Sunteți pe pagina 1din 11

JOURNAL OF COLLOID SCIENCE ~0, 267-277 (1965)

TRANSPORT CHARACTERISTICS OF SUSPENSION': VIII. A NOTE


ON THE VISCOSITY OF NEWTONIAN SUSPENSIONS
OF UNIFORM SPHERICAL PARTICLES

David G. Thomas
Oak Ridge National Laboratory, Oak Ridge, Tennessee
Received October 14; 196~, revised December 6, 196~

ABSTRACT
A critical analysis was made of the extensive experimental data on the relative
viscosity of suspensions of uniform spherical particles. By appropriate extrapolation
techniques, non-Newtonian, inertial, and nonhomogeneous suspension effects were
minimized. As a result, the scatter of the data was reduced from ± 7 5 % to ± 1 3 % at a
volume fraction solids of 0.50. The coefficients of different power series relating rela-
tive viscosity and volume fraction solids were determined using a nonlinear least
squares procedure. It was shown that a new expression containing three terms of a
power series with coefficients determined from previous theoretical analyses and an
exponential term with two adjustable constants fit the data as well as a power series
with six terms, either three or four of which were adjustable constants with the re-
maining coefficients being theoretical values.

INTRODUCTION
Previous papers in this series have been largely concerned with the
transport characteristics of non-Newtonian suspensions (1). One of the
major problems encountered in studies of such suspensions in the deter-
ruination of the limiting viscosity at high rates of shear (2). This Value is
necessary for use in turbulent transport correlations (1) and for verifica-
tion of asymptotic values predicted by different rheological equations of
state (2). In either case the viscosity of dispersed Newtonian suspensions
is of special interest, because it represents a mininmln value for the limiting
viscosity of non-Newtoniaa suspensions.
Since Einstein published an analysis for the viscosity of dilute suspen-
sions (3) in 1905, one of the most challenging rheological problems has been
the development of theoretical or empirical expressions for the viscosity of
concentrated suspensions. The results of these studies have been sum-
marized in recent review articles (4-6). As yet, there is no commonly ac-
cepted theoretical treatment for concentrated suspensions; perhaps the
1 Research jointly sponsored by The Office of Saline Water, U. S. Department of
the Interior, and U. S. Atomic Energy Commission under contract with the Union
Carbide Corporation.
267
268 THOMAS

principal reason for this is the difficulty of the analysis coupled with the
poor agreement among different experimental investigations (even though
the majority of these studies were carried out under carefully controlled
conditions). The object of this note is (I) to propose criteria for isolating
hydrongmic interaction of particles independent of such factors as floccula-
tion, particle size, particle size distribution, and type of viscometer and
(2) to apply these criteria to the experimental data and the evaluation of a
consistent set of parameters for several different expressions for the sus-
pension viscosity.
THEORY
Extreme Dilution
The theoretical calculation of the viscosity of suspensions of solid par-
ticles can be approached in several ways. First, the viscosity may be deter-
mined either from the velocity gradient (4, 7)
{OU, OU~
r,k = --~ \ ~ + b--~J' [1]

or from the energy dissipation in the bulk of the fluid (4, 7)


f
E~ = - ~ Jv OXk\OXk + OXi] dV. [2]

The limiting case of extreme dilution may be evaluated either by consider-


ing the perturbation in the flow field due to the presence of a continuously
decreasing number of particles per unit volume (8), or by considering the
interaction of a constant number of particles of continuously decreasing
size (9). The principal difference between these last two approaches is
that the first assumes no long-range interaction between particles whereas
the latter one does. In general, there is good agreement among the different
theoretical studies in that the viscosity of extremely dilute suspensions,
~,, is a simple function of the volume fraction solids:
,,/~0 = 1 + k~¢. [3]
For solid spheres with diameter large compared to molecular dimensions
but small compared to the characteristic length of the measuring instru-
ment, for no sllp at the sphere surface and for purely laminar flow, the
value of k~ is commonly accepted to be 2.5, although values as large as 5.5
have been suggested (9). In either case, Eq. [3] is valid only up to concen-
trations of the order of 0.01 volume fraction solids.
Concentrated Suspensions
With more concentrated suspensions, it is necessary to account for the
hydrodynamic interaction of particles, particle rotation, collision between
TRANSPORT CHARACTERISTICS OF SUSPENSIONS 269

particles, mutual exclusion of particles, doublet and higher order agglomer-


ate formation, and ultimately mechanical interference between particles as
packed bed concentrations are approached. Perhaps the greatest difficulty
in arriving at a theory for these concentrated suspensions is the fact that
the random structure of the suspension cannot in general be represented by
a simple model.
Many of the existing theoretical and experimental equations can be
expressed as a power series:
~8/#0 = 1 + k~¢ + k202 -4- k3¢~~ :i= . . . , [4]
in which the constant kt is generally assumed to have the value determined
by Einstein. Guth and Simha (10) obtained a value of ks = 14.1 by a
method of successive reflections in which it was assumed that additional
flow around a second sphere compensated for the disturbance in flow around
the primary sphere. In a later study, Saito (11) determined the effect of
the mutual volume of exclusion in addition to the above factor and obtained
a value, ks = 12.6. Vand (12) studied the problem by considering the effect
of adding an incremental volume fraction of spheres, de, to a dilute sus-
pension. After accounting for the hydrodynamic interaction of spheres in a
manner similar to that used by Guth and Simha (10) and also for the effect
of doublet formation due to collisions, Vand arrived at a value k2 = 7.349.
Later, Manley and Mason (13) showed that the period of doublet rotation
was larger than assumed by Vand, resulting in k2 = 10.05. In general, the
values of the coefficients of the higher order terms are less accurately known
since they must include more complicated interactions than it is possible
to handle theoretically. Values obtained from Guth and Simha's and Vand's
analyses indicate a contribution of from 16 to 50 to the value of k~ from
doublet formation and hydrodynamic interaction.
The main drawback to relations of the form of Eq. [4] is that termination
of the series after the ~ term means a 10 % or greater error in the viscosity
ratio for volume fraction solids greater than 0.15 to 0.20; inclusion of the
~3 term extends the validity of the series form to q~~ 0.40.
Several expressions have been proposed which give the relative viscosity
as a function of concentration in a closed form. Mooney (14) arrived at a
functional equation for the relative viscosity by considering two successive
additions of uniform spheres to pure fluid. In this way he accounted for
possible hydrodynamic interactions and the mutual crowding effect of the
two fractions of spheres on each other. Solution of the functional equation,
making use of Einstein's relation, Eq. [3], as ~ --~ 0 resulted in
#~/t~0 = exp [2.5~b/(1 -- a2~)], [5]
where the value of a~ was to be determined experimentally. From a modcl
based on the geometry of the cubic and face centered cubic lattices and
270 THOMAS

the fact that for spheres in contact the relative viscosity must be infinite,
Mooney suggested a value 1.35 ~ a2 ~ 1.91.
Ford (16) advanced the proposition that the experimental data may be
fitted more accurately by an expression of the form:
~0/~ = 1 -- 2.,~b -t- b~ 5 -- b~4)7, [6]
where values of the constants were taken to be b2 -- 11.0 and b3 = 11.5
oil the basis of Vand's data. Ford also suggested that the 4)5 term becomes
important at the onset of the inhibition of particle rotation and the 4)7
term becomes important at the onset of particle interlocking.
The finite size of the particles which effectively shields the central particle
from interaction with other than nearest neighbors at the concentration is
increased has been treated by Simha (15). The effect of the shielding re-
duces the result calculated on the assumption that the particles are points.
To obtain a solution to the hydrodynamic equations, the cage of particles
is replaced by a concentric spherical enclosure. As a result, the model is
only approximate for dilute concentrations. For dilute concentrations, the
result can be expanded to
~/#0 = 1 + 2.5¢ [1 + 25¢/4 f3 . . . ] , [7a]
whereas for concentrated suspensions the relative viscosity is approximated
by
lira ~/~0 = 1 -{- ( 5 4 / 5 f ) [ 4 ) 2 / { 1 - (¢/4) .... )}3], [75]
~b-*¢ m a x

where 1 < f < 2 but may be expected to vary within those limits over the
complete range of concentrations. It should be emphasized that Eqs. [7a]
and [75] are approximate expressions the use of which should be limited to
conditions of moderate ( 4 ) ~ 0.10) and very large concentrations, respec-
tively (35).
DATA ANALYSIS
Data from many different investigations of the viscosity of suspensions
of spheres are shown in Fig. 1. The data show a spread of from about :i:20 %
at ~ = 0.20 to about :i:75 % at 4) = 0.5. These data were obtained with
both rotational and capillary viscometers and represent, a range of particle
diameters from 0.099 to 435 microns. In the majority of cases, the particles
were rather closely sized. Since the data given by Eveson (19) show that
mixtures of spheres having a fourfold variation in diameter have less than
a 6 % difference in relative viscosity up to ~ = 0.20, it is believed that the
variation about the mean particle diameter in the investigations cited above
contributed very little to the scatter of the data. The particle materials
included polystyrene, rubber latex, glass, and methyl methacrylate. In all
TRANSPORT CHARACTERISTICS OF SUSPENSIONS 271

t0 2
REFERENCE
e I
O 25
[] 26 I
A 27
V t5
/
• 12 /
• 24 j B~
& 28
• 20 /' I
29 /
OD 25
(D 21
Ill 19 c~
2 in 30 /
=t pl 22 ~ 0 /
>_- G
'¢'
5!
3z /
B ()

o /
/ /,
10 Be /
I--
/
/
/ pr "'

/
in

I ~' o /
'• in

,,,fi ,~# ,o /

>7
f Fo-
• u/3
. o ~ ~'°
0 0.1 0.2 03 0.4 0.5 0.6 07 0.8
VOLUME FRACTION SOLIDS,

Fig. ~. C o l l e c t e d Relotive Viscosity Dora.

studies, either the density of the suspending medium was adjusted or the
viscosity of the suspending medium was sufficiently large that settling was
unimportant. Examination of the experimental procedure used in these
studies shows no basis for eliminating any of the data because of faulty
technique; consequently, there nmst be at least, one additional parameter
that has not been accounted for.
272 THOMAS

10 2

EQ, 70 AND 7b
WlTt TABLE I
. . . . EQ. 9

>.-
I-.
0
:
40
> #
laJ 8,, /0
>
I-
<Z
_1 6
laJ
cl:

./
~s

jv
,J'
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
VOLUME FRACTION SOLIDS,

Fig. 2. Reduced Relative Viscosity Data.


(See Fig. 1 f o r key to data)

According to de Bruijn (17), one parameter of importance is the absolute


value of the particle diameter. For particles with diameters less than 1 to
10 microns, colloid-chemical forces become important causing non-New-
tonian flow behavior (17). The result is a relative viscosity which increases
as particle size is decreased, but which decreases to a limiting value as the
shear rate is increased (2). For particles larger than 1 to 10 microns, de
TRANSPORT CHARACTERISTICS OF SUSPENSIONS 273

Bruijn (17) believes that inertial effects due to the restoration of particle
rotation after collision result in an additional energy dissipation and a
consequent increase in relative viscosity with increasing particle diameter.
In flow through capillary tubes, the increase in viscosity observed with
large particle size suspensions is opposed by a decrease in viscosity caused
by a tendency for particles to migrate toward the center of the tube as the
particle diameter is increased (18).
Examination of the data from which Fig. 1 was prepared showed that
in several cases the tests covered a sufficient range of shear rates or particle
sizes that it was possible to extrapolate to conditions where particle size
effects were negligible. For particles less than 1 micron diameter, the limit-
ing value of the relative viscosity was obtained as the intercept of either a
linear plot of 1/Dp versus u~/uo or a linear plot of 1/(du/dr) versus #.~/uo.
For particles larger than 1 to 10 microns, the limiting value of the relative
viscosity was obtained as the intercept of a linear plot of Dp versus u,~/uo •
In the event that large particle size data were also available as a function
of shear rate, the reduced particle size data were further corrected by
plotting against 1/(du/dr). T r e a t m e n t of the suitable data in this manner
gave a unique curve for which the m a x i m u m deviation was reduced from
three- to sixfold over t h a t shown in Fig. l, that is, to 4 - 7 % at 4, = 0.20
and to 4-13% at ~ = 0.5 (Fig. 2).

COEFFICIENT EVALUATION

Care must be used in the imlnericM analysis of the data to determine


coefficients in the different expressions because of the large absolute values
of the variation of u.~/uo from the mean for concentrated suspensions. ( T h e
data from concentrated suspmlsions are also the least precise owing to the
experimental difficulties of working with concentrated suspensions.) In
order to reduce the disproportionate effects of the large absolute values of
the variation from the mean, various weighting factors were tried. The
one which was ultimately most successful w~s 1 (/U~/lao -- 1).
When a curve with no inflection points was carefully fitted to the data,
the value of [(u.~/u0)- u,---~c]/(~-7/tto- 1) was substantially uniformly
distributed within a band of +0.16, - 0 . 1 4 for 0 < 4) < 0.40, and the
limits increased from these values to ± 0 . a a a t , = 0.60. Thirty-six of the
65 values were positive; the remainder were either zero or negative. With
the use of the weighting factor, the variauce of the data from this mean
curve was 0.013. The empirical curve was fitted with Eq. [7a] for 0 < ~ <
0.15 and with Eq. [7b] for 0.20 < ~ < 0.602. For Eq. [7a] the value of the
constant, f, was 1.111. In order to determine the value o f f in Eq. [Tb], it is
necessary to specify the value of ~ ...... This was determined by extrap-
olating a plot of 1/[(u,o/t*0) - 1] vs. ~ to zero giving ~ ..... = 0.625, a value
which corresponds very closely to the mean volume fraction solids of a
274 THOMAS

TABLE I
Coefficients for Eq. [7b]

(P p~/Po f

0.20 1.95 1.136


0.25 2.41 1.304
0.30 3.09 1.490
0.35 4.19 1.695
0.40 5.97 1.953
0.45 8.90 2.327
0.50 14.80 2.903
0.55 28.3 4.106
0.60 70.0 9.583

TABLE II
Coeffcients of Power Series
Iz8

IZo

Case ~ A2 A3 A4 As A6 A7 Variance

1A -880 -0.0121 646 925 0 0 0.289


1B -2240 -4.86 914 2480 0 0 0.338
2A -11.7 165 0 -1140 0 4000 0.154
2B -41.2 313 0 -1760 0 5000 0.261
3A (7.349) b 239 -- 1450 2350 0 0 0.207
3B (7. 349) 299 -- 1720 2640 0 0 0,341
4A (7. 349) 58.8 0 -- 657 0 3182 0.159
4B (7.349) 66.2 0 -- 725 0 3330 0.274
5A (10.05) 214 - 1380 2290 0 0 0.210
5B (10.05) 276 - 1660 2590 0 0 0.343
6A (10.05) 43.9 0 - 588 0 3065 0.161
6B (10.05) 52.4 0 - 668 0 3240 0. 276
7A (12.6) 191 - 1320 2230 0 0 0. 212
7B (12.6) 255 - 1610 2550 0 0 0.344
8A (12.6) 29.7 0 - 523 0 2960 0.164
8B (12.6) 39.4 0 - 614 0 3150 0.279

a All " A " cases: 0 < ~ < 0.60.


All " B " cases: 0.2 < ~ < 0.60.
b Fixed c o n s t a n t in least squares analysis.

r a n d o m l y p a c k e d b e d of u n i f o r m s p h e r e s d e t e r m i n e d b y e x t r a p o l a t i o n t o a
v a l u e of z e r o f o r t h e r a t i o of s p h e r e d i a m e t e r t o c o n t a i n e r d i a m e t e r ( 3 3 ) .
W i t h (~max ~ - 0 . 6 2 5 , t h e v a l u e s of f i n E q . [7b] w e r e d e t e r m i n e d ; s e l e c t e d
values are given in Table If
As pointed out b y Simha a n d Somcynsky (35), Eqs. [7al and [7b] were applied
over a greater c o n c e n t r a t i o n range t h a n i n t e n d e d originally (15). A more exact appli-
cation is t r e a t e d elsewhere (35).
TRANSPORT CHARACTERISTICS OF SUSPENSIONS 275

The coefficients of various combinations of terms in Eq. [4] were de-


termined using a nonlinear least squares procedure. The results are given
in Table II. The same program was used in an attempt to determine the
coefficients of Eqs. [5] and [6]; however, it was not possible to obtain con-
vergence on a curve that had no inflection points.
The terms
1 ~ 2.5~ -~ 10.05~ 2 [8]
account for over 97.5 % of the value of the relative viscosity for ~b < 0.25.
At this concentration the average separation of particle surfaces is esti-
mated to be only 0.35 particle diameter for a randomly packed array.
(This value was obtained from the geometric mean of values for the cubic
and close packed arrangements--a procedure which gives a value of 0.6229
for these two arrays when the particles are in contact. This latter value is
very close to the value ~ .... -- 0.625 determined empirically for use with
Eq. [7b].) For values of ~ > 0.25, the above three terms account for smaller
percentages of the relative viscosity, e.g., 60 % at ~ = 0.40 and 8.7 % at
= 0.60.
According to Vand's (12) theory, the three terms of [8] account only for
the effect of hydrodynamic interaction of spheres and particle-particle
collisions which form doublets. Consequently, it seems quite surprising
that there is less than 2.5 % deviation of [8] from the data even when the
particle surfaces in the suspensions are estimated to be only 0.35 diameter
apart. Clearly at larger concentrations there must be considerable rear-
rangement of particles in a suspension as the suspension is sheared. This
suggests that instead of continuing the series of [8], it would be more ap-
propriate to add a term which would be proportional to the probability of
a particle's transferring from one shear plane to another. The work of
Eyring et al. ( 3 4 ) suggests that this term should be of an exponential form.
Two different forms were selected: A exp (Be) and C1 exp [C~h/(1 - C~¢)].
The latter form did not converge when tried in the nonlinear least squares
curve fitting procedure. However, a satisfactory fit was obtained with:
~/~0 -~ 1 + 2.5¢ -~ 10.05~ 2 -~ A exp (B~b). [9]
The coefficients A and B for this expression were 0.00273 and 16.6, re-
spectively, and the variance of fit was 0.152. Thus, with only two adjustable
coefficients it was possible to obtain as good a fit to the data with Eq. [9]
as it was with three or four adjustable coefficients in the power series form
of Eq. [4].

DISCUSSION

The reduced data for relative suspension viscosity as a function of the


concentration of uniform spherical particles, shown in Fig. 2, were deter-
mined by extrapolation of data obtained for a wide variety of particle
276 THOMAS

sizes and shear rates with the object of minimizing secondary effects such
as those due to non-Newtonian behavior, inertial forces, and measuring
instrument wall effects. None of the data from Fig. 1 which were suitable
for extrapolation were eliminated in the preparation of Fig. 2. On the other
hand, because of the requirement for extrapolation to obtain a unique set of
data in which secondary effects were minimized, the results of several very
thorough investigations could not be included. This emphasizes the im-
portance of careful planning of future experiments to insure that a suffi-
cient range of conditions is covered to permit evaluation of the secondary
effects.

ACKNOWLEDGMENT
The author wishes to acknowledge the assistance of F. L. Miller, Jr., who per-
formed the nonlinear least squares analysis.

NOTATION
E~ = viscous dissipation per unit volume, ergs/crn. 3
f = shielding factor, dimensionless.
U = velocity, cm./sec.
V = volume, elYI . a

Subscripts
, = suspension.
0 = suspending medium.
niax = m a x i m u m value.

Greek Letters
= volume fraction solids, dimensionless.
r = stress, dynes/cm. 2
= viscosity, dyne-sec./cm. 2
~/~ = mean value of relative viscosity, dimensionless.

REFERENCES
1. THOMAS, D. G., A . I . C h . E . J . 6, 631 (1960); ibid. 7, 423,431 (1961); ibid. 8, 266,
373 (1962).
2. THOMAS,D. G., I n J. F. Masi and D. H. Tsai, eds., "Progress in International
Research on Thermodynamic and Transport Properties," p. 699. Academic
Press, New York, 1962.
3. EINSTEIN, A., Ann. P h y s i k 17, 459 (1905); ibid. 19, 289, 271 (1906); ibid. 34, 591
(1911).
4. FRISCn, H. L., .~NDSIMHA,R., I n F. R. Eirich, ed., "Rheology," Vol. 1, Academic
Press, New York, 1956.
5. HERMANS, J. J., Ed., "Flow Properties of Disperse Systems." Interseience, New
York, 1953.
6. RUTGERS,I. R., Rheol. Acta 2,202, 305 (1962).
TRANSPORT CHARACTERISTICS OF SUSPENSIONS 277

7. BIRD, R. B., STEWART,W. E., AND LIGHTFOOT,E. N., "Transport Phenomena."


Wiley, New York, 1960.
8. BURGERS, J. M., Kon. Ned. Akad. Wet., Verhand. 16, [No. 4] 113 (1938).
9. HAPPEL, JOHN, J. Appl. Phys. 28, 1288 (1957).
10. GUTH, EUGENE, AND SIMHA, R., Kolloid-Z. 74, 266 (1936).
11. SAITO, N., J. Phys. Soc. (Japan) 5, 4 (1950).
12. VAND,VLADIMIR,J. Phys. & Colloid Chem. 52,277 (1948).
13. MANLEY,R. ST. J., ANDMASON, S. G., Can. J. Chem. 33,763 (1955).
14. MOONEY, MELVIN, J. Colloid Sc/. 6, 162 (1951).
15. SIMHA, ROBERT, J. Appl. Phys. 23, 1020 (1952).
16. FORD, T. F., J. Phys. Chem. 64, 1168 (1960).
17. DE BRUIJN, H., Discussions Faraday Soc. 11, 86 (1951).
18. SEGRE, G., AND SI'~EERBERG,A., J. Fluid Mech. 14, 115 (1962).
19. EVESON, G. F., "Rheology of Disperse Systems," p. 61. Pergamon Press, London,
1959.
20. SAUNDERS, F. L., J. Colloid Sci. 16, 13 (1961).
21. ROBINSON, J. V., J. Phys. & Colloid Chem. 53, 1042 (1949).
22. MARON, S. H., AND SISKO, A. W., J. Colloid Sci. 12, 99 (1957).
23. MARON, S. H., AND FOX, S. M., J. Colloid Sci. 10, 482 (1955).
24. SW~ENEY, K. H., AND GECKLER, R. D., J. Appl. Phys. 25, 1135 (1954).
25. TING, A. P., AND LUEBBERS, R. M., A.I.Ch.E.J. 3, 111 (1957).
26. WARD, S. G., AND R. L. WHITMORE, Br. J. Appl. Phys. 1,286 (1950).
27. BOUGHTON, G., AND WINDEBANK,C. S., Ind. Eng. Chem. 30, 407 (1938).
28. MARON, S. H., AND PIERCE, P. E., J. Colloid Sci. 11, 80 (1956).
29. MARON, S. H., MADOW, B. P., AND KRIEYER, I. M., J. Colloid Sc/. 6,584 (1951).
30. EILERS, H., Kolloid-Z. 97, 313 (1941).
31. WILLIAMS, P. S., J. Appl. Chem. 3, 120 (1953).
32. BANCELIN, M., Comptes Rend. 152, 1382 (1911).
33. THOMAS, D. G., Unpublished.
34. EYRING, H., HENDERSON, D., STOVER, B. J., AND EYRING, E. M., "Statistical
Mechanics and Dynamics," p. 460. Wiley, New York, 1964.
35. SIMHA, ROBERT, AND SOMCYNSKY, THOMAS, J. Colloid Sci., In the press.

S-ar putea să vă placă și