bull. Chem. Soc. Japan, 39, 2378 (1966)
10. Billovits, G.E, and C. J. Durning, Macromolecules, 26, 6927 (1993).
11. Billovits, G.E, and C. J. Durning, Macromolecules, 27, 7630 (1994).
12. Tang, PH, C.J. Durning, C.J. Guo, and D. DeKee, Polymer, 38, 1845 (1997).
13. Hopfenberg, H.B., J. Memb. Sci., 3, 215 (1978).
14. Thomas N. L., and A. H. Windle, Polymer, 23,529 (1982).
15. Hui, C. Y, K.C. Wu, R.C. Lasky, and E. J. Kramer, J. Appl. Phys., 61, 51"~
(1987a).
16. Hui, C. Y, K.C. Wu, R.C. Lasky, and E. J. Kramer, J. Appl. Phys., 61, 51"~
1515

(1987b).
17. Durning, C.J., M.M. Hassan, K.W. Lee and H.M. Tong, Macromolecules, 28,
4234 (1995).
18. Durning, C. J., and M. Tabor, Macromolecules, 19, 2220 (1986).
19. Rodriguez, E, P.D. Krasicky, and R.J. Groele, Solid State Tech., 28, 125 (1985).
20. Roseman, T. J., and S. Z. Mansdorf, Eds., Controlled Release Delivery Systems,
Marcel Dekker, New York (1983).
21. Billovits, G.E, and C. J. Durning, Chem. Eng. Comm., 82, 21 (1989).
22. Cairncross, R.A., and C.J. Durning, AIChE J., 42, 2415 (1996).
23. Fu, T. Z., and C. J. Durning, AIChE J., 39, 6 (1993).
24. Painter, J. E, and A. C. Hindmarsh, "Livermore Solver for Ordinary Differential
Equations (Implicit Form)," Lawrence Livermore National Laboratory Report
L-316 (1982).
25. Brenan, K.E., S.L. Cambell, and L. Petzhold, Numerical Solution of Initial-
Value Problems in Differential-Algebraic Equations, Elsevier, New York
(1989).
26. Tang, P.H., " Analysis of Differential and Integral Sorption in Concentrated
Polymer Solutions," PhD Thesis, Department of Chemical Engineering,
Materials Science and Mining, Columbia University, New York (October,
1995).
27. Long, EA., and D. Richman, J. Am. Chem. Sot., 82, 513 (1960).
28. Durning, C.J., D.S. Cohen, and D.A. Edwards, AIChE J., 42, 2025 (1996).
29. Huang, S.-J., and C.J. Durning, J. Polym. Sci., Part B: Polymer Physics in
press.
1488

252. Ghosh,U.K.,Dey,K.N.,Gupta,S.N.,KtHnar,S.,Upadhyay,S.N.,Chem. Engng


.Commun.,43( 1986a)335.
253. Shanna,K.K.,Adelman,M.,Can.J.Chem.Engng.,47(1969)553.
254. Lin,H.-T.,Sllih,Y.-P.,Chem.Engng.Commtm.,4(1980)557. Also
7(1980)327.
255. Gorla,R.S.R.,Chem.Eng.Commun.,49(1986)13.
256. Gorla,R.S.R.,Polym.-Plast.Technol.Eng.,30(1991)75.
257. Skelland,A.H.P., AIChE J.,l 2(1966)69.
258. Mishra,I.M.,Mishra,P.,Indian J.Technol., 14(1976)375.
259. Kim,H.W.,Esseniyi,A.J.,J.Thennophysics Heat Transf.,7(1993)581.
260. Ruckenstein,E., Ind.Eng.Chem.Res.,33(1994)2331.
261. Coppola,L.,Bohm,U., Int.Commtm.Heat Mass Transf.,13(1986)77.
262. Kumar,S., Upadhyay, S.N., lnd.Eng.Chem.Fundam.,19(1980)75.
263. Wronski,S.,Szembek-Stoeger,M.,Inzyni. Chem.Procesowa,4(1988)627.
264. Hiial,M.,Bnmjail,D.,Colniti,J.,J.Appl.Electrochem.,21(1991 ) 1087.
265. Coppola,L.,Bohm,U.,Chem.Eng.Sci.,40(1985)1594.
266. Sedahmed'G'H"Mans~
J.App.Electrochem., 17(1987)583.
267. Potucek,F.,Stesjkal,J., Chem.Eng.Sci.,44(1989)194.
268. Hwang,S.-J.,Liu,C.-B.,Lu,W.-J.,Chem. Eng.J.,52(1993)131.
269. Edwards'M'F"Wilkins~ Chem.Eng.,No.257( 1972)310.
270. Desplanchaes,H.,Llinar,J.R.,Chevalier,J.L.,Can..l.Cllem.Eng.,58(1980)160.
271. Schugrel,K.,Adv.Biochem.Engng.,l9(1981 )71.
272. Moo-Young,M.,Blanch,H.W.,Adv.Biochem.Engng.,19(1981 )1.
273. Deckwer,W.D.,NguyenTien,K.,Schumpe,A.,Serpemen,Y.,Biotech.
Bioeng.,24(1982)461.
274. Chhabra,R.P.,Ghosla,U.K.,Kawase,Y.,Upadhyay,S.N.,in Multiphase
Reactor and Polylnerisation System Hydrodynamics, ed.N.P.Cheremisinoff,
Gulf, Houston (1996) p.539.
275. Hartnett,J.P. ,Hu,R.Y.Z., lnt.Comm.Heat Mass Transf.,l 3(1986)627.
276. Wang,A.T.,Hartnett,J.P., Wanne-und-Stoffubertragung,27(1992)245.
277. Floquet-Muhr,L.,Midoux,N.,Chera.Engllg.Process.,33( 1994)459.
278. Desplanchaes,H.,Gaston_Bonhomme,Y.,Chevalier,j.L., Int.Chem.Eng.,
34(1 994)225.
279. Liu,J.,Ye,L.,Liu,H.,J.Chem.Engng.Jpn.,28(1995a)210.
280. Liu,J.,Ye,L.,Liu,H., Int.Colmn..Heat Mass Transf.,22(1995b)359.
281. Rene,F.,Leuliet,J.C.,Lalande,M.,Chem.Eng.Res.Des.,69C(1991 ) 115.
282. Fossa,M.,Yagliafico,L.A.,Exp.Thennal Fluid.Sci.,10(1995)221.
283. Jaluria,Y.,Adv.Heat Transl.,28(1996)145.

S-ar putea să vă placă și