Sunteți pe pagina 1din 189

IMPERIAL COLLEGE OF SCIENCE AND TECHNOLOGY

Petroleum Engineering Section,

Department of Mineral Resources Engineering.

Mass Transfer and Interfacial Phenomena in


Oil Recovery

by

ERIC GORDON MAHERS

A thesis submitted for the degree of


Doctor of Philosophy of the University of London, and the
Diploma of Membership of the Imperial College.

August, 1983.
For Jane.
The heterogeneous structure of petroleum reservoirs results in regions of bypassed oil during
the normal practice of waterflooding, and leads to reduced oil recovery efficiency. Improved
recovery can be achieved by injecting chemicals into the oil formation, with subsequent mass
transfer between the oil ganglia and the surrounding fluid. This work has investigated the non-
equilibrium phenomena of diffusion, and capillary pressure driven interfacial and hydrodynamic
instabilities, which are important to these oil recovery processes.

Models have been constructed to enable oil to be isolated in volumes of up to several


pore diameters, and in dead-end pores. Mobilisation of oil from these stagnant regions can be
brought about by capillary pressure effects, resulting from non-uniform reduction in interfacial
tension. Such mass transfer mechanisms have been studied both at the pore scale (1 — lOOjtm)
using simulated porous media such as visual micromodels, and also at larger scales using
packed bead models and mathematical simulation.

The micromodel studies used two-dimensional pore networks constructed in transparent


materials, so that the in situ fluids could be directly observed. Networks have been produced
by etching into glass, and also by photoetching into nylon, from which replicas in epoxy resin
were cast. Computer graphics and microfilm facilities have been utilised to produce accurate
photomasks for the etching procedures. Displacement studies have used combinations of
decane, water and alcohols to represent immiscible and miscible processes. Real-time
holographic interferometry has been employed to determine fluid compositions within
micromodel pores as a function of both time and position. This technique has not previously
been used on such a microscopic level. A hologram plate holder/processing tank has been
designed and built to ensure exact superimposition of the object micromodel and the hologram
image. Mathematical simulation of diffusion between flowing and bypassed layers was
performed to scale observations to reservoir conditions.

These experiments have allowed a detailed study of mass transfer in porous media.
Instabilities during displacement processes due to capillary pressure forces, such as forward
snap-off, residual oil formation, and ganglion backflow, have been observed within micromodels
and theoretically analysed. Gravity forces have been demonstrated to play a much greater role
in the rate of transfer than previously assumed, and affect the improved oil recovery.
CONTENTS

Page No.

ABSTRACT iii
CONTENTS iv
ACKNOWLEDGEMENTS viii

Chapter 1 INTRODUCTION l
1.1 PROCESS EFFICIENCY 4
1.2 SCOPE OF THIS PROJECT 5

Chapter 2 FUNDAMENTALS OF FLOW THROUGH POROUS MEDIA 6


2.1 DARCY'S LAW 6

2.2 CAPILLARY PRESSURE 8

2.2.1 Interfacial Curvature 8

2.2.2 Contact Angle 10

2.2.3 Capillary Rise 11

2.2.4 Convergent-Divergent Capillary 13

2.2.5 Wettability and Hysteresis 13

2.3 CAPILLARY PRESSURE IN POROUS MEDIA 14

2.3.1 Fluid Saturation 14

2.3.2 Residual Oil Saturation; Capillary and Bond Numbers 15

2.4 CHARACTERISATION OF POROUS MEDIA 19

2.4.1 Laboratory Studies 19

2.4.2 Single-Phase Flow Through Heterogeneous Media 20

2.4.3 Layered Media 23

2.4.4 Two-Phase Flow 24

2.5 CHEMICAL FLOODING 25

2.5.1 Dispersion 26

2.5.2 Taylor Dispersion 27

2.5.3 Dispersion in Porous Media 31

2.5.4 Theoretical Capillary Network Model 32

2.5.5 Macroscopic Dispersion — Heterogeneous Media 35

2.5.6 Layered Media 35

2.5.7 Dispersion-Capacitance Model 38

iv
2.5.8 Dispersion in Two-Phase Flow 39

2.5.9 The Effect of Viscosity and Gravity on Dispersion 40

2.6 DISPERSION AND MICROSCOPIC RECOVERY PROCESSES 41

Chapter 3 MICROMODELS 42
3.1 PREVIOUS STUDIES 43

3.2 MICROMODEL CONSTRUCTION 44

3.2.1 Nylon Etching — Resin Micromodel 44

3.2.2 Etched Glass Micromodel 48

3.3 CAPILLARY PRESSURE IN MICROMODEL NETWORKS 51

3.3.1 Pore Shape 51

3.3.2 Surface Wettability 51

3.4 NETWORK DESIGN . 51

3.4.1 Computer Drafting 51

3.4.2 Unit Cell Design 52

3.5 MICROMODELLING TECHNIQUE 57

3.5.1 Fluid Systems 58

3.5.2 Fluid Visualisation 59

Chapter 4 NETWORK AND FLUID PROPERTIES 61


4.1 SINGLE-PHASE FLOW 61

4.1.1 Permeability 61

4.2 TWO-PHASE FLOW 63

4.3 INTERFACIAL STABILITY 65

4.3.1 Filament of Oil Surrounded by Water 65

4.3.2 Instability of an Oil Ganglion in Single Capillary Displacement 67

4.4 IMMISCIBLE DISPLACEMENT IN PORE NETWORKS 68

4.4.1 Hydrodynamic Instability — Haines Jump 68

4.4.2 Interfacial Instability — Snap-off 71

4.4.3 Stable Oil-Water Distributions 75

4.4.4 Bypassing of Wetting Phase in Drainage Displacement 76

4.4.5 Residual Oil Formation 77

4.4.6 Stability of Oil Ganglia 82

v
4.4.7 Effect of Viscosity and Density on Residual Oil Formation 84

4.5 EFFECT OF SAMPLE SIZE ON DISPLACEMENT EFFICIENCY 86

4.5.1 Edge Effects 87

4.5.2 End Effects 87

4.6 SUMMARY 89

Chapter 5 THE ROLE OF MASS TRANSFER IN TERTIARY

OIL RECOVERY 90

5.1 H Y D R O D Y N A M I C STABILITY - NON-UNIFORM I FT 92

5.1.1 Marangoni Instability 92

5.1.2 Dynamic Capillary Pressure Effects 94

5.1.3 Capillary Pressure Displacement of Bypassed Oil 98

5.2 MASS TRANSFER INTO BYPASSED ZONES 101

5.2.1 Layer Model 101

h» 5.2.2 Numerical Solution to the Layer Model 105

5.2.3 Results of the Numerical Simulation 106

5.2.4 Summary of Mass Transfer Results 115

*
Chapter 6 REAL TIME HOLOGRAPHIC INTERFEROMETRY 116

6.1 THEORETICAL BASIS OF HOLOGRAPHIC INTERFEROMETRY 116

6.1.1 The Hologram 116

6.1.2 Linear Response and Diffraction Efficiency of Photographic Materials 120

6.1.3 Holographic Interferometry 120

6.1.4 Reference Fringes 122

6.1.5 Rotatable Plate Method for Producing Reference Fringes 124

. 6.1.6 Laser Speckle 126


i-
6.2 METHOD OF HOLOGRAPHIC INTERFEROMETRY 127

6.2.1 Apparatus 127

6.2.2 Experimental Procedure 131

• 6.3 PRELIMINARY EXPERIMENTS 134

6.3.1 Wedge Model 134

6.3.2 Cover Slip Model 136

vi
Chapter 7 STUDY OF MASS TRANSFER INTO DEAD-END PORES 137
7.1 EXPERIMENTAL METHOD 137

7.1.1 The Micromodel 137

7.1.2 Fluid Properties 138

7.1.3 Experimental Procedure 141

7.2 THE MASS TRANSFER EQUATIONS 141

7.2.1 Miscible Fluids 141

7.2.2 Partially Miscible Fluids 143

7.3 MASS TRANSFER RESULTS - MISCIBLE FLUIDS 144

7.3.1 Mass Transfer Time Constant 156

7.3.2 Mass Transfer Coefficient Correlation 160

7.3.3 Effect of Orientation on the Time Constant 161

7.4 MASS TRANSFER RESULTS - PARTIALLY MISCIBLE FLUIDS 162

7.5 SUMMARY 169

Chapter 8 SUMMARY AND CONCLUSIONS 170


8.1 CONSTANT IFT DISPLACEMENT PROCESSES 170

8.1.1 Summary 170

8.1.2 Conclusions 171

8.2 VARIABLE IFT EOR PROCESSES 171

8.2.1 Experimental Studies — Summary 171

8.2.2 Conclusions 171

8.2.3 Theoretical Studies — Summary 172

8.3 MASS TRANSFER IN POROUS MEDIA 172

8.3.1 Summary 172

8.3.2 Conclusions 173

8.4 SIGNIFICANCE OF THESE STUDIES 174

8.5 SUGGESTIONS FOR FUTURE WORK 174

8.6 FINAL REMARKS 175

REFERENCES 176

vii
ACKNOWLEDGEMENTS

The author would like to express his sincere thanks to Dr. R.A. Dawe for his guidance and

encouragement throughout this work.

I am indebted to Dr. R.J. Wright for his many hours of fruitful discussion, advice and
encouragement. Thanks must also be expressed to Mr. M. Hughes for much technical help,
Dr. W.A. Wakeham, who suggested using holographic interferometry, the Physics Department
of Imperial College for their help with initial holography trials, Dr. M. Allmen for the
algorithm used in the numerical calculations described in chapter five, and
Professor C.G. Wall for his support.

I am grateful to my colleagues, past and present, for their advice and friendship:
Dr. Andy Lever, Leo Coucoulas, Mark Bilsland, Mike Wheat, Christine Gnanathurai, Dr. Prins
Casinader and Dr. Teck Choon Tan. Acknowledgement is also due to the Computing
Department of Imperial College for their help and facilities.

The author is grateful to the Science and Engineering Research Council and the Society
of Petroleum Engineers for their financial support, and the U.K. Department of Energy for
their interest.

Finally, the author would like to thank his wife, Jane, for her patience and forebearance
during these studies, and for her hard work in the preparation of much of this thesis.

E.G. Makers

viii
CHAPTER ONE

INTRODUCTION

Oil is found together with water, and sometimes gas, deep underground in porous rocks,
commonly sandstone or limestone, which are overlaid by impermeable strata. Inspection of
samples of the porous rock indicates that many have a layered nature due to the changing
environmental conditions during the formation of the rock. Faults and impermeable shale
streaks affect the fluid flow patterns in the reservoir.1 Clearly there is a range in scale from
the pores of the reservoir rock (1 — lOO/am) to the core test samples (0-01 —lm) through to
the grid block dimensions of reservoir simulators (1 —103m) and to the whole reservoir (1 —
100km), which is shown diagramatically in figure 1.1. There is thus a scale range of some
1010 over which studies have to be made and interrelated. Shale streaks and the layered
texture of the porous media are shown in figure 1.2. Petroleum reservoirs are therefore
naturally heterogeneous porous media,2,3 however most theories of flow through porous media
assume that the medium is homogeneous and uniform.4,5,6 This work emphasises the
importance of heterogeneity on the mechanisms of oil recovery.

Oil can be displaced from the porous rock of an oil reservoir by:7,8

• Using the natural reservoir pressure (natural drive-forces or reservoir energy due to
expansion of fluids within or in contact with the reservoir);
• Injection of water to sustain pressures or deliberately sweep out oil (waterflooding);
• Natural gas injection to maintain pressures or displace oil downwards.

In spite of the huge worldwide experience gained in the recovery of crude oil from
underground reservoirs over the past fifty years, oil production operations using current
engineering practices leave behind in the reservoir over half the oil originally in place.
Therefore, a number of enhanced oil recovery (EOR) processes have been proposed to improve
recovery, including:

• thermal recovery of heavy oil by steam injection or in situ combustion;


• chemical flooding, where chemicals such as surfactants and polymers are added to the
flood water;
• miscible flooding, where fluids such as hydrocarbon gases (enriched in propane and
butane), carbon dioxide and possibly nitrogen, are injected after, or in place of, flood
water. Enriched gas is miscible (often only after multiple contact) with the in situ oil
and gas, whilst carbon dioxide is soluble in both the oil and the water.

EOR techniques can be applied to either virgin or waterflooded reservoirs.

1
r Impermeable
Gas Oil Water
strata

RESERVOIR
SCALE

1 km

Cement
MICROSCOPIC
SCALE
Sand
grains
j 100 [lm

Figure 1.1 Diagramatic representation of the scale range in a petroleum reservoir.

2
Scale, 1:2

Figure 1.2 Typical samples of North Sea reservoir material (numbers represent

depth in metres).

3
1.1 Process Efficiency

The efficiency of any oil recovery process is dependent on both the macroscopic sweep
efficiency, which is the ability of the displacing fluid to reach all parts of the reservoir, and
the microscopic displacement efficiency, which is the ability of the process to mobilise oil on
the microscopic pore scale within the swept regions. The main factors are:

Macroscopic Sweep Efficiency Microscopic Displacement Efficiency


• Geometry of the reservoir and • Pore geometry (aspect ratio) and local
macroscopic heterogeneity, such as network geometry;
faulting, shale streaks and permeability
layers, and the well pattern;
• Gravity effects due to differences in # Gravity effects;
fluid densities;
• Fluid properties such as mobility ratio, • Capillary pressure (i'nterfacial tension
and capillary pressure; and wettability) and phase properties;
• Dispersion. # Diffusion and dispersion.

The macroscopic sweep efficiency has been studied elsewhere,9,10 but the mechanisms
controlling the microscopic displacement efficiency of oil recovery, especially EOR processes,
are not well understood, in particular the aspects of mass transfer as affected by pore
geometry and gravity.

Residual oil is mobilised by the transfer of chemicals from the bulk fluid to the oil
droplet interfaces in surfactant flooding, or transfer between the bulk phases with swelling or
dissolution of the oil, as in condensing and vaporising gas drives. In multi-component miscible
displacement, e.g. carbon dioxide or nitrogen flooding, pressure, temperature and mixture
composition determine whether the displacing fluid is immediately miscible with the in situ
phases or whether miscibility is reached after multiple contacts with mass transfer occurring
between the phases. The mechanisms of mass transfer are complicated as can be discerned
from the many Chemical Engineering texts, articles and conferences which have been
dedicated to the subject. 11-16 These complications are increased in petroleum reservoirs
since the processes are occurring within porous media, and then further exacerbated when the
medium is heterogeneous. In this work some of these phenomena have been experimentally
studied on the microscopic pore scale using pore network micromodels.
1.2 Scope of this Project

The objective of this study has been to gain a better understanding of the basic

parameters affecting the microscopic mechanisms of chemical and miscible EOR processes.

Specifically, this thesis reports mass transfer and interfacial effects, which are relevant to the

description of the physics of EOR methods. Novel experimental techniques were required to

measure mass transfer on the microscopic pore scale.

A summary of the theories of flow through porous media is given in chapter two. The

design, construction and use of pore network micromodels is given in chapter three and the

results of immiscible and miscible displacement processes are described in chapters four and

five. The bulk phase mass transfer of chemicals to areas of entrapped oil is also described

theoretically in chapter five and the results discussed in terms of reservoir conditions. Some of
the studies of mass transfer on the microscopic pore scale have been conducted using real

time holographic interferometry. The principles of this technique are described in chapter six

and the results given in chapter seven. In the remaining chapters these experimental and

theoretical results are discussed in the light of reservoir applications outlined in chapter five,

and the consequences for future oil recovery further considered.


CHAPTER T W O

FUNDAMENTALS OF FLOW THROUGH POROUS MEDIA

This chapter introduces the fundamental physics of flow through porous media and fluid

entrapment. The porous medium is often considered to be a continuum where a property at a

particular point is the average value over a representative volume. Thus for instance, for a

meaningful determination of porosity, the averaging volume must be larger than a single pore

and should include a sufficient number of pores to be statistically representative of the

medium.17 For heterogeneous media the value of a continuum parameter is therefore

dependent on both the averaging volume and the method of measurement. Many of the

properties will be introduced in this chapter under steady-state conditions, but their dynamics

(time or history dependence) will be discussed in later chapters.

2.1 Darcy's Law

The pore structure of reservoir rock is extremely complicated with some 109 pores in a

lcm 3 sample with a surface area of the order of lm 2 , and exact solutions of the velocity and

potential fields are currently unobtainable. It is perhaps remarkable that a simple relationship

has been found to exist between the average flow rate of a single fluid through the medium

and the applied potential gradient at low Reynolds number (Re=(pvW/p)<l, creeping flow,

where d is the average pore diameter) and is expressed by Darcy's law which, in various

forms, is

(2.1)

where Q is the volumetric flow rate, A is the bulk cross-sectional area through which the flow

rate is measured, p and p are the fluid density and viscosity, P is the pressure, g is the

vector acceleration due to gravity (positive direction upwards), ^ is the pressure potential, $ is

the fluid potential, z is height above a datum level, and k is termed the permeability which

has dimensions of L2. ( ¥ is the energy potential per unit volume, <t> is per unit mass.)
The average interstitial velocity is

v (22)
= A •
where 0 is the ratio of the pore void volume to the total bulk volume and is termed the porosity.

Therefore,
v = _ A v * 5 (2.3)
<Pn

from equations 2.1 and 2.2.

The permeability is a measure of the ability of the medium to transmit fluid, and is a
property of the medium which is assumed to be independent of the fluid. If the medium is
anisotropic, the permeability becomes a second order tensor with nine components.17

Darcy's law has been expanded, without any apparent theoretical justification, to multi-phase flow:

K
QI I
f = ( 2 A )

Z e , = Q '
i-l

where Q( is the flow rate and kt is the effective permeability of the ith phase at some saturation, and N

is the number of phases.

For many purposes the relative permeability, kri, is found to be a useful concept, where

K, = J • (2.5)

The sum of the relative permeabilities is usually less than unity, which indicates an interaction
between the individual phases. The multi-phase Darcy law has been much discussed in the
literature6 8 - 7 , 1 8 - 2 0 and forms the basis of many reservoir engineering calcu-
lation procedures.

ft

7
2.2 Capillary Pressure

2.2.1 Interfacial Curvature

There is a discontinuity in the fluid pressure at a static, curved interface due to


interfacial tension (IFT) forces; the interfacial tension for liquid-air and liquid-vapour systems
is usually termed the surface tension. Consider a surface element as shown in figure 2.1.
Balancing the forces normal to the element gives:

(PA —PB)dxdy = 2y d y s i n a + 2y dxsinfi ,

where a and /3 are small angles, PA and PB are pressures in fluids A and B, and y is the interfacial ten-

sion. Figure 2.1 shows that sin a = d x / 2 R a and sin/3 = dy/2Rp, and substitution into the above equa-

tion gives the Young-Laplace equation:

The interfacial curvature is defined as

A P

Ra are Rp are the principal radii of curvature. Figure 2.2 shows two types of surface

curvature, synclastic and anticlastic (saddle shaped). The points O a and Op for the synclastic

surface (figure 2.2a) lie on the same side of the interface, the radii Ra and Rp are positive,

and PA~PB>0. A sphere is a special case where Oa and Op are identical points in space and

Ra=Rp. The points Oa and Op for the anticlastic surface lie on opposite sides of the

interface and the two radii, Ra and Rp, have opposite sign. If Oa lies in fluid A and PA—

PB>0, then Ra is positive and Rp is negative, and Ra<—Rp. A catenoid has zero curvature

where Ra=—Rp, although the magnitude of the radii is not necessarily constant.

The difference in pressure, PA~PB, is called the capillary pressure, Pc. For an oil-water

system the capillary pressure will be defined as

p ' = P O
~ P W = Y
{~IT + = ^ = (2-8)

a p

where R is the harmonic mean radius of curvature, PQ and Pw are the oil and water pressures respec-

tively. Rv is positive if Ov lies in the oil; v = a or /3.

When the fluids are not static the pressures are replaced by the normal stresses; the

stress fields are no longer isotropic.

8
_ ! 5

Figure 2.1 Forces on a surface element.

Figure 2.2a Synclastic surface

Figure 2.2b Anticlastic surface.

Figure 2.2 Surface curvature.

9
2.2.2 Contact Angle

When a stationary oil-water interface is in contact with a solid surface, the tensions

between the oil and solid, yos, water and solid, yws, and oil and water, yow, are in

equilibrium. This is illustrated in figure 2.3, and resolving forces parallel to the solid surface

gives the classic Young equation

TOJ = Tm/j + 7 0 w cos 0 , (2.9)

where 0 is the angle between the line of action of the oil-water tension at the solid and the

solid surface, measured through the aqueous phase, and is termed the contact angle. If the

solid is totally wetted by one of the fluids then 0 = 0 and 180° for water-wet and oil-wet

respectively.

Figure 2.3 Equilibrium state at a three-phase contact.

There may be an apparent finite contact angle, 0*, when a very smooth surface is totally

wetted by one fluid in the presence of another fluid. This is because the stress field in the

thin wetting film, shown in figure 2.4, is anisotropic and the normal stress is greater than the

bulk pressure; the difference is termed the disjoining pressure.21 The curvature therefore

increases near the solid surface giving the appearance of a finite contact angle. The disjoining

pressure only becomes significant for films less than lOnm thick. When large pores are

considered, such as those between sandstone grains (pore>l/xm wide), this effect may be

neglected, although it would occur within the fine structure of clay minerals (pores<100nm

wide), particularly if surface active agents are present.22 The physics of thin films becomes

relevant when the coalescence of oil drops is considered, where the separating water film thins

as the water drains away and the two interfaces become unstable at a limiting film thickness.

The presence of surfactants stabilises the water film and increases the characteristic

coalescence time.

10
Solid

Figure 2.4 Apparent contact angle.

2.2.3 Capillary Rise

Capillarity can be illustrated by the phenomenon of capillary rise, shown in figure 2.5. If
the pressure at the bulk oil-water interface is P, then

Pw = P ~ P w g h ,

P0 = P-p0gh,

and

P 0 ~ P , = (Pw - Po)gh = *Pgh = P c . (2.10)

The interfacial curvature is given by (figure 2.5e)

r = J - = ?cos0, (2.11)

where a is the capillary radius. Substitution into equation 2.8 for Pc from equation 2.5 and T from equa-
tion 2.11, gives
Apgh = cos6 . (2.12)

The interface was assumed to be spherical, which ignores any variation in capillary pressure
along the interface due to gravity, but is justifiable for small capillaries. The error can be
estimated if we let the radius of curvature at a height h—a be R* and 0 = 0 then the
capillary pressure is

AP G ( H - A ) = £ ,

11
which with equation 2.12 gives

27
a-R* = Apg/t A pg(h-a) 1
a 2y
A pgh -f

a-R*
The error — - — « 0 for a h.

In figures 2.5c and 2.5d the contact angle, 6, is greater than 90°, cos0 < 0, and h is a negative
quantity from equation 2.12.

Oil
• Po
h
1 _

Water -
Pw

•capillary
(a) (b) (c) (d)

Figure 2.5 Capillary rise.

12
2.2.4 Convergent-Divergent Capillary

Figure 2.6 shows a circular capillary which has a linearly varying radius. The interface
has a contact angle of 0, but the interfacial curvature is given by

J - = §cos(0 + a ) , (2.13)

where Q+a is the effective contact angle, which determines the capillary pressure, a is
positive for the interface at position 1, figure 2.6, and negative for position 2. When the
interface is displaced through the tube the capillary pressure will vary periodically since a is a
function of x. If

|0 + a | > 9 O ° ,

and
\6-a\ < 9 0 ° ,

the interfacial curvature, and thus capillary pressure, will alternate positive and negative.

Figure 2.6 Convergent-divergent capillary

2.2.5 Wettability and Hysteresis

The contact angle of a particular solid-oil-water system is to a certain extent an intrinsic


value, however it also depends on the contact history, for example whether the oil or the
water first contacted the surface or whether the surface was initially clean. High energy solid
surfaces such as quartz, which has a surface charge, tend to be water wet. This is because
the water molecules are polar and prefer, energetically, to be surrounded by polar molecules
of either liquid or solid. Adsorbed molecules, especially surfactants, on the solid surface can
change the surface charge and hence the wettability.

13
When the oil-water interface is displaced over the solid surface, two contact angles are
observed: one when wetting phase is replacing the non-wetting fluid (imbibition) — advancing
— and one when wetting fluid is being replaced by the non-wetting phase (drainage) —
receding. This contact angle hysteresis is due to both the finite time required to replace
interfacial molecules of one species by those of another, and also because wettability can
depend on exposure.

2.3 Capillary Pressure in Porous Media

2.3.1 Fluid Saturation

Figure 2.7 illustrates a simple experiment to measure fluid saturation as a function of an


average capillary pressure. The sand column is initially fully saturated with the wetting phase,
water, such that the fluid levels in both the burette and the sand pack are equal to the top
of the sand. The burette is lowered slightly and the system left to reach equilibrium. The
volume of water in the sand column is calculated from conservation of mass and the water
pressure in the sand above the level in the burette (ignoring the density of air) is

Pw = (1 atmosphere) - pwgH ,

/ initial
— water
level
packed
sand

( a ) (b)

Figure 2.7 Capillary pressure — fluid saturation relationship

14
and the capillary pressure is
Pc = (l atmosphere) - Pw = pwgH .

Pc is plotted against water saturation (volume of water divided by the pore volume) in figure 2.7b. If
oil were used instead of air, the capillary pressure would be

Pc = ( p w - p 0 ) g H . (2.14)

Equation 2.14 expresses the variation in capillary pressure with height above datum in a
reservoir when both oil and water are continuous. The datum level is the position of a free
bulk interface at equilibrium with the porous medium and is the height of zero capillary
pressure.

In a water-wet reservoir the water exists as wetting films on the solid and in the crevices
at particle contacts. Reducing the volume of water at these contacts increases the interfacial
curvature and thus capillary pressure, due to the particle geometry. Above a certain height,
called the transition zone, a small decrease in saturation produces a large increase in capillary
pressure.

The water becomes less mobile as its saturation decreases, and the saturation at which it
becomes immobile is termed the irreducible water saturation. The water initially present in the
reservoir is termed "connate" water, and the initial water saturation, which may be greater
than the irreducible value, is therefore termed the connate water saturation. If the oil is
displaced by oil, the connate water remains stationary since an increase in saturation would be
accompanied by a decrease in capillary pressure. However, in a waterflood the connate water
builds up to larger saturations and becomes mobile ahead of the main displacement front.
Fluid displacement is termed drainage and imbibition when the displacing phase is non-wetting
and wetting respectively.

2.3.2 Residual Oil Saturation; Capillary and Bond Numbers

In an imbibition displacement some of the non-wetting phase is entrapped in the form of


discrete droplets, called ganglia, although in heterogeneous media these discrete volumes may
occupy many pores. A ganglion becomes entrapped when the viscous drag and pressure forces
are insufficient to overcome the resisting force of capillary pressure. This is illustrated in
figure 2.8, where Qx and 02 are tbe receding and advancing contact angles, Rx and R2 are
the downstream and upstream mean radii of curvature. Balancing the forces for a stationary
ganglion, gives
p = P + ? 7 .

P = P4 +12. (2.15)
2 Pj '
P P
2~ 3= P0^COS/3,

where 0 is the angle between the upward vertical and the vector length of the ganglion
(positive in the average flow direction), Pj, P 2 , and are pressures, and and P 2 are

mean radii of curvature as shown in figure 2.8. The ganglion is assumed to block the pore
and therefore viscous drag forces are not considered. Rearranging equation 2.15 gives

P2- P3 = POGZ*OS0 = P)-P4 + 2Y(J- - .

Figure 2.8 Entrapped oil ganglion.

16
The pressure drop in the water is

PL-P A = + PWG2COS/8 ,

where is the drop in pressure potential. Substitution for P{ — P 4 gives

~ = A * w + ApgCcos/3 . (2.16)

where Ap = pw - p0.

If there is an adjacent water pathway as shown in figure 2.8b, then A ^ may be estimated
by the Hagen-Poiseuille equation:

(2.17)

a2 A*w
v =
8p « '

where v is the average velocity in the channel, a is the channel radius, and p is the viscosity
of water. Substitution for from equation 2.17 into equation 2.16 gives

Rearranging equation 2.13 gives

l. (( -l 1- )
V/?J R2J
C= o 4 —V • (2.19)
8pv Apgcosff v
2
7a 7

If we approximate the volume of the ganglion, V, by ttrji, then

1 1
0 • 25TR 2
\RL R2J
V = W R H = — Pz — P - . (2.20)
z pv Ap a g cos ft
T 87

Let us now define two dimensionless groups, the capillary number, Nc, and the Bond number, NB,
by

uv
aq= (2.21)

Apga 2
NB = . (2.22)

17
The capillary number is the ratio of the viscous and capillary forces; the Bond number is the ratio of

the buoyancy and capillary forces. Substitution of equations 2.21 and 2.22 into equation 2.15 gives

V = - — — ( 2 23)

If the quantities in equation 2.23 are replaced by statistical averages, we may expect

S = constant
or AQ + 0 • 125 <cos/8>7Vfi '

<cos/3> = 0 for horizontal flow and correlations between Sor and Nc similar to figure 2.9 have been

published in the literature. 23-25

'or

-6 -2 N(
10 10

Figure 2.9 Capillary number — residual oil saturation relationship.

Reservoir waterflood conditions give a capillary number of around 1 0 - 7 — 1 0 - 6 , and it

has been shown that it must be increased to 1 0 - 4 — 10~ 2 in order to reduce significantly

the residual oil saturation. The only feasible method of achieving this is by reducing the

interfacial tension, 7, by three or four orders of magnitude. Certain surfactants are capable of

reducing the IFT to very low values under critical conditions of salinity, temperature, oil and

surfactant compositions.26 In miscible displacement there is no interface between two totally

miscible fluids and therefore the IFT is zero.

18
Morrow and Songkran27 have examined the relationship between the Bond number and

residual oil saturation. In their definition of the Bond number they used a particle radius

instead of a pore radius. They correlated SoR with a weighted summation of Nc and NB

which for vertical displacement was (in the notation of this thesis):

r2
Nc + 0 - 001412/Vb — , (2.25)
a2
where r is a particle radius. If we approximate r by 10a, then expression 2.25 becomes

Nc + 0- 1412NB .

For vertical flow, 0-125<cos/8>=0-125, which indicates that the assumptions contained in the

derivation of equation 2.24 are reasonable. The relationship between Sor and expression 2.25

given by Morrow and Songkran was not linear, which indicates that the numerator in equation

2.23 cannot be treated as a constant and may be expected to vary between samples.

Chatzis and Morrow 24 conducted horizontal flow experiments in water-wet sandstones,


and measured residual oil saturation at different values of the Capillary number, and by two
methods:

1. Waterflood at low Nc, increase the flow rate, and thereby Nc, and determine the new

value of Sor This method mobilises discontinuous oil.

2. Waterflood at different flow rates, and measure Sor for each value of Nc. In this
method the mobile oil is continuous.

Differences between the two Sor — Nc curves showed that the residual saturation was less for

the second method especially in the transition region A Q = 1 0 - 5 to 1 0 - 3 . This indicates a

difference between the dynamics of the two processes, which will be commented upon in

chapter four.

2.4 Characterisation of Porous Media

2.4.1 Laboratory Studies

Many laboratory displacement experiments are conducted on sandpacks and cores. Well

packed sand columns can be homogeneous, and core material is usually selected for uniform

properties such as permeability. Unfortunately these experiments are "black box", where only

the input and output data are known, although techniques using microwave absorption and

radioactive tracers can determine local saturations over small lengths of the core. From these

19
experiments, average parameters such as permeability, absolute and relative, fluid saturation

and dispersion coefficients can be calculated. If the material is isotropic, uniform and

homogeneous, then these quantities are characteristic of the porous structure and relationships

such as Sor as a function of Nc and NB are useful tools.

These experimental techniques do not allow us to look at displacement mechanisms on

the microscopic pore level, and so understand for example the mechanisms of residual oil

formation. Correct modelling of microscopic phenomena in terms of both applied and local

boundary conditions is essential when trying to model and correlate flow through the more

complicated heterogeneous media.

Reservoir media are heterogeneous and usually anisotropic, at least macroscopically, and

experiments on homogeneous materials give no direct information on the behaviour of these

particular media. Experiments on samples of reservoir rock to determine average properties do


not directly correlate with the displacement properties of larger samples when the

characteristic size of a heterogeneity is comparable to or greater than the size of the test

sample.

2.4.2 Single-Phase Flow Through Heterogeneous Media

All porous media are heterogeneous on the microscopic level, but may appear statistically

homogeneous as the sample size increases. However, many natural media have a layered

structure or contain lenses of higher or lower permeability than the surrounding matrix. Let us

consider a specimen containing a lens of low permeability, whose characteristic size is less

than, but comparable to, the sample size, and examine the results of experiments on either the

whole specimen or test samples of it. The different boundary conditions are represented in two

dimensions in figure 2.10. Figure 2.10d represents the case when many samples are taken and

would indicate the presence of the lens, and permit correct interpretation of the average

properties of the whole specimen. The boundary conditions shown in figures 2.10b and 2.10c

are different and so are the average permeabilities for the two cases. The applied and local

potential gradients are colinear and independent of y, however in 2.10c

d^/dx=constant, whilst this is not true for 2.10b. Let us first consider the boundary

conditions of 2.10b, and calculate the average permeability. Conservation of mass for steady-

state flow of incompressible fluid ensures that the volumetric flow rate is constant, and

Darcy's law (equation 2.1) gives


kx
O = H W — — — •
rr
^ " p xx
kx +
= H W — - •
u
m lL - xx -C '
i * • (2.26)

< KH >
= FJ W " i 1
n n FI L
where W is the depth of the sample and < kH ) is the average permeability. Rearranging equations
2.26 gives

~ * ^ = INVTX> (2-27a)

*(xx) - *(xx + C) = 1 , (2.27b)


2

/ uO L — x. — £
+ 8) - % = ^ ^ , (2.27c)

Adding equations 2.27a, 2.27b and 2,27c gives

r _ \T> = J . ,
+ £ ). (2-28)

1 2 H W \ kx k,

which with equation 2.27d gives


L L ~ ®+ ® (2 29)
( k H y kx k2

Thus from equation 2.29, the average permeability, <kH>, for the boundary conditions of
figure 2.10b is the length weighted harmonic mean of kx and k2.

The boundary conditions for figure 2.10c and Darcy's law give

, . kx ty - y
Q x = ( H - h ) W - j - l j - 1 , (2.30a)

k2 -

Qi = hW — , (2.306)

Qx + Q2 = H W ^ - ^ - 1 . (2.30c)
Adding equations 2.30a and 2.30b, and comparing this with equation 2.30c gives

H (kAy =(H -h)kx + hk2. (2.31)

21
Equation 2.31 states that the average permeability, ( k A >, is the width weighted arithmetic mean of
kY and k2. If the dimensions H, h, L and C are known then kx and k2 can be calculated from the two
average permeabilities.

It should be noted that the velocity and potential fields for the whole specimen, figure 1
2.10a, are two-dimensional even though the applied potential field is one-dimensional.
Assumptions of homogeneity and one-dimensionality would result in incorrect modelling of the
flow characteristics of this sample. However if H^>h and the presence of the lens has

only a limited effect on the average properties of the medium. Thus the degree of
•4
heterogeneity that is important depends on the size of the porous medium.

kj> k 2

L and H > h

(a)

i /
H >f|i k, ; k2 kl i vy2

L>je and H < h

(b)

Figures 2.10a and 2.10b

22
L < 4 and H > h

(c)

i / / / ip
wi
H ^ i It, or k2

J v t - x t '
0 L

L < <and H<h


(d)
Figure 2.10 Boundary conditions for flow through heterogeneous samples.

2.4.3 Layered Media

Reservoir materials are often layered on a macroscopic scale, and also have an apparently

laminated structure within the core plug scale. This layered texture results in anisotropy of

permeability, but when the sample size is small the boundaries can also perturb the flow field.

This is demonstrated by figure 2.11. The trivial case is when k2=0, fc^O; in figure 2.11a no

flow would be possible, whilst for figure 2.11b some flow would occur.

If the no-flow boundaries were aligned with the layering plane ( a = 0 ) or normal to the

layering (a=90°), layer width weighted arithmetic and harmonic mean permeabilities

respectively would be measured. These would be the principal permeability components of the

second order permeability tensor describing the flow through this anisotropic medium for the

case when the boundaries extend effectively to infinity.

Reservoir material also contain shale intercalations, which may or may not be correlated

over large distances. Several workers 1 - 3 ' 2 8 , 2 9 have attempted to model the behaviour of

shaly media, although generally only for single-phase flow. Shale streaks form impermeable

barriers which perturb the flow field in a manner similar to that discussed here.

23
No-flow boundaries

2.4.4 Two-Phase Flow

Figure 2.10a shows the two-dimensional, single-phase streamline pattern for a lens of

lower permeability. The streamlines for immiscible displacement are somewhat different at low 4

flow rates due to the effect of capillary pressure. In a drainage process the flow around the

lens is amplified and may totally bypass it, leaving water entrapped within.

In an imbibition displacement the greater capillary pressure driving force at the lens pulls ^

the wetting phase into it and tends to reduce the effective permeability contrast. Similarly, for

a high permeability lens, the contrast is amplified for drainage and reduced for imbibition.

These effects can be demonstrated using translucent packs of glass beads and dyes to

indicate streamlines and the displacement front.9 The permeability contrast can be created by ^

using beads of different sizes. These experiments illustrate that fluid may be trapped not only

as small ganglia, but also in larger zones within complicated heterogeneous media.

24
2.5 Chemical Flooding
In surfactant and miscible, e.g. C 0 2 , flooding the injected fluids, or components of it,

are soluble in one or both of the reservoir fluids — oil and water. It is necessary to

understand the mechanisms of fluid mixing in order to describe the physics of miscible

displacement. This is especially important for enhanced oil recovery, where the EOR

chemicals are expensive and it is desirable to inject a "small" volume, or slug (i.e. a fraction

of the reservoir pore volume) and follow this with cheaper fluid. Clearly, this fluid mixing

degrades the effectiveness of the slug as illustrated in figure 2.12.

C
Co
Crnax

Figure 2.12 Chemical slug degradation.

25
EOR chemicals are usually only effective above a critical concentration and therefore if

the maximum concentration in the slug drops below this critical value, the recovery efficiency

decreases markedly. In simplistic terms the volume of oil recovered is:

VEO = ( S o r - S o r * ) V s , (2.32)

where V S is the reservoir volume swept at a concentration above the critical value, SOR and

SOR* are the oil saturations within V S at the start of the flood and after the E O R process.

The mechanisms of oil mobilisation can also depend on the composition gradient at the

displacement front, which will be further discussed in chapter five.

2.5.1 Dispersion

Fluid mixing in pipes and porous media is termed dispersion. Dispersion is caused by

both convection and diffusion or more usually a coupling of both mechanisms. Convective

mixing in porous media occurs because there is a variation in velocity between adjacent

streamlines both within a single pore and between adjacent pores. Therefore, there is a

distribution in the residence time of molecules flowing through the medium, and marked

molecules initially contained within a small fluid element spread out to occupy an increasingly

larger fluid volume as the displacement progresses, as illustrated in figure 2.13. Diffusion

transports molecules from regions of high concentration to regions of low concentration, as

denoted by Fick's law, but diffusion also transfers molecules between streamlines, from low

velocity streamlines to ones of higher velocity and vice versa, hence the coupling between

diffusive and convective mixing, often termed "hydrodynamic" dispersion.

N m a r k e d particles

Flow
volume, 5Vj
volume, 6V2
6V2>

Figure 2.13 Hydrodynamic dispersion.

26
Pure convective dispersion (negligible diffusion) is reversible, which was demonstrated by

Hiby 30 and Heller.31 They injected dye into a transparent porous medium, commenced

displacement and observed the dispersion of the dyed fluid. The fluid flow was then reversed

and the dispersed dye was observed to reform its initial configuration. Diffusive mixing is

irreversible and therefore hydrodynamic dispersion is irreversible or only partially reversible.

The main factors affecting dispersion are

• Heterogeneities in the porous medium;

• Viscosity and density effects (dependent on composition);

• Molecular diffusivity;

• Time.

The mixing zone between two dispersing fluids grows with time irrespective of the dispersive

mechanism, but the rate of growth is dependent on the dominant mechanism, which itself may

change with time.

This discussion is not intended to be a comprehensive review of dispersion

literature,c-8'32'33 but rather an outline of the physics. It is important to note that the

mechanisms of dispersion in a laboratory core flood are not necessarily representative of the

mechanisms occurring within the reservoir, because of the variation of scale as already

illustrated in figure 1.1. There is a greater variation in residence times in heterogeneous media

due to the greater variation in permeability; this is often termed "macroscopic"

dispersion.31'34

2.5.2 Taylor Dispersion

Taylor 35 ' 36 analysed the dispersion of solute within solvent flowing through a capillary,

and found that dispersion could be described by a Fick's law type equation for low flow rates.

The diffusion-convection equation for a capillary of circular cross-section, as shown in figure

2.14, with steady laminar flow, is

(2.33)

where u is the average velocity, 2w(l—r 2 /^ 2 ) is the parabolic velocity distribution for

Poiseuille flow, c is concentration, a is the capillary radius and D is the molecular diffusivity

of the solute in the solvent. The boundary condition is

(2.34)

i.e. impermeable walls.

27
Figure 2.14 Circular capillary.

Approximate analytical solutions to equation 2.34 can be found for two particular

conditions:

1. Negligible diffusive transport.

If the initial boundary conditions for the concentration are:

c(x, t = 0) = 0 , x > 0 ,

c{x = 0, t) = cQ,

then the average concentration over the tube cross-section is

c* = cJ I - —) 0 < x <2ut;
°V 2 utJ (2.35)
c* = 0 x>2ut.

Therefore the solute is dispersed over a length 2ut, which thus grows proportionally with time.

This is shown in figure 2.15.

t= 0

ut 2ut x

(a) (b)

Figure 2.15 Dispersion with negligible diffusive transport.

2. Radial diffusion dominates

Here the time for convection to make appreciable changes in c is much greater than the time

for radial variations in c to be damped out by diffusion. The time for radial diffusion can be

estimated from equation 2.33 by setting

dx

28
i.e. ignoring axial (longitudinal) diffusion. A solution, by separation of variables, with a finite value at

r = 0, is
00

c == (2.36)
n=i

where J0 is the Bessel function of the first kind, of order zero. The boundary condition equation 2.34

ensures that
J (ajaj5) = 0,
x

where Jx is the Bessel function of the first kind, of the first order, and the first root gives

aJaxD =3-8. (2.37)

Equation 2.36 shows that c decays to 1/e of its initial value at time

1 a2
t = — = — , (2.38)
«i . 3-82z) '

from equation 2.37.

If the solute is dispersed over a tube length, lm, then the time for convection to change
significantly the concentration is of the order / m /2w, and therefore for an approximately
uniform concentration across the capillary,

^ » —-— . (2.39)
2M 3-8 2D

If the average value of c over the tube cross-section is c*, then the solution to equation 2.33 is given
by35
d2c* dc* dc*
= (2.40)
1 dx2 d x d t
where the longitudinal dispersion coefficient is

48Z>

Aris37 expanded Taylor's analysis to tubes of arbitrary cross-sectional shape and showed that
for a circular section, the inclusion of axial diffusion gives

a2 u2
= — . (2.41)

29
If the tube is initially filled with solvent ( c = 0 ) which is displaced by solvent containing
solute at a concentration c 0 , the boundary conditions for this step function are:

c(x,0) = 0; x> 0;
c(x,0) = c 0 ; x<0;
(2.42)
c(—oo,t) = c0; t > 0;
c(+oo,/) — 0 . t> 0.

The solution to equation 2.40 for an infinitely long tube with the boundary conditions of
equations 2.42, is

c 2
(2.43)
0 \ YJKLT'

where the error function,

7= f e * d i . (2.44)
yj 7t J

The form of equation 2.43 is shown in figure 2.16.

x - ut
Figure 2.16 Error function solution to equation 2.40.

The mixing zone length, l m , has often been defined as the distance between the 0 - l c 0
and 0-9c 0 points. Error function tables and equation 2.43 give

x — ut
= -0-9062,
LYJKFT
x + lm — ut
M = +0-9062,
IFKFI

and therefore

= 3-625 jKLt (2.45)

30
If x* = x(c=jc0) then t = x*/u and equation 2.45 becomes

(2.46)

Equation 2.46 shows that for a constant velocity, u, the mixing zone grows as the square root

of the mean distance travelled. Axial dispersion is irreversible and one-dimensional, since the

concentration is practically uniform over the tube cross-section.

2.5.3 Dispersion in Porous Media

Although one may visualise Taylor type dispersion within a pore or channel, flow through

porous media is much more complex. There is a velocity distribution between channels due to

both variation in conductance and the angle of the pores to the average flow direction. In a

capillary the fluid velocity vector is in one direction (parallel to the tube axis) although the

magnitude of the velocity varies with radial position, ux=ux(r). In porous media the velocity

field is three-dimensional although the average velocity may be in one direction. In

heterogeneous media, the local average interstitial velocity is a function of position in both

magnitude and direction.

Dispersion in porous media occurs not only in the average flow direction, but also in the

two orthogonal directions, which is termed lateral or transverse dispersion. Equation 2.40 is

often applied to sandpack and core displacements. Transverse gradients are ignored and it is

assumed that

c(x,y,z,t) = c*(x>t) >

where

and the average flow direction is along the x-axis. The assumption here is that either
transverse dispersion is rapid, or that convection and longitudinal dispersion are invarient over
the cross-section and therefore transverse gradients are not created. This is termed one-
dimensional dispersion.

31
Blackwell38 calculated transverse dispersion coefficients for sandpacks. He used a co-axial

dual injection arrangement to measure the amount of tracer, injected into the central part of

a sand packed column, that transferred into the annular effluent. Lateral dispersion was found

to be less than longitudinal mixing, and therefore showed that dispersion is anisotropic even in

homogeneous media.

Equation 2.40 can be extended to three dimensions to give

V-K-Vc-V-(vc) = —, (2.47)

where K is the second order dispersion tensor and v is the average interstitial velocity vector.

For homogeneous media K possesses rotational symmetry about its principal axis, which lies in

the direction of v. For uniform density and porosity, divv=0 and equation 2.47 can be

rewritten as39

(2.48)

where the axis xl lies in the direction of v, and x2 and x3 are two mutually perpendicular

axes transverse to xv

2.5.4 Theoretical Capillary Network Model

Saffman 4 0 - 4 2 used a theoretical, three-dimensional, randomly orientated and statistically

isotropic network of identical straight, circular capillaries to calculate the degree of dispersion.

He calculated the flow rate in each capillary from the Hagen-Poiseuille law (equation 2.17),

where the potential gradient was set to (d\P/dx)cos0, where d was the angle between the axis

of the tube and the average flow direction, x-axis. The longitudinal and transverse dispersions

were expressed statistically in terms of the probability distribution of the displacement of

marked particles after a given time.

Saffman considered two limiting cases for dispersion (similar to those for the single

capillary):

• Z)«(v£ where £ is the individual capillary length;41

• Z ) » v £ where diffusion is not negligible.42

32
In each tube there is Taylor dispersion provided that 9.>5a, where a is the tube radius. Both

models coupled the residence time distribution within each capillary with the randomness of

the streamlines. The velocity in each capillary is dependent on 0 and this controls the average

residence time for the tube. The dispersion in each tube governs the residence time

distribution for the capillary. The average residence time of a capillary was assumed

independent of those of adjacent tubes.

When Z)«:vfi, the mixing zone length was found to be proportional to the square root of

the average displacement, and the dispersion coefficients were proportional to the average

velocity, i.e.

/ = y/x*X constant;
171 *
x* = vt;

K l = VX constant; (2.49)

KT = v X constant;

KT « KL .

When diffusion was not negligible, the dispersion coefficients were more complicated expressions,

but at very low rates, D » vfi, they were given approximately by

(vc)
KR = D+K 1 '
l 15D '
(2.50)
(vC)2
W 1 40D •

where

D ==
D • (2.51)
tortuosity factor

The tortuosity factor in equation 2.51 is usually about 1-5. The form of equations 2.50 is the

same as that for Taylor dispersion, equation 2.41.

Hiby 30 conducted experiments on beads of glass spheres and determined an empirical


correlation for KL in terms of w, C (here the particle diameter) and D, given by

^ _ , 0 • 65v£
K L = 0 - 67D H . (2.52)
1+6-7 / A

33
Saffman's model was found to be in good agreement with equation 2.52 when £=5a. Data of
Blackwell43 and others32 for consolidated sandpacks form similar curves to equation 2.52,
although they do not all overlap. In their review, Perkins and Johnston32 suggest

KL = DE + \ • 75v£ , (2.53)

whilst Blackwell et al , 4 gave

^ = 880 ( i f f 7 - w > 0 ' 5 ' {1S4)

where <>
/ is the porosity.

Equation 2.52 for D«:v£ and equation 2.53 indicate the convective nature of dispersion in
homogeneous media for what are usually short packed beds. Here the process may not be well
modelled as a diffusion-like mechanism.44 Convection becomes even more important in
heterogeneous media.

Von Rosenberg45 conducted dispersion experiments, under equal viscosity and density
conditions, in sand columns -276, -579 and l-18m long, and at three different velocities in
each column: 6 X 10" 6 ms _ 1 (1-7 ft/day), 4-8 X 10"5ms-1 (14 ft/day) and 3-6 X
10-4ms_1 (102 ft/day). Some of Von Rosenberg's data are given in table 2.1. For the
short column, the mixing zone length (longitudinal) is almost independent of velocity, which
implies that KlQCV, and therefore convection dominated. For the longer columns KL/V shows a
velocity dependence, implying that the dispersion mechanism changes with length. It is
therefore uncertain that relationships derived from laboratory experiments hold on the scale of
the reservoir, especially since petroleum reservoirs are inhomogeneous media.

Table 2.1

Av. interstitial Column Mixing zone Kl X 10 9 £xio4

velocity length length


(ms~l) M M (m2s~l) M
6
6 X 10" 0 276 0 0392 2- 54 4-24
6
6 X 10" 0 579 0 0535 2- 26 • 3-76
6 X 10" 6 1 180 0 0673 1 75 2-92
5
4 - 8 X 10" 0 276 0 0394 20- 55 4-28

4 - 8 X 10" 5 0 579 0 0645 26 25 5-47


5
4 - 8 X 10" 1 180 0 0920 26 20 5-46

3 - 6 X 10" 4 0 276 0 0378 141 83 3-94


4
3 • 6 X 10~ 0 579 0 0695 228 55 6-35

3 - 6 X 10" 4 1 180 0 1103 282 46 7-85

34
2.5.5 Macroscopic Dispersion — Heterogeneous Media

Reservoir heterogeneities may be: random on a large scale, and may be greater than the

laboratory core size; correlated, such as layering; or ordered, such as a decrease in

permeability with distance.

Warren34 was able to reproduce core flood data by ignoring microscopic dispersion and

using a model with spacial fluctuations in permeability. He simulated the results with a three-

dimensional array of uniform, homogeneous blocks of varying permeability which were

randomly arranged. The effluent concentration was an average of the tracer effluent over the

outlet cross-section. Increasing the variance of the permeability distribution increased the

macroscopic dispersion coefficient. Heller31 showed that when microscopic (hydrodynamic)

dispersion is negligible, macroscopic dispersion is reversible; he reversed the flow in a


numerical simulator and reconstructed the initial concentration distribution.

Heller39 also performed a miscible flood in a sample of Berea sandstone (normally

considered to be homogeneous) using catalysed styrene monomer for the fluids. Clear fluid was

displaced by dyed fluid until half a pore volume had been injected, where upon flow was

stopped and the styrene allowed to polymerise. The core was then sectioned and polished.

Examination revealed flow irregularities in the form of channelling. Analysis of the effluent

concentration of this displacement using a solution to the one-dimensional dispersion equation

would not have correctly characterised the dispersion mechanism.

Random variation in permeability produces a Gaussian distribution in residence time


distribution, which can always be made to fit a dispersion law. However, the dispersion
mechanism must be correctly modelled in order to predict the dispersion behaviour at different
size and time scales.

2.5.6 Layered Media

Sandstone core material is often layered, and when these layers are correlated between
the inlet and outlet faces, the apparent dispersion coefficients can be orders of magnitude
larger than for homogeneous media. This can be illustrated using a bundle of capillaries
model. If there were no dispersion in each tube the effluent composition would reflect the
distribution in conductance of the capillaries. For n capillaries of circular section the average
effluent concentration is

35
^ttt a)uiCi
_=i
= —nn . (2-55)

/= !

where ai is the radius of the ith tube, wf is the average velocity in the capillary, and c{ is the concentra-

tion in the effluent from the ith tube.

For laminar flow, the average velocity, ut, is given by the Hagen-Poiseuille law (equation 2.17):

= a? D* '
dx

where /z »s the fluid viscosity and ^ is the pressure potential. If the viscosity is not a function of con-
DV
centration and the total potential drop is the same for each tube, then —— is identical for all capil-
dx
laries. Substitution into equation 2.17 for u( from equation 2.55 gives,

I
n
AID
'=1

I
c* = —nn • (2.56)

a?
i=i

Equation 2.56 shows that c* is a weighted mean which amplifies the contribution from the large capil-

laries. For large n, equation 2.56 can be represented by

j' n(a) a4 c{a) da

c* = ^ 00 , (2.57)

"4 da
o

where the number of channels whose radii lie between a and a + da is n(a)da, and

n(a) da — N , (2.58)

where N is the total number of channels.

36
If there is no dispersion within each capillary, then

c(a) = 0, t< — =r(a);


"M (2.59)
c{a) = c0, t> r(a),

where L is the length of the capillaries, and therefore,

00
it
n(a)a c0da

c*(t) = ^ , (2.60)
n(a)a4 da
*
o

a
t

J n(a) a4 c(a) d a = 0 ,
o

since c(a) = 0 for a < at, where L

X dx 1 '

from equation 2.17. n(a) can be chosen to fit the dispersion solution, equation 2.43 and under
these conditions the effective dispersion coefficient is not rate dependent, provided that laminar
flow conditions exist throughout. However when dispersion is included within each capillary,
equations 2.59 no longer hold and the effective dispersion coefficient would be rate dependent.

If we now introduce cross links, there will be transfer between capillaries due to lateral
dispersion, diffusive and convective. Solute is lost from the fast channels into surrounding
slower ones (where solute is less concentrated), resulting in a sharpening of the effluent
profile. This is analogous to the effect of lateral diffusion in the capillary, which produces
Taylor dispersion.

The time for lateral dispersion to damp out transverse concentration gradients is
approximately £2/I3KT, where £ is the "diffusion" distance, and 0 is a constant dependent on
the geometry. For example if fast channels are periodically interspersed with slower channels
with a period of 2£, then the analogous Taylor condition (equation 2.39) is:

7 » ^ •

where 0 = 3 • 8 2 for cylindrical geometry, and 0 = 7r2 for plane layers.

37
The right hand side of inequality 2.62 is a constant for correlated layering, but the left

hand side grows with time, or distance. Thus for a layered system, the dispersion mechanism

is initially convective, controlled by the distribution of layer permeability, and the mixing zone

length grows with distance, but is independent of the average flow rate. However as the

mixing zone length increases, transverse dispersion becomes more important, and when 4

inequality 2.62 is satisfied the dispersion becomes Taylor-like. Koonce and Blackwell46, Marie

et al.47, and Lake and Hirasaki48 have studied the effect of lateral dispersion on miscible

displacement in two-layer systems. At long times, when equation 2.62 holds (/ m is the mixing

zone length in one layer), it was concluded that lateral dispersion damps out transverse

gradients and the two-layer system behaves as a single layer, with the 0-5c 0 profile travelling

at the mean interstitial velocity, but with a larger effective dispersion coefficient than that

predicted by one-dimensional mixing theory.

Thus, a dispersion coefficient derived from experiments where inequality 2.62 is not

satisfied, does not describe the dispersion behaviour at long times when inequality 2.62 is

valid.

2.5.7 Dispersion-Capacitance Models

For a step change in concentration in a packed bed displacement, the error function

solution (equation 2.43) to the dispersion equation (equation 2.40) is symmetrical about the

mean displacement distance. However, effluent curves for consolidated cores are often, indeed

usually, asymmetric, and tail off at high concentrations. For short cores this can be explained
by the finite core length (more correct solutions to the boundary conditions are given in

references 49 and 50) or because the concentration is measured at a fixed distance rather

than at a fixed time (equation 2.45), and the mixing zone may grow significantly during the

time period of the measurements.49,51 Nevertheless, considerable asymmetry is observed even

when these two factors are negligible. Although asymmetry can be explained in terms of

heterogeneity, either layering or non-random (on the same scale of the core sample) variations

in permeability, the dispersion-capacitance model of Turner52 has been widely used to fit

effluent concentration data. The model was initially proposed to characterise chemical

engineering packed beds, especially beds of porous catalyst pellets.53 The fluid contained

within the porous particles is stagnant and mass transfer with the flowing fluid is diffusional.

Extending equation 2.48 to a medium with mobile and stagnant fluid gives

d2cf / d2cf d2cf \ dcf dcf dc


f ^ + / + 4 ) '-/ v = / -i + (1 " / ) -i • (2.63)

38
where Cj- and cs are the concentrations in the flowing and stagnant fluids respectively, / is the

volume fraction of mobile fluid, and v is the interstitial velocity averaged over the mobile

fluid volume.

The model used by Coats and Smith54 ignored any transverse concentration gradients and

equation 2.63 reduces to

d2cf dcf dcf $c

where KL* = f KL and v* = / v. v* is the interstitial velocity averaged over the total pore volume.

They modelled solute transfer into the stagnant fluid by a first order equation:

(1 - f ) ^ = M(cf-cs), (2.65)

where M is a mass transfer coefficient with dimensions of T_l. Equation 2.64 is not

necessarily a rigorous representation of the dispersion mechanisms in a laboratory core flood

experiment, and is perhaps rather more a curve fitting exercise than a characterisation of the

process. It is therefore somewhat suspect to extend laboratory dispersion data to the reservoir

scale.

2.5.8 Dispersion in Two-Phase Flow

Dispersion-capacitance models have been extended to two-phase floods, especially at

connate water or residual oil saturation® 8-55 where only one phase is mobile, and steady-state

two-phase flow.56 In an EOR flood, e.g. with carbon dioxide, of a water flooded core, there is
mixing of C 0 2 with the aqueous phase and also mass transfer with the entrapped oil.

Shearn and Wakeman57 extended equation 2.64 to include a barrier phase between the

stagnant oil and the mobile phase. The dispersion-capacitance equation would appear to be an

applicable model, however it is by no means certain that the fluid mixing is characterised by

a Fick's law-type dispersion equation and even less so by the one-dimensional version.

Nevertheless Van Deemter et a/.58 used a model similar to equations 2.64 and 2.65 with two

phases (one stagnant) to successfully explain non-ideality in chromatographic columns, in terms

of longitudinal dispersion and resistance to mass transfer.

39
2.5.9 The Effect of Viscosity and Gravity on Dispersion

Miscible displacment at unequal viscosity or density ratio is an unsteady-state


displacement; the potential and velocity fields are modified as the displacement progresses.
Unsteady flow patterns can develop when the displacing fluid is less viscous or less dense, if
the displacement is inclined upwards, than the displaced phase. This is analogous to the free
convection which develops when a more dense fluid is poured on top of a less dense one.
Under unstable conditions, peturbations in an isoconcentration surface grow with time and
fingers develop.43,59

Heller60 analysed displacement stability in terms of the family of moving isoconcentration


contour surfaces and their displacement velocities, vd, where

(V-K-Vc)Vc

where K is the symmetrical dispersion tensor. If the viscosity, p, and density, p, are functions of con-
centration only, let
Dp
p* =
~dc '
(2.67)

dc

Darcy's law (equations 2.1 and 2.3) gives

Substitution for v in equation 2.66, and taking the curl of vd gives 60

v ln
+ 7 >< k + |vcl2 (2-68)

V-AjVc
-Vl v " ) X Vc .
|v,

Equation 2.68 describes the contribution to the vorticity of the flow patterns from the fluid
properties, viscosity and density gradients (first term on the right hand side), inhomogeneities
in the texture of the porous medium (second term), and the effect of hydrodynamic dispersion
(third term).

40
If the vector (p*/p)vd + kp*g/pcj) is parallel or anti-parallel to the concentration

gradient, their cross-product is zero. When these vectors are anti-parallel the displacement is

superstable, and when they are parallel the flow is unstable, since a small angle between the

two vectors causes this angle to grow in time.

If k/<t> is constant the second term on the right hand side of equation 2.68 disappears.

When the system is superstable the fluctuations due to inhomogeneities will tend to be

damped out.

The third term in equation 2.68 represents the effect of dispersion, which tends to

stabilise the front. Heller assumed that hydrodynamic dispersion is independent of viscosity,

density and permeability effects, which combine to create macroscopic dispersion. Unstable

displacement increases the degree of macroscopic mixing.

Slug displacement efficiency in layered media has been studied by Wright 9,10 in terms

of mobility ratio (viscosity ratio for miscible displacement), and the layer permeabilities and

aspect ratios. He demonstrated that a non-unit mobility ratio changes the potential field to

produce cross-flow between the layers. For a stable viscosity ratio, the cross-flow behind the

displacement front is from the high permeability layer into the surrounding layers, which

reduces the channelling effect of the permeability contrast. For unstable displacement, cross-

flow in the opposite direction amplifies the effect of permeability heterogeneity.

2.6 Dispersion and Microscopic Recovery Processes

EOR efficiency is dependent on maintaining the maximum concentration in the slug

above a critical value, e.g. the concentration at which miscibility is reached. It is therefore

imperative that dispersion does not degrade the slug below this critical value, and slug sizes

greater than one pore volume may be required in many reservoirs. The mechanisms for

mobilising entrapped oil at the displacement front depend on the local velocity and potential

fields, and also the local concentration gradients. Concentrations averaged over a flow cross-

section may yield gradients which are less than the local gradients (as in the parallel tube

model). Transverse dispersion reduces lateral concentration gradients, and becomes more

dominant as time progresses. Therefore at long times, the dispersion may become Taylor-like,

although with a much greater effective longitudinal dispersion coefficient. Under these

conditions, the local longitudinal concentration gradient is approximately equal to the gradient

of the mean concentration (averaged over the flow cross-section). Concentration gradient

dependent recovery mechanisms are discussed in chapter five.


CHAPTER THREE

MICROMODELS

The mathematical description and prediction of fluid flow behaviour can be assisted by direct

observation. Although this is not possible in real porous media, artificial media can be made

in transparent materials. These may be glass bead packs, which are translucent unless the

refractive index (RI) of the solid particles is matched with that of the surrounding fluid, and

are very useful for observing displacements of dyed fluids on a semi-macroscopic scale. When

the refractive indices are matched, it is possible to observe residual oil ganglia within the

packing.

Monolayer bead packs, small matched refractive index models and etched networks are

usually termed "micromodels" or "micro-visual models". Here we shall define a micromodel as

a porous- medium wherein pore level events can be directly observed throughout the model.

Etched network models have been used in this work because of the advantageous control over

network design and pore geometry (size and shape). There are several ways micromodels and

their results may be used:

1. As a purely visual aid to gain insight into the physics of displacement processes within

real media;

2. Measurement of volume average properties such as fluid saturation, permeability and

dispersion coefficients, and relate these to network parameters.

3. To study pore level events, such as the mechanics of oil ganglia and fluid snap-off, in

terms of local pore topology and imposed boundary conditions (velocity and pressure

fields).

The first way has proved to be valuable and is also the forerunner of any more quantitative

studies. The essence of micromodelling is not only to seek the answers to questions of fluid

flow, but also to pose the questions which need to be answered.

It is tempting to consider the micromodel as totally representative of real media, but

although the pores are three-dimensional, the networks are not, and two-dimensional networks

can behave differently from three-dimensional ones. Bicontinual flow in two dimensions exists

only in the form of parallel channels, whereas in three dimensions intertwining pathways are

possible, e.g. in the manner of a double helix. It is therefore erroneous to measure micromodel

permeability, dispersion coefficients and fluid saturations, and to treat these as absolute

quantities which can be compared directly with values derived from real, three-dimensional

media. Measurements of, for instance, residual oil saturations can be valuable in a relative

rather than absolute sense.


In this work, controlled pore and network geometries have been used to study the pore

level physics of displacement processes and mass transfer phenomena;61'62 statistical averages

and random networks have not been used, but could be developed in any future work.

3.1 Previous Studies

Egbogah,63 Schechter64 and Scriven65 used micromodels to study the mechanics of oil

ganglion entrapment, mobilisation and movement. Egbogah and Scriven used glass bead packs

whilst Schechter used a cryolite chip packed model (the refractive index of cryolite is

approximately equal to that of water). Egbogah used monolayer bead models, which have also

been used by Chatenever66 for simple observations of displacement sequences. Scriven

matched the refractive indices of the continuous fluid and the glass. A certain amount of

information on the pore geometry is lost in matched refractive index models, except in as

much as this is reflected by the shape of the ganglia. These models were all uniformly packed

and heterogeneity effects were not studied.

Mattax and Kyte,67 Michaels et a/.,68 Davis et al.,69 Wardlaw 70 " 73 and Chatzis74

have used etched glass models. Mattax's network comprised of a rectangular array of straight

channels of similar widths but varying lengths. He used this model to study the mechanisms

of water flooding, with regard to relative permeabilities and wettability. He described the fluid

distributions and the effect of wettability on areal sweep efficiency. He clearly demonstrated

that connate water in water-wet models can become mobile ahead of the bulk displacement

front due to flow through thin wetting films. This mechanism is important for the transport of

EOR chemicals, e.g. surfactant, dissolved in the aqueous phase to oil ganglion interfaces which

are in hydrodynamically stagnant pores.

Michaels et al. used Mattax's model to analyse how changes in surface wetting might

mobilise entrapped oil. This was achieved by injecting aqueous hexylamine, and they concluded

that the observed stimulation of oil production was the result of transient changes in

wettability.

Davis et al. made use of a commercial overlay shading medium to produce an( irregular,
random micromodel network. This model was used to demonstrate qualitatively the
displacement of oil and water by the microemulsions used in the various Maraflood processes.
A film describing these displacements is available from Marathon Oil Company.
Chapter four describes the effect of pore connectivity and pore body to throat size ratio

on displacement mechanisms. This has also been studied by Wardlaw, 7 0 - 7 2 who used a

heterogeneous, rectangular network with varying pore widths.

3.2 Micromodel Construction


•4

Mattax and Kyte67 coated glass with paraffin wax and scribed straight channels in the

wax with a stylus, thus selectively exposing the surface of the glass. Davis et al.69 coated the

glass with photosensitive resist, and used a photographic masking technique similar to that

described here; Wardlaw73 and Chatzis74 have further improved glass etching techniques. In

each case the pores are formed by etching the glass away with hydrofluoric acid. 4

The photomask is a high contrast transparency of the network, where the pores are in

black and the pore walls are the transparent regions. The development of photomasks in this

work has permitted the transformation of complicated patterns into pore networks, within the ^

constraints of the etching process such as the minimum channel width which can be

satisfactorily etched. The networks can be drafted at large scale and reduced when

photographed onto high contrast line film such as Kodak Kodalith ortho film. In this work

computer graphics have been used to produce accurate network designs directly onto

microfilm. The advantages of this method are the high drafting precision and the ability to

vary readily the pore sizes by changing a data file and re-running the program.

This study has used a nylon etching technique, which has also been employed by

Lenormand,75'76 although the glass etching technique was adapted to produce the uniform

depth model for the holographic interferometry, mass transfer studies (chapters six and seven).

3.2.1 Nylon Etching — Resin Micromodel

The nylon etching method is commonly used for making printing plates and has been

further developed to suit the small scale of micromodel networks. This is a physical etching

process and the method is illustrated in figure 3.1. The photosensitive nylon (BASF Nyloprint

was used) was shielded by the photomask and exposed to ultra violet radiation, which

polymerises the nylon monomer. The unexposed regions were washed away in a bath of

ethanol and water (80 per cent, ethanol by volume) at 30 — 35 °C. The resultant pores have

44
a trapezium cross-section and steep walls. This technique produced controlled pore widths down

to at least 20/zm, and a pore depth equal to the width for widths less than 50/um and fairly

constant at 50 — lOO^m for larger pores.

Ultra Violet Light

Kill Negative

• A i r Gap

(a) ' Photosensitive Nylon

Rigid Backing Plate

(b)

Unexposed regions etched away,


and then r e - e x p o s e d t o u l t r a violet.
(c)

Silicone Rubber
• • •
(d)

I Perspex Plate

Resin
(el

Figure 3.1 Micromodel construction procedure.

To produce steep pore walls, the ultra violet light must be collimated; a double iris was
found to be adequate. The exposure time is dependent on the source and the optical

arrangement; here a Philips HPW 125W F / 7 0 / 2 bulb was used 300mm above the nylon, with

an exposure time of about 30 minutes. A strong source and short exposure time reduces the

degree of cross-linking with unexposed monomer, which reduces the pore width and increases

the pore wall angle.

The nylon polymer absorbs solvents and dyes and is not in itself a suitable micromodel
medium. Consequently silicone rubber moulds (Hopkins and William Silastic 3110 RTV and
Dow Corning Catalyst 1) of the etchings were made, from which rigid, non-absorbent epoxy
resin replicas were cast. Araldite MY753 with MY951 hardener, was found to be a suitable
resin for this purpose. Holes were drilled in the casting and tubing secured in place for the
fluid inlets and outlets. To complete a micromodel a sheet was sealed on top of the relief
structure. Epoxy resin sheet, cast in a silicone rubber mould of a microscope slide, was used
in order to preserve uniform wettability. The model was sealed by one of five methods:

45
The pores are filled with wax and then resin poured on top. The wax may be melted by

either heating the model or by exposure to infra-red radiation (from a tungsten light

bulb) and displaced by a solvent such as kerosene; a similar method was used by

Lenormand.75,76 Filling the pores, and only the pores, with wax without leaving a film

on top of the pore walls, is extremely difficult.

A temporary seal can be formed using very thin resin film and external pressure, as

illustrated in figure 3.2. However, the seal is not perfect and the film tends to depress

into the pores, resulting in an unknown and variable pore depth. This method is,

nevertheless, convenient and useful for preliminary studies.

Resin sheet forms a weak seal with the casting when they are clamped together and

placed in an oven at 50 — 60°C for one to two hours. This is not always successful, but

is more reliable when both the resin sheet and casting are freshly made. The resin

becomes more inert when cured at 50°C. Again, the resin sheet deforms and may even
block up large spaces such as entry and exit ports.

Printing plate method. The tops of the pore walls are very thinly coated with araldite

using a rubber roller, and a resin sheet pressed on top. This is more successful for large

pores and was used for the holographic interferometry model.

The method with which the author has had most success is to coat the network

perimeter with epoxy and gently press the sheet on top, ensuring that resin does not

penetrate into the pores, in order to seal the network boundaries. When this has set, a

resin emulsion is formed by mixing with ethanol (20 per cent, resin by weight) and

adding water until it becomes cloudy. The emulsion is injected into the model and left

for one or two hours. Ethanol is then flushed through the pores to remove excess resin

and leave a bonding film in the tiny crevices between the top of the pore walls and the

sheet. Air is injected into the network and the model is evacuated to remove the solvent,

and left for one to two days. The micromodel is heat treated at 50°C to increase its

inertness and is then ready for use. This, again, is a delicate technique which, although

not always successful, has produced the best working micromodels.

• j &
11 i *
i
Resin Casting
6 mm Perspex Resin Film
Plates ^
Silicone Rubber
Cushion

Figure 3.2 Micromodel seal by external pressure.

46
Figure 3.3 shows example scanning electron microscope (SEM) photographs of resin

castings. The pore depth and wall angle were calculated from stereo photograph measurements,

which gave depths up to 120Aim with pore walls inclined at 15° to the network plane normal.

J-100 p.m

Figures 3.3a, b and c.

47
j 100[lm

Figure 3.3 SEM photographs of resin castings. In figures 3.3a, b and c the resin was

freeze-fractured.

3.2.2 Etched Glass Micromodel

A glass etching technique was used to make the uniform depth micromodel for the mass

transfer studies (chapter seven). Uniformity in depth was achieved by etching into the glass

through to an inert backing and therefore the pore depth was equal to the thickness of the

glass sheet. A glass microscope cover slip was spin coated with photoresist material, exposed

to ultra violet light through a photomask, and developed to expose the parts of the glass

which were to be etched away by the hydrofluoric acid.

Procedure

A 25mm square, 115/um thick glass microscope cover slip was:

1. cleaned in acetone and baked for 30 minutes at 200°C under vacuum;

2. attached to the spin-coater using "blue-tac" (double sided tape was used for thicker

glass), as shown in figure 3.4a, and a drop of resist fluid (Micro-Image Isopoly MR450)

placed on top of the glass surface at rest. The cover slip was rapidly accelerated and

spun at 6000 r.p.m. for 10 seconds;

3. prebaked for 15 minutes at 80°C under vacuum;

48
4. covered with the photomask and placed in a vacuum frame to ensure intimate contact
between the mask and the resist, and exposed to ultra violet light for 15 minutes;

5. spray developed (figure 3.4b) in Isopoly Resist Developer;

6. spray rinsed in Isopoly Resist Rinse;

7. dried with compressed air;

8. postbaked for 30 minutes at 130°C under vacuum;

9. glued to a resin sheet with araldite;

10. immersed in hydrofluoric acid (60 per cent, by weight) for 12 minutes;

11. neutralised in aqueous sodium carbonate;

12. washed in water;

13. inlet and outlet holes drilled and tubing sealed in place;

14. resin sheet sealed on top using the roller technique, method 4 in section 3.2.1;

15. resin surfaces polished with "brasso" followed by "Duraglit".

Glass sheet
~'Blue-tac'

>
Compressed
air —
Spray
— Electric motor

| Developer

(a) (b)

Figure 3.4 Spin-coater and spray gun.

Undercutting of the resist coating was observed, which may be reduced by evaporating a
coating of copper onto the glass prior to the resist.73 The copper adheres well to the glass,
and the copper-resist adhesion is greater than the glass-resist adhesion. Although this would be
investigated in any further work, satisfactory etchings were produced by the method described
above. SEM photographs of the pores are shown in figure 3.5.

49
j 250(im

Figure 3.5 S E M photographs of the glass/Araldite model. Figure 3.5b shows severe
undercutting of the resist mask.

50
3.3 Capillary Pressure in Micromodel Networks

3.3.1 Pore Shape

The capillary pressure is given by equation 2.8:

where y is the interfacial tension, and Ra and Rp are the principal radii of curvature. The

radii Ra and Rp are controlled by the pore width, the pore depth and the contact angle. In

order to study the effect of pore shape by variation in the width, equation 2.8 shows that the

pore depth must be of the order of the pore width. In this respect the nylon etchings have
superior pore shape to thick glass etchings which tend to be shaped like a shallow T . Other

etching techniques which have been developed by the semi-conductor industry, although

capable of producing accurate network patterns, are unable to satisfy this pore width-depth

criterion.

3.3.2 Surface Wettability

Glass micromodels have a surface chemistry similar to silica sandstone, and are water-wet
when the surfaces are clean. Resin models are not strongly water-wet and show contact angle
hysteresis. Decane completely wets the resin in the presence of air, but water and air have a
finite contact angle. The contact angle for decane and water is zero for decane displacing
water, but finite for water displacing decane. For decane-water-alcohol systems the aqueous
phase tends to wet the resin completely, but not spontaneously when the surfaces are initially
contacted by decane.

3.4 Network Design

3.4.1 Computer Drafting

Computer graphics programs were written to draw precise networks directly onto
microfilm at actual size. The University of London computing and microfilm facilities were
used in conjunction with Imperial College graphics routines. The advantage of this method was

51
that similar designs, but with different pore sizes, could be drafted simply by changing the
data file and re-running the program.

The networks were in general composed of a repeated unit cell design. The unit cell
parameters governing the pore widths were set constant for the whole or for portions of the
network (e.g. figure 3.7), although random networks can be drawn by specifying different
values for each cell.

The microfilm was however found to have insufficient contrast for the etching process,
and lithographic film copies were made by contact printing.

,3.4.2 Unit Cell Design

The unit cell designs were composed of straight lines, where the end point coordinates
were defined, or circular arcs, where the circle centre, radius and two angles were defined.
The lines drawn on microfilm were transparent against a black background, and therefore the
pore walls, rather than the pores, were drawn. The width of a line was 15pm for the beam
intensity used, and blocked areas were formed by drawing parallel lines spaced 10pm apart.
The outline of the pore walls was drawn first, taking into account the finite line width, and
then shaded using the least number of lines spaced at 10pm or less. For rectangles and
triangles the filling Unes were drawn parallel to the longest side. This is illustrated in figure
3.6a where two parameters, R{ and R2, are defined. The coordinates of the corners of the
rectangles are friJ^)' (x2,y2), (x2,yl) where
xj = xq + rj + 8,
x2 = x^ — Rl — 8 ,
yl = y0 + R2 + 8,
R B
YI = Y L - 2- -
If jc 2 —^i>>'2—filling lines are drawn parallel to the x-axis and if x2—xx<y2—y^ filling
lines are drawn parallel to the y-axis. 8 is the half line width, which was set to 7-5pm
although this was defined as a variable.

Figure 3.6b shows a unit cell composed of four rectangles with three specified
parameters, R{, R2 and R2. For given values of these three parameters, the thicknesses of the
pore walls, and thereby the lengths of the throats, are determined by the size of the unit cell.
The unit cell size was set constant for the whole network, but was a variable for each design.
A network composed of this unit cell design (figure 3.6b) is shown in figure 3.7, where the
unit cell parameters, Rlt R2 and R2, were re-defined for each of the six regions.
(Xo.YL) (xL.yL)

R-

Ri [t*vy2) R,

5
1 -I
i «

t*2.yi>j

1
(x0,Yb) (xL.y0)

(a)

j
R3
t
f
R 2 — R2
1
r.

1
R3
f

(b)

Figure 3.6 Principle of the unit cell design.

53
Figure 3.7 Variation of the pore parameters (pores in white).
•4
Networks composed of triangles are shown in figures 3.8 and 3.9. The unit cell for figure

3.10 is shown in figure 3.11. Here the outline was composed of circular arcs and blocked in

with straight lines (more efficient in terms of computing time than a series of concentric arcs

and circles). .

0
Figure 3.8 Doublet network. Figure 3.9 Network with a coordination

number of four and eight.

Figure 3.12 is a dead-end pore network, figures 3.13 and 3.14 contain heterogeneities

which entrap fluid in large regions (bypassed zones) in comparison with the small residual oil

ganglion in homogeneous networks. Straight and tortuous channels without cross-links are

shown in figures 3.15 and 3.16.

54
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx
xxxxxxxxxxxxxxxxxxxxxxxxxxxxxx

E.G.MflHERS.IMPERIAL COLLEGE

Figure 3.10 Regular network of curved channels.


Figure 3.12 Dead-end pore design.

ti
¥
k
¥

in
¥
i
i
i
i

Figure 3.13 Bypass network.

Figure 3.14a Bypass network incorporating end-effect regions.

56
IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII
I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I IB IB I I I I I I I I I I
IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII
Figure 3.14b Magnification of design shown in figure 3.14a.

Figure 3.15 Straight channels. Figure 3.16 Tortuous channels.

3.5 Micromodelling Technique

The micromodels were evacuated and filled with the first fluid, except of course when

this was air. Fluids were then injected using microsyringes (e.g. 500/d) and a variable rate

syringe pump. Fluid velocities were controlled at realistic rates of less than 50mm/hr ( l - 4 m / s ,

4 ft/day). Figure 3.17 shows a simple valve and tubing (usually teflon) arrangement to control

injected fluid composition; eliminating the effects of dispersion in the tubing. The

displacements were observed through a microscope and recorded in colour on videotape or still

photographs.

Inlet 1

Figure 3.17 Valve arrangement.

57
3.5.1 Fluid Systems

Imbibition processes were studied using decane and air, whilst decane-air and decane-

water systems were used in drainage experiments. Decane-water-alcohol systems were used to

study reduced IFT and miscible processes. Various alcohols were used to cover a spectrum of

phase properties including propan-l-ol, propan-2-ol, butan-l-ol and amyl alcohol. Figure 3.18 is

the ternary phase diagram at 20 °C and atmospheric pressure for decane-water-propan-l-ol,

which was the system used in the majority of EOR experiments. Refractive index and density

were measured for compositions on the binodal curve, and used to determine the end point

compositions of the tie lines. Refractive index and density are plotted against mass fraction of

propan-l-ol in figure 3.19.

PROPAN-1 - O L

Figure 3.18 Ternary phase diagram for decane-water-propan-l-ol; the compositions are

mass fractions. The binodal curve and tie-lines were determined

experimentally.

58
M a s s F r a c t i o n of P r o p a n - 1 - o l in Fluid Mixture

Figure 3.19 Refractive index and density, both at 20 °C, along the binodal curve of
figure 3.18.

3.5.2 Fluid Visualisation

In many experiments the fluids were dyed to aid fluid visualisation. Water was dyed with
Methylene Blue and ICI Lissamines, whilst ICI Waxolines were used to dye decane. Alternate
injection of dyed and undyed, but otherwise identical, fluids highlighted the flow pathways and
clearly indicated regions of stagnant fluid, as shown in figure 3.20. However, dyes are often
surface active and may change the surface wettability, which must be taken into account
when quantifying observations. In decane-water-alcohol systems the dyes become soluble in
both phases and may also precipitate, but do not diffuse at the same rate as alcohol, water or
decane. Dyes were not used in the mass transfer studies described in chapter seven.

59
Water

Water

Figure 3.20 Doublet model (early hand drafted version of figure 3.8) initially fully
saturated with Carnation oil, which was partially displaced by water,
followed by a dyed oil flood. Subsequent injection of undyed oil
highlighted the stagnant regions.

60
CHAPTER FOUR

NETWORK AND FLUID PROPERTIES

This chapter discusses fluid flow through porous media and analyses properties, such as residual

oil formation by fluid instability, in terms of pore and network geometries; the pore aspect

ratio (ratio of body to throat size) was found to be a critical parameter. These processes were

observed experimentally using the micromodel networks which were described in chapter three.

These studies have indicated some of the essential properties of pore networks, which must be

incorporated in order to model correctly the displacement mechanisms in porous media.

4.1 Single-Phase Flow

T
Let us first consider some of the important differences between single capillary, two-

dimensional and three-dimensional networks.

4.1.1 Permeability

Consider a porous medium consisting of a bundle of identical straight, circular capillaries.

Laminar flow of Newtonian fluids in a single, uniform capillary is described by the Hagen-

Poiseuille equation (equation 2.17):

= -£T ^L
" 8M dx '

where u is the average velocity and a is the capillary radius. Comparing this equation with

Darcy's law, equation 2.3 gives:

* = (4.1)

where k is the permeability. The porosity, </>, is Nira2, where N is the number of capillaries

per unit bulk cross-sectional area. Let us now consider a two-dimensional network composed of

identical straight capillaries connected randomly, which is therefore statistically isotropic (in

two dimensions). The proportion of its pores inclined at an angle between 0 and 0 + J0 to the

x-axis (average direction of flow) will be given by:

R dd 2dd
T
(4.2)
IT
JR dd

If the potential gradient in a capillary inclined at 0 is ^ f c o s 0 , then the average velocity in this

capillary, again given by the Hagen-Poiseuille law, is

61
a2
ua = — cos 0 ,
9 8 p dx

and the component of this velocity in the x-direction is

a2 d if? 2
U = UA cos 0 = — - —cos ( (2.46)
6 8m dx

Thus the mean interstitial velocity in the two-dimensional network is

v = UJIDE
X 7T

= LL f cos Odd
47tjli dx J

A2 DIF?
(cos 2 0 + 1 )dd
8Tcp dx .

a2 d ty i

a2 d<f?
16 m dx

Comparison with Darcy's law gives the permeability as

a24>
k = (4.4)
16 '

For a three-dimensional network, as for instance used by Saffman,40 the proportion of channels

that are inclined between 0 and 0 + dd to the x-axis is

•6 = 2*-
r1 sin 0 dd d§
6=0
/•6=2ir i* | = sin Odd

I
J 6=0 J 0
rhismddddb

where r is the radial component of a spherical coordinate system (r,0,6). Thus, using equation 4.3 the
mean interstitial velocity is

a2 d ^
sin 0 cos 2ddd
8 n dx .

a2 dif? 3 i

a2 dif?
24m dx

Again comparison with Darcy's law gives


(4.5)
K 24

62
Thus in summary, one dimension:
k = ao2±' (4.1)

two dimensions:
(4.4)
K 16 '

three dimensions:
(4.5)
K 24 '

The above analysis indicates that statistical values derived from two-dimensional and three-

dimensional networks should not be directly compared. However, the two-dimensional visual

model can give valuable indications into the physics occuring in real media.

4.2 Two-Phase Flow

Many workers such as Arriola et a/. 77,78 have studied ganglion mechanics in single

capillaries containing constrictions but no cross-links. Figure 4.1 shows a capillary of circular

section, but non-uniform cross-sectional area. Since there are no cross-links, and assuming

incompressible fluid, the flow rate must be constant throughout and the Hagen-Poiseuille law

(equation 2.17) gives


7raA d V
Q = — = constant,
8 p dx

and therefore
d^ constant
—• — . (4.6)
dx a4

Thus, at a constriction the potential gradient is greater than on average. If a non-wetting oil
ganglion were trapped at a constriction, preventing flow, the pressure drop would increase until
the resisting capillary forces were overcome. In a capillary with non-circular section, such as
used by Arriola, water can flow past the ganglion, but, since the area for flow is reduced, the
velocity is increased which increases the potential drop. For this case, there are both pressure
and viscous drag forces acting on the ganglion.

a = >
'L

Figure 4.1 Circular capillary of varying radius.

63
Figure 4.2 illustrates a pore doublet where there are two alternative flow routes. From

the Hagen-Poiseuille law,

3 (4 7fl)
a - v -v - • -

& = V -t- 3
• (4 76)
-

Let

Q = Q\ + 02 = constant, (4.7c)

therefore,

02

where and <22 are the volumetric flow rates in channels 1 and 2 respectively.

If a ganglion were trapped in channel 1 the resistance to flow would increase, thus

decreasing the effective value of a{ in equation 4.7a. In order to maintain condition 4.7c more

fluid would be directed through channel 2. Increasing Q 2 results in an increase in the

potential drop ( ^ j — a s shown by equation 4.7b. Therefore even though a parallel route

exists for flow, the potential drop across the ganglion would increase, provided that the total

flow rate is maintained constant.

When many parallel routes exist, such as in a three-dimensional network with high

coordination number (number of channels that meet at a junction), the increase in potential

drop will be less than for the pore doublet. Thus an entrapped ganglion would have less effect

on the local pressure field in a network than in a single capillary.

In order to study ganglion mobilisation using single capillary models, the boundary

conditions at the constriction (point of ganglion entrapment) must be similar to those for large,

high coordination networks and real media. The important boundary condition is the pressure

gradient, and the average capillary flow rate is an insufficient condition for comparison with

real media. Flow rates much less than typical field values may be needed to meet the

64
pressure gradient requirement at the constriction. Therefore in this work, high coordination,

two-dimensional networks have been used to investigate the pore scale mechanisms of fluid

displacement, thus ensuring that the required boundary conditions of potential gradient and

fluid velocity were met. Micromodels with realistic pore sizes were used to scale correctly the

capillary pressure and viscous forces. The networks were designed to investigate particular

aspects of pore and network geometry (size and shape), for instance pore body to throat size

ratio (pore aspect ratio), but not to simulate the statistical properties of real media.

4.3 Interfacial Stability

4.3.1 Filament of Oil Surrounded by Water

Figure 4.3 illustrates the perturbation of a cylindrical oil filament surrounded by water.
The wave line is described by

(4.8)

L
Figure 4.3 Perturbed oil filament.

The surface is generated by rotating this wave line about the x-axis, which conserves the

volume of the oil. The curvature of the waveline, equation 4.8, can be obtained from

Newton's formula:

d2y
dx2
3 ' (4.9)
R

The other principal radius of curvature is R2, where

,=
2
Y
COS0 '

65
2n sin2 0
a 1—
1 _
cos2fl 0 / fly \ z
(dy
tan 0 = - ' —
cos2 0 cos 2

therefore,

m
1
cos0

1 +

which gives
1 _ cos0

The surface curvature is

R R
L 2
d2y
r = ^ — 3 +

H&7
The derivatives of equation 4.8 are given by

dy lira 2-irx
dx L L
d2y 4ic2a . 2tx
—r^ = sin .
dx2 l2 L

Substitution for the derivatives in equation 4.10 gives

t= ^ r +
(l + ^cos2^)' {R +asin^) JI + ^ C o s 2 ^

Let

= r(* -

r
2= r
(* = t)

Thus from equation 4.11

r - • 1 •
1 L2 R+a '
4T 2 a 1
r, t- +
L2 R-a

The pressures Pl and P2 are given by


R
I = rr,;

66
If Px > P2 the filament will be stable since this condition would tend to make the filament cylindri-
cal, but if Pl < P2 the perturbation will tend to grow. Thus for stability equations 4.12 and 4.13 give
47r2a 1 At1 a 1
L2 R+a L2 R-a '
2
8tr a 1 1 _ la
L2 R~a R+a R2-a2'
i.e.
4tr2 1
L2 > R2-a2 '
therefore if a <5C R 2,

L<2ttR, (4.14)

for stability.

Higher order harmonics would have wavelengths less than L, therefore expression 4.14
states that a filament is unstable if its length is greater than its circumference, a conclusion
which was reached by Plateau.79 Rayleigh80 studied the dynamics of a thin cylindrical
column of incompressible perfect liquid and found that the instability was a maximum for a
wavelength of about 1-44 X circumference. Rayleigh81 also examined the case of a filament
of viscous liquid, and found that the instability was maximum for infinite wavelengths. In
these studies Rayleigh did not consider the effect of the surrounding fluid.

Taylor82 conducted experiments on the stability of cylindrical threads formed by viscous


drag in definable shearing flow fields of another viscous liquid. He discovered that when the
two fluid viscosities were comparable, the filament became unstable when the flow was
stopped and then broke up into regularly spaced droplets. This implied that when the
surrounding fluid is also considered, there is a finite wavelength of maximum instability
corresponding to the droplet separation. Tomotika83 considered theoretically the effect of the
surrounding fluid and found that for the viscosity ratio in Taylor's experiment there was a
maximum instability for a wavelength of 1-76 X circumference. Taylor's photographs showed
the droplets to be spaced at twice the circumference, which compares satisfactorily with
Tomotika's prediction.

4.3.2 Instability of an Oil Ganglion in Single Capillary Displacement

Arriola (see section 4.2) observed that at the capillary constriction a ganglion was often
mobilised by drops snapping off at the downstream end rather than by displacement of the
whole ganglion through the throat. The instabilities occurred at the upstream side of the
constriction and can be explained in terms of the above analysis, although the particular
shearing flow field is not well defined.

67
4.4 Immiscible Displacement in Pore Networks

4.4.1 Hydrodynamic Instability — Haines Jump

Figure 4.4 illustrates the displacement of water by oil in a water wet pore. In order to

displace the interface through the constriction, termed the pore throat or neck, the resisting

capillary forces must be overcome. Assuming the end oil-water meniscus to be hemispherical

in shape, this threshold capillary pressure is given by

(4.15)
c RR '

where Rr is the average radius of curvature of the throat. Thus, oil will invade the right

hand pore of figure 4.4, when

27
(4.16)

where PQ and Pw are the oil and water pressures respectively.

/ / / / / ,
Figure 4.4 Oil invading a water-wet pore.

As oil begins to fill the pore the interfacial curvature decreases (R 2 increasing), which

reduces the capillary pressure and lowers the oil pressure in the right hand pore. The

increased pressure gradient results in a rapid invasion of the pore, termed a Haines jump. 84 A

Haines jump sequence in a micromodel network is shown in figure 4.5.

Figure 4.6 illustrates the imbibition of wetting fluid. As the interface is displaced from

point A to point B, the interfacial curvature increases, causing a reduction in the interfacial

water pressure and an increase in the oil pressure at the interface. The increasing pressure

gradients accelerate the interface; Haines jump in imbibition mode. As the interface passes

point B its curvature reduces and displacement will only continue to point C when

P < rP + — (4.17)
r o ^ w ' f?
A r
P
r >P
r
w ^ o ft

68
t 0 * 0-04 s

100 [im
t0 + 0- 08 s to- 0-12 s

Figure 4.5 Haines j u m p phenomenon; decane (light tone) invading a water (dark) filled pore.

Water

Figure 4.6 Haines j u m p in imibition mode.

69
The speed of a Haines j u m p depends on the pore size and shape, the viscosity of the

fluids and the local fluid distribution, which control the rate of supply of fluid. A Haines

j u m p sequence in imbibition mode is shown in figure 4.7.

j 250 (Im

Figures 4.7a, b, and c.

70
J 2 5 0 [im

Figure 4.7 Imbibition Haines jump; decane (wetting phase) displacing air.

4.4.2 Interfacial Instability — Snap-off

Let us now consider the stability of the oil-water interface during the drainage
displacement described in figure 4.4. Assuming that the end meniscus is hemispherical in
shape, the oil in the neck is cylindrical, and

3 ' (4.18)

then

(4.19)

71
If P5>P4 water will flow out of the neck, but if P5<P$ there will be a tendency for the water

to be forced into the neck as it is displaced by the oil entering the pore, and may result in

interfacial instability in the throat leaving a droplet of oil isolated in the right hand pore.

Thus from equation 4.19, if

P5 - P4 < 0 , R2 » 2R l' (4.20)

Haines jump with snap-off may occur, but if

P5 - P4 > 0 , R2< 2R l' (4.21)

Haines jump without snap-off takes place.

The critical value of R2/Ri depends on the pore throat shape. A value of R2/Rx = 2

was deduced by assuming that the oil filament in the neck is cylindrical, but for short throats

the filament will be anticlastic (see chapter two) and the critical value will be higher; the

critical value is approximately three for a toroidal throat shape. However in this thesis the

critical value will be taken as R2/Rx=2.

The limiting value of R2 is controlled by the size and shape of the pore body. The pore

body radius is often defined as the radius of the largest sphere which will fit into the pore

body. However, this definition is inadequate when the depth is less than the width. Thus here,

the pore body radius, RB, will be defined as the mean radius of curvature of the largest

ellipsoid which will fit into the pore body. The throat radius, R j , will be defined as the mean

radius of curvature of the largest ellipse which will pass through the neck. Therefore if

(4.22)

forward snap-off may occur during drainage. RB/RT is termed the pore aspect ratio. Snap-off

may not always occur even though expression 4.22 may be satisfied, if the Haines jump is

sufficiently rapid because expression 4.20 may only be momentarily satisfied before R2

decreases as the meniscus is forced into the next throat. Forward snap-off is demonstrated in

figure 4.8. In figure 4.8a the body radius was 75/im and the throat radius was 25fim. The

droplet radius was about 50nm, justifying expression 4.20 for this pore shape.

Figure 4.8b shows re-invasion of the pore containing the oil droplet. In this process the

droplet is deformed and the invading oil lobe is not hemispherical, and the interfacial

curvatures are higher than for initial invasion. When coalescence occurred, figure 4.8d, the

resultant lobe was not hemispherical since the volume in the pore was too large, i.e. R2 would

be greater than RB, and the capillary pressure did not drop sufficiently for expression 4.20 to

be satisfied.
to • 0 04 s

j 100 |im

to • 0- 08 s t 0 * 0-12s

Figure 4.8 Forward snap-off (light tone is decane, dark tone is water).

Figure 4.9 shows snap-off in a pore where the body radius was again 75pm, but the

throat radius was reduced to 10pm. In figure 4.9a the droplet radius was about 20pm. In this

example multiple re-invasions and snap-offs occurred since and the capillary pressure

at coalescence satisfied expression 4.20.

Snap-off, or choke off, was first observed by Roof 8 5 in a water-wet circular capillary

with a constriction. He displaced water by oil but stopped the flow a f t e r the interface had

passed through the throat. He found that a collar of water gradually formed in the neck and

eventually snap-off occurred. W h e n a groove was cut along the inside of the capillary, the

time for snap-off was greatly reduced, indicating that the process is controlled by the
86 87
resistance to the flow of water into the throat. Mohanty, Chatzis et a/., and Chatzis and

Dullien 7 4 have also discussed snap-off, but have not experimentally studied snap-off in

networks with realistic pore sizes.

73
t0 • 0-32s t0• 0-48s

Figure 4.9 Multiple re-invasion and snap-off.

The micromodel studies conducted by the author have demonstrated that snap-off can
occur during rapid Haines jumps, although a large aspect ratio does not always result in snap-
off.

74
4.4.3 Stable Oil-Water Distributions

Let us consider a cylindrical oil filament in a single capillary composed of a series of


pore throats and bodies, as shown in figure 4.10. Let the filament be perturbed sinusoidally
similar to section 4.31, where the wave line was given by equation 4.8,

Assuming this to be true for large perturbations, then snap-off would occur when a=R.
Therefore if RB<2R, snap-off would not be possible. However R<RJ, and if RB>2RT the
filament shown in figure 4.10a would be unstable provided L>2TR.

(b)

Figure 4.10 Oil filament.

It is not possible to form a filament such as that shown in figure 4.10a during high
interfacial displacement, since equation 4.16 states that a minimum capillary pressure equal to
2Y/RT is required. It was also demonstrated in the last section that although snap-offs occur
for high aspect ratio, multiple re-invasion reconnects the oil droplet until the oil volume in the
pore exceeds 4-KRB/2 and the interfacial curvature, and thereby capillary pressure, increases
and expression 4.20 is no longer satisfied. The resultant oil structure will be stable because
the constraints of pore geometry and relative fluid volumes prevent spherical drops forming.
Even if the throat length exceeds 2KRT the same considerations prevent the formation of

75
spherical droplets. At very low interfacial tension, long, thin oil filaments can be formed,
which are unstable when the length is greater than the circumference and break up into small
drops, thus forming an emulsion. However in this chapter, we shall only consider high
interfacial tension processes.

4.4.4 Bypassing of Wetting Phase in Drainage Displacement

Figure 4.11 illustrates wetting phase entrapped within a pore connecting two channels. If
at junction A the threshold capillary pressure is greater for channel 2 than for channel 1,
invasion of channel 2 will not occur. Subsequent displacement of wetting phase by oil in
channel 3, isolates the water in channel 2 provided the threshold capillary pressure has not
been reached at point A. Such bypassing is demonstrated by the micromodel network shown in
figure 4.12.

Figure 4.11 Bypassing of wetting fluid.

Figure 4.12 Entrapment of wetting fluid in a hand-drafted doublet network (light


tone is decane, dark tone is water).

76
4.4.5 Residual Oil Formation

Section 4.4.4 described the entrapment of wetting phase during drainage. Now consider a

network containing non-wetting oil in the presence of connate water. Water is injected into the

matrix to displace the oil, as illustrated by figure 4.13. Equations 4.18, 4.19, 4.20 and 4.21

are still appropriate, and snap-off will occur when either

R2 > 2RJ-2 ,

or

r 2 > 2RFI >

or
R2 > 2RT4,

etc. For example let

^b 12 ^ 7.R t 2 ,

^b 12 ^ 2 R T 2 >

but

Rb 12 ^ 2 -^r4 •

Snap-off will occur at throat 4 when R2>2RT4. Figures 4.13b and 4.13c show that for
RGY>2RT4 there is a range of positions for which R1>2RT. However, the Haines jump
analysis showed that the position shown in figure 4.12b occurs during the slow advancement,
whilst the position in figure 4.13 occurs during the rapid jump. Thus, for reservoir flow rates
snap-off occurs at the position illustrated by figure 4.12b, which has been verified by
micromodel experiments. We may also note that if RN=:RT2—RT3=RT4—RT, snap-off
occurs at throat 2 if R ^ ^ R j - , throat 3 if R ^ ^ R ^ and at throat 4 if and

the resultant ganglion will occupy a single pore body. Therefore, if the throats of a network
have equal size small ganglia will be formed in pores where the body radius is greater than
twice the throat radius, as shown in figure 4.14. Larger ganglia will be formed only when
there is a distribution of throat sizes. Figures 4.15 through 4.17 show residual non-wetting
phase formed in a micromodel network, where the pore aspect ratio was on average equal to
the critical value of two; there was a slight variance in pores size due to the etching process.

Figure 4.18 shows imbibition in a micromodel network in which the aspect ratio was less
than two. This photograph clearly shows that ganglia were not formed. An etching
imperfection produced an area of larger pore bodies within this low aspect ratio. Figure 4.19
shows that non-wetting fluid in this area was bypassed because of the lower capillary pressure
driving force. A large ganglion was formed when snap-off occurred at the downstream throats.

77
This occurred in both primary imbibition (network initially saturated with non-wetting phase)

and secondary imbibition (network containing non-wetting fluid with connate fluid). In the

primary imbibition case, wetting fluid continued to imbibe into the bypassed area and

occupied the small channels, as shown in figure 4.20. Oil can therefore be trapped in

relatively large volumes in heterogeneous media containing lenses of larger body size and low

aspect ratio.

RB > 2 RT

RB>2R T

Figure 4.13 Residual oil formation.

78
Figure 4.14 Residual oil in high aspect ratio network.

Decane

(a)

Figure 4.15 Residual oil formation in a critical aspect ratio network.

79
80
Figure 4.17 Ganglion occupying four pore bodies.

Figure 4.18 N o residual oil formed in a low aspect ratio network.

Figure 4.19 Large ganglion.

81
30 u.m Decane Solid

4.4.6 Stability of Oil Ganglia

Figure 4.21 illustrates the formation of a ganglion by snap-off at throat 3 when

Rb^2K2RY2> but Rb^2T2RJ3. I f we let R N = R T , snap-off occurs when R2 = 2RT to form the

ganglion shown in figure 4.21c. As previously discussed in section 2.2.2, the ganglion will be

trapped if the pressure field within the oil is hydrostatic, i.e.

P2~P3 = P 0 gtcos(3,

where £ is the ganglion length and ft the angle subtended by the vector ganglion length and

the upward vertical. Equation 2.16 states that

A * „ + ( Pw - Po )g£cosft - 2 7 Q- +

for stability. R{ must be reduced to RT in order to displace the ganglion through throat 3. It

has been demonstrated that R2=2RT at snap-off and if we assume that R2 remains constant

whilst R^ is reduced to Rp, then for displacement of the ganglion,

+ (Pw ~ cosft > 27 " ,

+ [pw PQ) g£ cos ft > . (4.23)

To mobilise the ganglion the magnitude of the potential gradient must be increased or the

interfacial tension reduced sufficiently to satisfy inequality 4.23. Note, however, that if

RB23>2RT or RB34>2RT the ganglion may break in two during the displacement by snap-off

at throat 3.

82
Oil
/ /
7 X

R n '
I /
/ / / X ' V , X
kB12 ' RB23 Rb34
(a)

Figure 4.21 Ganglion formation.

83
Chatzis83 conducted fluid displacement experiments in a glass micromodel consisting of a
series of pore doublets and multiplets. He showed that oil is not trapped in the large limb
during imbibition unless there is a constriction and entrapment is by snap-off. These
observations are in agreement with those of the author. The author, Chatzis,83 and Mattax
and Kyte40 have all observed the transport of wetting fluid through the thin films of the
connate phase. The author observed this transport in micromodels, and also in a larger model
filled with crushed glass, which is shown in figure 4.22.

VA ID
Bit *nlimitiltl *. "-
Outlet

Figure 4.22 Connate water bank ahead of injected water.

4.4.7 Effect of Viscosity and Density on Residual Oil Formation

Displacement of more viscous fluid by less viscous fluid, or more dense fluid by less
dense fluid in an upward direction, is unstable and fingering occurs. The probability that
larger ganglia will be formed by the bypassing and snap-off mechanism is increased. This is
demonstrated by figure 4.23, where an oil finger has formed because the body radius at point
A is larger than for the surrounding matrix. A ganglion will be formed if snap-off occurs at a
throat downstream, for example at point B. Equations 4.18 through 4.21 ignore any density
difference between the oil and water (or assume horizontal flow) and the potential gradient in

84
the bulk phases. If the potential gradient in the displacing aqueous phase and gravity are
included then (figure 4.23):

R R
27.
O2 W2 P
JV '
2

POL = ^wl + if = R W2 + V ^ • C - pwg£cos0 + ± ,

and therefore,

po2 - - - -V^h, • c + (pw •- pJgficos/5 + t - 7 (4.24)

where 0 is the angle subtended by the finger vector length and the upward vetical, and the
oil in the throat at point B is assumed to be approximately cylindrical.

Dl

Water

Figure 4.23 Oil finger.

If the oil in the finger is stationary then

P o 2 - P o i - P o g * cos/? = 0 , (4.25)

but if
P o 2 - P o l - p o g Z c o s 0 < O , (4.26)

snap-off will occur at the throat at point B to form a ganglion, and if

P o 2 - P o l - P o g Z c o s ( 3 > Q , (4.27)

snap-off will not occur and the meniscus at point A will move forward. Equation 4.24 and inequality
4.26 give

-V9W • fi + {pw - pJgficos/3 + J " - jjr ) < 0 ,

for ganglion formation, i.e.

(4.28)

85
If V\fQ,=0 and cos/3=0 inequality 4.28 reduces to expression 4.20. If V ^ = 0 but cos/3<0, i.e.

gravity unstable displacement, inequality 4.28 shows that snap-off can occur at B even when

R2<2RV Similarly, if V^w=0 and cos/3>0, snap-off is less probable. represents the

viscous losses which are reduced by decreasing the viscosity. Therefore in a low pore aspect

ratio network, unstable displacement increases the probability of ganglion formation and

thereby increases the residual oil saturation. In a high aspect ratio matrix, ganglia will

probably be formed before large fingers are formed, and before the right hand side of

inequality 4.28 becomes significant.

Increasing the flow rate in a stable displacement increases the magnitude of VSF^ and

reduces the probability of ganglion formation. This explains the anomoly in the experimental

results of Chatzis and Morrow, 24 described in section 2.3.2 and shown in figure 4.24. In these

experiments the capillary number was controlled through the flow rate, which increases the

magnitude of the term. Therefore, the physics of residual oil formation is inadequately

described by the capillary number concept.

s' o r
Discontinuous oil
0-35 H

Continuous oil

0-14
v-5 -4 -3
10 10 10'

Ne
Figure 4.24 SOR — N C relationships (after Chatzis and Morrow 24 ).

4.5 Effect of Sample Size on Displacement Efficiency

This section discusses the influence of network size on the mechanism and efficiency of

immiscible displacement at high interfacial tension, where capillary pressure dominates, in the

light of the above analysis.

86
4.5.1 Edge Effects

Edge effects were discussed in Chapter 2, section 2.3, for single phase flow in

heterogeneous media and briefly for immiscible displacement. Capillary pressure controls

displacement through a random pore network composed of. a distribution of pore sizes. For

drainage, the pores with the largest throats are entered first. Displacement continues until it is

more favourable to invade by other routes. For a thin sample, i.e. small cross-sectional area,

the freedom of choice of pathway is restricted. In an unconsolidated pack or badly fitting core

sleeve, there may be high conductance routes along the impermeable walls. In these cases

displacement would not be typical of larger samples.

4.5.2 End Effects


*

Consider a network initially saturated with water which is displaced by non-wetting oil.

At very low flow rates the applied pressure drop, AP, is accommodated by the capillary

pressure at the interfaces. If the capillary pressure at a pore throat is less than AP, the oil

0 invades the adjacent pore body. The average flow rate depends on the viscous resistance in

the oil and in the water. Thus oil will invade pores that are connected to the inlet face by

throats with threshold capillary pressures less than AP. If AP is increased more invasion

occurs until equilibrium is reached when all interfaces are at pores with threshold capillary

„ pressures greater than AP. This is one of the principles involved in porosimetry.

When oil reaches the outlet face, termed breakthrough, a continuous pathway is formed

and since there is no capillary pressure for this route, the pressure drop AP decreases. If the

pressure drop is maintained the flow rate increases until the viscous losses equate the applied

drop. However if a constant flow rate is maintained no further displacement of water takes

place after breakthrough. If the network length is large enough, the applied pressure drop may

be larger than the maximum threshold capillary pressure, even for low flow rates, and high

displacement efficiency is achieved until breakthrough. It is possible to estimate the effect of

+ network length, L, from the Hagen-Poiseuille equation (substituting A P / L for —Vq>)

_ a2 AP
V
8/1 L '

where a is the effective channel radius, and the threshold capillary pressure is

Per = 3T • (4.29)

87
If

T » (4.30)

then good displacement efficiency, at least for a certain length, is achieved!

Let us estimate values for the parameters in inequality 4.30 by:


-3
p = 10 ~3Pa.s
-5
a = 10 pm = 10~ 5 m

Rc = 5pm = 5 X 10~ 6 m
- 1
7 = 35 X 10 ~3N m~l .

Inequality 4.30 then gives

L v > 3 • 5 X 10 ~5 m2s~l .

If the velocity, v, is maintained at a typical reservoir value of 1 0 - 5 / m j - 1 , then

L > 3 • 5m .

Thus for this example, a length greater than a few metres is required for high displacement
efficiency. The length L is the end effect region and saturation measurements should only
cover the network up to a distance equal to L from the outlet face.

There is also an end effect region at the inlet face since there is fluid access to all
pores, although if AP is less than the threshold pressure for all of them some will not be
entered, which is not necessarily the case further into the network. Thus the region at the
inlet face should also be discounted from saturation measurement.

For a "black-box" core displacement, a simple mass balance does not account for end
effect regions, unless these are negligible in terms of the total size, which is not the case for
short cores. In some experiments the core is sandwiched between porous material to eliminate
this, and average saturation is determined by weighing the core (using the pore volume, and
the densities of the two fluids).

In some micromodel network designs a section with large pores at the inlet and one with
small pores at the exit have been incorporated to control drainage displacement, and vice versa
for imbibition. This was found to be essential for uniform displacement in networks with very
narrow throat size distribution. The exit section prevents premature breakthrough and reduction
in pressure drop.
A displacement experiment should therefore be conducted at fixed pressure drop rather

than fixed flow rate provided that the displacement velocity does not become too high and

affect time dependent processes. This inevitably means that long samples of natural media are

required. In micromodel networks the outlet end section which prevents breakthrough, enables

a study of the displacement efficiency in a short network provided the pressure drop is

controlled.

4.6 Summary

Micromodels have been successfully used to study imbibition and drainage displacements,

entrapment of wetting fluid, and residual oil formation. The control over pore size and shape

of the nylon etching process has enabled a critical study of the effect of pore geometry on

these processes. These experiments have verified that capillary pressure and the pore geometry

(aspect ratio) play the major role in the physics of immiscible displacement. At low aspect

ratios (less than two) residual oil is not formed. In high pore aspect ratio networks, residual

oil ganglia are formed in most pores, and tend to be small (occupying one or two pore

bodies). Larger ganglia are formed in networks where there is a distribution of pore body and

throat sizes. Heterogeneities can result in large, discrete volumes of entrapped fluid.
CHAPTER FIVE

THE ROLE OF M A S S TRANSFER IN TERTIARY OIL RECOVERY

Miscible and surfactant EOR methods rely on mass transfer to mobilise residual oil.

Mobilisation requires transport of components from the bulk fluid of one phase firstly to the

oil interface and then into the bulk fluid of the oleic phase. A very recent and thorough

review of the miscible and multicomponent processes is given by Stalkup.88 Briefly, in the

vapourising gas drive process, light hydrocarbon components transfer from the oil into the gas,

and the enriched gas may then become miscible with the oil at the displacement front.

Alternatively, in the condensing gas drive process, components transfer from the gas into the

oil, and miscibility may be achieved behind the displacement front. Carbon dioxide is soluble

in both oil and water, and has a complicated and not fully understood phase behaviour, which

is dependent on pressure, temperature and oil composition.

Currently, the task of reducing the complex mechanisms of multi-component, multi-phase

flow through porous media to a tractable problem has usually required several major

simplifying assumptions 4-6 :

• the porous medium is uniform, homogeneous and isotropic;

• hydrodynamic dispersion is negligible;

• no capillary pressure effects are present;

• local equilibrium exists everywhere.

The first assumption has been necessary because it has proved exceedingly difficult to

characterise real porous systems. Dispersion is ignored because in numerical methods,

numerical diffusion is often much greater than physical dispersion. Capillary effects have been

neglected previously, because very little information is available on how to include them when

modelling heterogeneous media; especially when mass transfer is involved. Local equilibrium is

assumed because non-equilibrium effects have not been fully characterised. Local equilibrium

can only be justified when the characteristic time for mass transfer is much less than the

characteristic time for convection to change the concentration. Local equilibrium is feasible for

small residual oil ganglia in homogeneous media,89 but a more exact description is required

when considering heterogeneous media.

In this chapter we shall consider some of the non-equilibrium effects which are involved

in the recovery of entrapped oil from a waterflooded reservoir by miscible displacement, where

the EOR chemical is soluble in both oil and water. Some of these effects are also relevant to

surfactant flooding. In particular we will discuss heterogeneity, local equilibrium and capillary

90
pressure effects. Dispersion is assumed to occur and will be expressed in terms of the mixing
zone length. Changes in the relative phase volumes will be ignored at low concentrations of
the EOR chemical, but the solute will affect the interfacial tension and thereby the capillary
pressure. The hydrodynamic stability of entrapped oil will be considered as affected by
changes in interfacial tension. Numerical methods will be used to determine diffusional mass
transfer into bypassed oil zones, which will be analysed in terms of non-equilibrium effects.
The general interaction of these mass transfer effects is summarised in figure 5.1.

HYDRODYNAMIC DISPERSION CONCENTRATION GRADIENTS & INHOMOGENEOUS


(convective and diffusive) H RATE OF CHANGE OF CONCENTRATION PERMEABILITY
1/r,

MASS TRANSFER

UNIFORM FLUID DISTRIBUTION HETEROGENEOUS

EQUILIBRIUM TIME CONSTANT

LOCAL
/
EQUILIBRIUM PHASE
INTERFACIAL TENSION
REDUCTION
\
NON-EQUILIBRIUM SPONTANEOUS
EQUIUBRtUM BEHAVIOUR — J SWELLING
EFFECTS EMULSIFICATION
« r, > T,

DISSOLUTION

CAPILLARY PRESSURE HYDRODYNAMIC INTERFACIAL TENSION


EFFECTS INSTABILITIES GRADIENTS

BREAK-UP GANGLION MOBILISATION COALESCENCE

INCREASE IN FLOWING OLEIC SATURATION

OIL BANK FORMATION

Figure 5.1

91
5.1 Hydrodynamic Stability — Non-uniform IFT

5.1.1 Marangoni Instability

Interfacial tension gradients can arise during mass transfer, due to fluctuations in solute

interfacial concentration, and also because of temperature gradients. The net force acting along

the fluid interface produces surface transport from regions of low IFT to regions of high IFT.

The movement along the interface causes movement in the bulk fluid boundary layers, setting

up convective currents.90,91 IFT gradient effects are generically termed Marangoni effects.

The Marangoni effect can be readily demonstrated by swilling brandy or whisky around in a

glass, and observing the film of liquid on the glass surface above the general level of the

liquid. The alcohol evaporates more rapidly than the water, the IFT of the film increases, and

the resultant IFT gradient produces transport into the film from the bulk liquid, resulting in

thickening of the film. As the film thickens it becomes unstable due to buoyancy forces and

flows back down the glass, appearing like tears.

Interfacial convection during diffusional mass transfer between two phases affects the rate
of transfer by changing the component of the concentration gradient normal to the interface.
Marangoni effects may cause the interface to become unstable. Sorensen91 showed that a
critical concentration gradient at the interface must be exceeded for Marangoni instability to
develop.

Figure 5.2 shows hydrodynamic and interfacial instability of an oil drop within a
micromodel pore. Mass transfer of alcohol into the drop was anisotropic and the resultant
Marangoni effect, coupled with flow within the drop due to changes in capillary pressure,
caused deformation and snap-off to occur. This occurred at a high concentration gradient at
the micromodel inlet but Marangoni effects were not seen elsewhere, although dynamic
capillary pressure effects were observed.
Figure 5.2 Marangoni effect.

93
5.1.2 Dynamic Capillary Pressure Effects

In section 4.4.5 we examined the mechanisms of residual oil formation. Equation 4.23
stated that in order to displace a ganglion either the magnitude of the potential gradient must
be increased or the interfacial tension reduced.

Let us consider the stability of a ganglion when the IFT is reduced by mass transfer of
solute to and across the oil-water interface. If this mass transfer is anisotropic the IFT will
not be uniform. Refering to figure 5.3, equation 2.11 becomes

iV
(5.1)
2 'lvl

(R{ and R2 are the mean radii of curvature of the downstream and upstream menisci). If

P2-P3-P0gtcosp = 0, (5.2)

the ganglion will be immobile, but if

P 2 - p 3 - P o 8 * cos/3>0, (5.3)

the ganglion will move forward, and if

P2-P3-poglcos(3<0, (5.4)

the ganglion will move backward.

Water

m / 1 / / / / .

Figure 5.3 Entrapped oil ganglion.

94
If 7j and y2 are reduced but 7 ! = 7 2 , inequality 5.3 is satisfied and the ganglion will
move further into the throat, which reduces the value of Rx, until a stable configuration
satisfying equation 5.2 is reached. If inequality 5.3 is still satisfied when is reduced to the
throat radius Rn, the ganglion will be displaced through the throat.

Now consider the case when 7 2 <7i and inequality 5.4 is satisfied. The ganglion will
retract from the neck which increases the value of R { and increases the right hand side of
equation 5.1. The ganglion will continue to move backward until equation 5.2 is satisfied.

The analysis of section 4.23 showed that the ganglion is formed by snap-off at throat 3
(figure 5.3), with RBl2>2RT3, i?512<2J?77, RB23<2Rn, and R2=2Rn. Backward motion of
the ganglion causes R2 to decrease, and if it is reduced to Rn, the ganglion will move
backwards through throat 1.

If the IFT difference is maintained the backward motion will persist until the values of
Rt and R2 no longer satisfy inequality 5.4. As the ganglion moves backward it will contact a
higher solute concentration, decreasing y2 which helps to perpetuate backflow. If the initial
IFT is 7j equations 5.1 and 5.2 give

but

therefore
' 2 1
^ - ^ R R W J - {5-5b)

For backflow through throat 1, equation 5.1 and inequality 5.4 give

F2 P3 Pog1 cos/3 = Pl— P4 + yk— — < 0, (5.6)


71

where R{* is the new value of the upstream radius of curvature and Rj* > Rv Substitution for Px —
P4 in equation 5.6 from equation 5.5b gives

for backflow.

RT3 < R^ < 2RT3, but is R{ » RT3 or is R{ 2 R T 3 = Rfl For example, a typical (perhaps

even high) field potential gradient is 1 psi/ft, 22 • 5 X 10 3 Pam~K If we let RT3 = 5pm, R2 = 2R T 3 =

10pm, 7 0 = 35 X 10~3 Nm~\ and the ganglion length, fi, be 50pm, then equations 5.1 and 5.2 (7! =

72 = T0) 8ive
95
Pj — 7*4 = 2 • 25 X 10 4 X 50 X 10 6 = 2y0 ( —

where Rl is the downstream mean radius of curvature in pm. Therefore,

J = 1 - 6 X 10"5,
R{ 10

J l_
Rl 10 '

and

Ry « 10 = R2

If the ganglion length were 500pm, R1 would still be approximately equal to R2.

To displace the ganglion forward by reducing the IFT (yl = y2 = y) then

PI-PA _ 27 ( \ 1 (5.8)
2 2
70 YQ\R* R*

If R2* = R2 = 2RT3 = 10pm and Rx* = RT3 = 5pm, then

1 - 6 X 10-5 = I - - L ),
To V 5 10 1

i.e.

= 1 • 6 X 10-6
To
for a ganglion 50pm long. Therefore the IFT must be reduced by six orders of magnitude for mobilisa-
tion. For a ganglion length of 500pm, the IFT must be reduced by five orders of magnitude. 22 • 5 X
10 3 Pam _ 1 is a high potential gradient, and for a lower value the IFT would have to be reduced fur-
ther still.

Letting Rx* = Rx= 2RT3 in inequality 5.7 gives

72 7i
<0,

i.e.
R
7 t > 2 R^ 7 2 . (5-9)

If we let 7, = 7 2 = 8y, inequality 5.9 becomes

(5.10)

96
noting that RT3 < RTl < 2RT3 otherwise the configuration shown in figure 5.3 would not occur. If we

let RTX = 1 • 5R T 3 , inequalities 5.9 and 5.10 give

7i > 1 '33372,

or
^ > 0-333,

for backflow.

The relationship between interfacial tension and solute concentration is dependent on the
particular fluid system, however the IFT may often be represented by an exponential function
of concentration, for y/y0 down to about 0-01 to 0-001, since

8y = ydcX constant, (5.11)

and integration gives

7 = 70exp(-ac), (5.12)

where yQ = y(c = 0) and c is the solute concentration (M L~3 or moles L~3).

If the IFT behaviour is characterised by equation 5.12 then inequality 5.9 becomes

> 2- 15
R
T\ '

where cx and c2 correspond to yx and 7 2 . If we let cx = c2 + 5c then


2Rj*3
—a 5c > In —5— ,
Ktx

(8c is a negative quantity). If RTX = 1 • 5R T 3 and a = 6 • 8/c 0 (which fits data of Morrow 92 ) then,

15c j
Vc > 0 • 04 ,
o
for backflow. Therefore if the ganglion length, 2 > 0 • 04/ ganglion backflow is probable.

Figure 5.4 shows a ganglion backflow sequence observed within a micromodel during
displacement of water by propan-l-ol, which is soluble in both the decane and the water (see
chapter three for phase properties).

The effects of backflow are:

• it may remove -a ganglion from a pore with an abnormally small throat into a channel of
more favourable aspect ratio;
• coalescence of backflowing ganglia produces larger ganglia which are easier to displace
forward;
• aids in oil bank formation.

97
Flow direction i i 250 \im

Figure 5.4 Ganglion backflow (average flow from left to right).

Under reservoir conditions the concentration gradient may be too small for backflow of
small ganglia. Backflow is, however, likely to be present in laboratory core or sandpack
experiments, where mixing zone lengths are very short.

5.1.3 Capillary Pressure Displacement of Bypassed Oil

In heterogeneous media large zones of bypassed oil remain entrapped after water flooding.
Now consider the hydrodynamic stability of such a zone during mass transfer of solute from
the surrounding mobile phase. Since the oil is initially stable there is no potential gradient
within the oil in the bypassed zone, and thus the capillary pressure compensates for the
potential gradient in the flowing fluid. If the concentration along the boundary between the
mobile and immobile zones is not constant, the IFT will not be uniform, and a potential
gradient will be present within the oil phase. The potential gradient is such that oil flows into
the regions of lowest IFT. If this flow causes the interfacial curvature (defined in chapter
two) to increase, increasing the capillary pressure, a new stable configuration may be reached.

Figure 5.5 illustrates a head meniscus at the boundary with the mobile zone. If the head
meniscus is as shown in figure 5.5a, then the capillary pressure is

Pc= |r. (5-13)

98
(a)
A

(b) ,

A
-vc

FLOW

Figure 5.5 Instability due to non-uniform IFT.

If the interfacial tension of the head meniscus changes by A7, whilst remaining unchanged
elsewhere, the resultant potential gradient pushes the meniscus into the throat until the
curvature is sufficient for stability, i.e.
-v -t- A'v 9«v
P = ->7+_£7 = £ 7 (5.14)
R RB-

Thus,
R _ 7 + A7
RZ 7

If the stability equation 5.14 requires that R < RT, a stable configuration will not be found and the
meniscus will be displaced through the throat. Therefore the condition for capillary pressure displace-
ment is
7 + A7 Ryj
7 < R b >

i.e.
. RT
(5.15)
Y RB1

(A7 is a negative quantity). Substitution into equation 5.15 for 7 from equation 5.12 gives
e-a(c+Ac) ft^

i.e.
R*P
-a Ac < In-n*- ,
KB

or
RB •
a Ac > In ,

(Ac is a positive quantity). Substituting = 2 and a = 6 • 8/c 0 gives

co 6-8
Therefore if the bypassed zone length is greater than 0- \lm capillary pressure displacement is probable.

99
Inequality 5.15 does not set a limit on the aspect ratio, R B / R T , although if RB>2RT

snap-off may occur. Micromodel examples of capillary pressure displacement from bypassed

regions are shown in figure 5.6.

Mixture of water a n d propan-1-ol j250|lm

Decane in bypassed zone


i i 100 |im

Mobile
zone

Bypassed
zone

Figure 5.6 Capillary pressure displacement from bypassed zones (decane-water-

propan-l-ol system).

100
5.2 Mass Transfer into Bypassed Zones

Section 5.3 described the hydrodynamic instability of a zone of bypassed oil under the

influence of an interfacial tension field. IFT is a function of the solute concentration, and

therefore the IFT field is related to the concentration field,, which is dependent on dispersion

in the flowing fluid and mass transfer into stagnant fluid. Mass transfer will be diffusional

until the bypassed oil becomes hydrodynamically unstable, whereupon convection will contribute

to the total mass transport. Dispersion will be characterised by the mixing zone length. This

section analyses diffusional mass transfer in bypassed zones using numerical solution methods,

and assesses the effect of mass transfer on the concentration field and thereby the probability

for the onset of hydrodynamic instability. The solution technique assumes that the mobile and

immobile zones may be treated as fluid continua. The results will be used to assess the

conditions for the onset of capillary pressure displacement of oil out of the bypassed zone.

5.2.1 Layer Model

Consider a medium consisting of layers of mobile and immobile zones, as shown in figure

5.7. A theoretical model composed of the symmetrical unit (figure 5.7b) was used to analyse

diffusional mass transfer into bypassed oil.

F y
2(W-g)
w
Oil ( • c o n n a t e w a t e r )
W-g-
Water

( • r e s i d u a l oil g a n g l i a )

0 x
L
(a) (b)

Figure 5.7 Layer model.

101
The analytical solution c(&) for diffusion into a semi-infinite slab of width 2£, for the
boundary conditions

c = 0, / = 0, -£<f<£;
(5.16)
c = cx, t > o, r = -£, r = £.

where c is the solute concentration, is11

00
(~l)w -D(2n+ l) 2 7T 2 r (2/i+l)irf
exp cos (5.17)
2/2+ 1 4£ 2

n=0

The average concentration, c*, is

c =

_8_ V 1 — D (2/2 + 1)2tt2/


= c, 1- — eXP (5.18)
IT2 Z J (2/2 +
4 l)2 4£2
n=0

and

dc* = 2c x D Y —Z) (2/2 + l) 2 x 2 /


(5.19)
dt e Z|CXP 4£2
n=0

Let us define the mass transfer coefficient, ATm, by

F=Km(cx-c*), (5.20)

where F is the molar flux per unit surface area into the slab, and by conservation of mass,

F = £— = K{c, — c*). (5.21)


q dt mV 1 ;

Ignoring terms in n greater than 0 in equations 5.18 and 5.19, and substituting for ^ and c* in equa-

tion 5.21 gives


2c, D -Dirh ^ 8c j -Dirh
exp = Kmm —rr exp — ,
K 4£2 ir2 v 4£2

or
p. 7T
" 2 D
— ~~ (5.22)
m 4 £

102
Let us also define the equilibrium time constant as

c*(/=oo)

TDC*
c*(/=0)
T.„ = (5.23)
eq c*(f=oo)
i— ^

dc*
I
c*(t=0)

or
00

f DC* ;
n d t

T„„
eq =
oo
dc* ,
t dt
dt

Equation 5.19 gives


00 00
2D f V ~D(2N+ \)VT ,

0 71 = 0
32 £2
-g^Xl-OlS,

and therefore,
1 £2
T = - It . (5.24)
eq ^ D

We now have a measure of the rate of mass transfer and time for transverse equilibrium,
through Km and r , for the layered model of figure 5.7 for the particular boundary
conditions
/) = 0 0 <x < L , 0 < y < W , t < 0;
c{x,y, t) = cQ 0 <x <L , 0 <y < W t > 0;
c{x,y, /) = 0 0 < a: < l W-%<y <W .
However we are interested in examining mass transfer into a stagnant zone within the
macroscopic medium of the reservoir where solute dispersion may produce a mixing zone
length larger than the bypassed zone length, L. It is inappropriate to introduce a step function
of solute concentration at *=0, therefore ramp and complementary error function
concentration profiles were injected. Since both profiles were similar at low concentrations,
only ramp profile will be considered here. The ramp profile is shown in figure 5.8. The
relevant dimensionless groups are

103
W '

r = — •
(5.25)
dX = £v_
dT LD 5

c = — •
^ c '
max
L =
x(C = 0 • l) — x(C =0-9) /„
L

where v is the magnitude of the average interstitial velocity in the mobile layer, cmax is the
maximum solute concentration in the aqueous phase, and D is the effective diffusivity in the
oil zone.

C(X=0) <

Let us now define a characteristic time, 7y, for convection to change the solute concentration,

7 = v ; (5.26)

Ty is the convection time constant. One may expect transverse equilibrium to be achieved approximately

when,

7 » >

or
- 1I7T
f »
i, (5.27)
3D
by equations 5.24 and 5.26. Rearranging equation 5.27, and substituting from equations 5.25, gives

1 £2v L
1» - s
3 LD i '
m

LM»~ 1—
dX > (5-28)
3 dT
for local equilibrium.

104
Local equilibrium means transverse equilibrium i.e. c{x,y,t)=c(x,t), but if / m « L and
TJ»T t capillary pressure displacement may occur because longitudinal equilibrium is not
present. However we shall restrict the analysis to the case when lm^>L and consider the effect
of mass transfer on the solute concentration field.

5.2.2 Numerical Solution to the Layer Model

Solute transport within the mobile zone was by convection (fluid flow) in the longitudinal
direction (x-axis), and diffusional in the transverse direction. Within the immobile zone
transport was diffusional and only in the transverse direction. Longitudinal diffusion was
neglected in both zones, which is justified when The solid structure of the layers was

neglected, and each zone was considered as a continuum fluid so that the diffusion coefficients
were effective diffusivities. The results were computed using a finite difference algorithm.

The finite difference mesh used was 40 transverse points by 50 longitudinal nodes. At
each convection time step the nodal concentrations in the mobile zone were moved along one
node, the values at x = 1 . 0 discarded and new values allocated at x = 0 according to the rules
of the injected concentration profile. After each convection step, transverse diffusion was
allowed (ten time-steps were found to be adequate) and then convection was again permitted
in the mobile region.

In all the simulations certain parameters were kept constant:

^ =14-4; (5.29)

(x(C = 0) — x(C = 1) = 181).

Although the diffusivities in each zone could be made different, they were set equal. The
solubility of solute in the aqueous and oleic phases was allowed to vary but the relative
solubility was assumed to be independent of concentration and characterised by an equilibrium
partition coefficient defined As

O ,
MM — -PR- (5.30)

where C0 and Cw are the dimensionless concentrations in the oil and water respectively at equilibrium.
Equilibrium was assumed at the boundary, y — W — and

C0(y= W-Z) = mCw(y= W-Q

Three partition coefficients were examined, 0 - 5 , 1 - 0 and 2 • 0.

105
Two values of the dimensionless speed, - j f , were used:

dX 5 , , ,
= - = 1 . 667 ; case 1
dT 3
dX 125
= = 41 • 667 . easel
dT 3

i.e.

1 //V
- — = 0 • 555 « 14 • 4 = L ; case 1
3 dT
1 z/y
- — = 13-888 « 14-4 = L m , easel
3 dT

Thus case 1 satisfies the condition for transverse equilibrium, inequality 5.28, whilst case 2
examines this criterion by setting Teq=z.Tf.

To put these dimensionless groups in perspective case 1 is equivalent to


D = lO^mV1;
v = 1 • 04 X 1 0 - 5 ;
| = 2MM;
L = 25 mm ;
W = 8 mm ;
lm = 0 - 3 6 m .

If the other parameters are constant case 2 is equivalent to £=10mm and W= 40mm.

5.2.3 Results of the Numerical Simulation

A mass transfer coefficient was calculated at X = 0 from equation 5.21, by

and the dimensionless mass transfer coefficient, A, was defined by

= Ay • (5.31)

Equation 5.22 gave A=0-257r 2 =2-47 for a constant boundary concentration at y=W—
Initially A was high but decreased with time to an asymptotic value, given in Table 5.1 and
the rate of approach is indicated by the time at which A was within 10 per cent of the
asymptotic value.

The IFT was estimated from equation 5.12, setting

7 = e"6'8S (5-32)

106
where here y is the IFT normalised with respect to the IFT at zero solute concentration. The
value of a = 6 . 8 fitted experimental data of Morrow92 for a brine-iso-octane-propan-2-ol system,
and aqueous alcohol concentrations up to 60 per cent, by volume.

Table 5.1 Mass transfer coefficient for ramp concentration profile at X = 0 .

m A

0-5 2-52 for T > 1-8


within + 1 0 % for T > 0-5

1-0 2-14 for T > 2 - 0


within + 1 0 % for T > 0-4

2-0 1-645 for T > 2 - 2


within + 1 0 % for T > 0-37

The concentration data were plotted at time intervals equivalent to L/v, i.e. AT = (jf-) •
Figures 5.9 through 5.14 show selected data for the two cases studied.

The simulator was also run for the case when transverse equilibrium is a good
approximation, r e q «.rp with dX/dT— 0-00167. Figure 5.15 shows that the isoconcentration
profiles are normal to the flow direction, which was expected for transverse equilibrium.
However, figure 5.16 shows that the mass transfer increased the magnitude of the
concentration gradient by a factor of W/(W—£)\ the concentration drop, AC, along the layer
boundary was increased from 0-055 to 0-075 approximately. For case 1, Teq<Tp AC was
0-115, whilst for case 2, t carp AC was 0-085. These values of AC show that the boundary
concentration drop is greater for case 1 when transverse equilibrium may be expected, than
for case 2, the non-equilibrium case. This is because there is almost transverse equilibrium at
X=0, but not downstream. The concentration drop will be maximised when the transverse
equilibrium time constant is equal to the time for flow to transport solute from X=0 to
X=L, i.e. when

T v L > x1
'eq v T ' m '
m

Substitution for r eq from equation 5.24 gives

= K
3 D v'

and from equation 5.25, the dimensionless speed is

dX _ Z2v
= 3. (5.33)
dT LD

The dimensionless speed for case 1 was 1 • 667.

107
RAMP INJECTION
PARTITION COEFFICIENT = 1 - 0 0 0 0 TIME=2.40000

FRACTIONAL DISTANCE IN FLOW DIRECTION

Figure 5.9 Isoconcentration curves, case 1.

RAMP INJECTION
TIME=0.60000

FRACTIONAL DISTANCE, X

Figure 5.10a

108
RAMP INJECTION
TIME=1 .20000

FRACTIONAL DISTANCE, X
Figure 5.10b

RAMP INJECTION
TIME=2.40000

FRACTIONAL DISTANCE, X

Figure 5.10c
Figure 5.10 Boundary concentration profiles, case 1.

109
RAMP INJECTION
PARTITION COEFFICIENT = 1 . 0 0 0 0 TIME = 2.40000

FRACTIONAL DISTANCE, X

Figure 5.11a Concentration profiles, case 1.

RAMP INJECTION
PARTITION COEFFICIENT = 1 .00 a =6.80 TIME = 2 . 4 0 0 0 0
1
1 -0 1— 1 —r

— Interfociol Tension Along Y»0


— Interfaciol Tension Along Boundory
Interfociol Tension Along Y»1

0 o.l 0.2 0.3 0.4 0.5 0-6 0.7 0-8 0-9 1.0
FRACTIONAL DISTANCE, X

Figure 5.11b IFT profiles (equation 5.32), case 1.

110
RAMP INJECTION
PARTITION COEFFICIENT = 1 . 0 0 0 0 TIME=0-09600

Immobile

Mobil*

.0 0.1 0.2 0.3 0.4 0.5 0.6 0-7 0.8 0.9 1.0
FRACTIONAL DISTANCE IN FLOW DIRECTION

Figure 5.12 Isoconcentration curves, case 2.

* RAMP INJECTION
T If1E=0 .02400

FRACTIONAL DISTANCE. X

4
Figure 5.13a

111
RAMP INJECTION
TIME=0.04800

FRACTIONAL DISTANCE, X

Figure 5.13b

RAMP INJECTION
TIME=0.09600

FRACTIONAL DISTANCE, X

Figure 5.13c

Figure 5.13 Boundary concentration profiles, case 2.

112
• RAMP INJECTION
PARTITION COEFFICIENT = 1 - 0 0 0 0 TIME = 0 - 0 9 6 0 0

°°§.0 0-1 0.2 0.3 0.4 0.5 0-6 0.7 0.8 0.9 1.0

FRACTIONAL DISTANCE, X

Figure 5.14a Concentration profiles, case 2.

RAMP INJECTION
PARTITION COEFFICIENT = 1 . 0 0 a = 6-80 TIME = 0 - 0 9 6 0 0
1 ,q ^ —-l — y 1 | 1 1 1 1 r

0.0 o.l 0-2 0.3 0.4 0.5 0.6 0-7 0 0.9 1.0
FRACTIONAL DISTANCE, X

Figure 5.14b IFT profiles (equation 5.32), case 2.

113
RAMP INJECTION
PARTITION COEFFICIENT = 1 . 0 0 0 0 TIME = 240 . 0 0 0 0 0
1 1
1 1 1 1 1 1—'—1—'—1 —1—'—1—'—1 1
Immobile

Mobile

n . I I I. 1
I . I .1 I I. J I1 .' I1. I I—L.i
I .!_, 11 .
0-0 0-1 0.2 0.3 0.4 0.5 0-6 0.7 0-8 0.9 1.0
FRACTIONAL DISTANCE IN FLOW DIRECTION

Figure 5.15 Isoconcentration curves, transverse equilibrium.

RAMP INJECTION
PARTITION COEFFICIENT = 1 - 0 0 0 0 TIME = 240.00000

0.250

0.225

0.200

0.175
O
S 0.150

£
^ 0.125
g 0.100
o
CD

0 .075

0.050

0.025

°-°°S.O 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
FRACTIONAL DISTANCE, X

Figure 5.16 Concentration profiles, transverse equilibrium.

114
5.2.4 Summary of Mass Transfer Studies

Capillary pressure displacement of oil from a bypassed zone occurs because longitudinal

equilibrium does not exist and there is an IFT gradient along the boundary with the mobile

zone. For longitudinal equilibrium to be approximated the mixing zone length would need to

be many times longer than the length of the bypassed layer. The probability of hydrodynamic

instability is increased by transverse mass transfer, which increases the concentration drop

along the boundary between the zones. When transverse equilibrium is a good approximation,

Teq<s.Tj-, the relative layer thickness controls the increase in the magnitude of the concentration

gradient. The results show that the probability of hydrodynamic instability is not greatest when

transverse equilibrium is not a valid assumption, but is a maximum for

When T eq =Tj, transverse equilibrium is approximated at the upstream end ( Y = 0 ) , but not

downstream because mass transfer has increased the magnitude of the concentration gradient.

For this maximum case the relative thickness of the mobile layer does not influence the

concentration drop when W > i 4 l as shown by figure 5.9.

115
CHAPTER SIX

REAL T I M E H O L O G R A P H I C INTERFEROMETRY

Real time (live fringe) holographic interferometry was used to determine fluid composition
within micromodel pore networks as a function of both time and position, and has enabled
quantitative assessment of mass transfer between miscible fluids. Although measurements of
diffusion coefficients by this method have been reported, for instance the diffusivity of
potassium chloride in water,93 this technique has not been used previously on such a
microscopic scale as the micromodel network.

The advantage of holographic over other interferometric techniques, such as Rayleigh or


Mach-Zehnder, is that expensive optical components are not required, and the reference system
is the state of the object at the time the hologram was exposed. This technique maps lines of
constant pathlength change, which for uniform depth corresponds to lines of constant refractive
index. However, the composition must be uniquely defined by the refractive index in order to
convert this information into a composition map. Therefore in general only binary systems can
be studied, although even some binary systems, for instance water-methanol, are not uniquely
defined by the refractive index.

The real time holographic interferometer is shown in figure 6.1. The elements were
clamped magnetically to an optical table. Vibrations were damped by floating the table on six
car inner tubes. The hologram plate was not moved after exposure, but was processed in situ
in a specially designed tank/plate holder. This ensured that the hologram image was exactly
superimposed on the object. The resultant interference pattern was projected onto a 120 film
camera, and observed in detail using a television camera.

6.1 Theoretical Basis of Holographic Interferometry

6.1.1 The Hologram

The name "hologram" was coined by Professor Dennis Gabor and derived from the Greek
word holos which means "the whole", because a hologram records the whole information of an
object wave — both phase and amplitude. Gabor developed the idea as a method for
increasing the resolution of the electron microscope. He produced the first hologram in 1948
using a point source of a mercury spectral line. 94 " 96 However the real breakthrough in
holography arrived with the invention of the laser, an abundant source of coherent light, and
which was first used for holography by Leith and Upatnieks in 1963.97

116
Figure 6.1 Holographic interferometry apparatus.
117
The hologram is a record of the light distribution scattered from or transmitted by an
object, in which the depth, and therefore three-dimensional, information has been encoded such
that it can be retrieved. All the information, both amplitude and phase, about the wave at the
hologram recording material is encoded into the hologram. After the hologram material has
been processed the recorded wave may then be reconstructed to give a perfect replica of the
original wave at the hologram. An image of the object is seen in the same position as the
object by looking through the hologram. The image is truly three-dimensional (displays
parallax), and is completely indistinguishable from the original object; it is as though one were
looking through a window at the object itself.

In conventional photography, the lens system transfers the light distribution at a plane
within the object onto the plane of the film, which records the light intensity distribution as a
function of "opacity"; the photographic image is therefore two-dimensional. The phase
information of the light is not recorded, because the intensity is the square of the complex
amplitude. If the amplitude is represented by A0then the intensity is A0e"i'XA0e~"t'=
A02, and the phase information represented by el<t> is lost. Light from other planes in the
object is lost by defocus, and no depth or parallax information is available when viewing a
photograph.

All the information in the object wave is recorded in holography by superimposing a


coherent reference wave and the object wave on the holographic material (usually a fine grain
photographic emulsion). Since the two waves are coherent the two waves interfere, producing a
constant interference pattern (provided that the waves are not altered by, for example, moving
the object). The amplitude at exposure at a point of the photographic plate is the sum of two
simple harmonic terms, one representing the amplitude and phase of the coherent wave issuing
from the object, A0 exp(z'0), the other the amplitude and phase of the coherent background,
Ar exp(z'0). The photographic plate records the intensity of the light and can be processed as
a transparency such that the intensity of transmission at any point is proportional to the
square of the intensity at exposure (i.e. with a "gamma" of two). The amplitude transmittance,
Ta, of the plate is given by

(6.1)

where K is a constant. When the processed plate is illuminated by the reference beam only then, from
equation 6.1, the transmitted amplitude is

U = TaAre

118
therefore,
22 -/ (0-2^)
U = K{A I + A L)Areiyp + K A 0 A + KA 0 A re (6.2)

The first term in equation 6.2 is similar to the incident reference wave and represents

light that passes straight through the plate. The other two terms are diffracted waves; the

second term is the reconstructed object wave, identical to the original except in relative

intensity, and the third term is a ghost on conjugate image. Since the hologram encodes the

amplitude and phase of the wave, no lens system need be used to transfer object distributions

to the plane of the photographic plate.

The source of the coherent light is a laser which is split into two beams: object and

reference. A simple optical arrangement for a transparent or translucent object is shown in

figure 6.2. Reconstruction of the object wave is illustrated in figure 6.3. When the object and

reference beams are both planar and a is the angle between them, the interference pattern

recorded on the photographic plate consists of a series of dark parallel lines. The line spacing

is A/sina (see figure 6.4.) where A is the wavelength of the monochromatic light in vacuo,

since the optical pathlength difference between lines is one wavelength. The photographic

emulsion must be able to resolve lines of at least A/sina spacing in order to record the

hologram. The resolution of the hologram image is equal to the recorded line spacing, and

therefore the resolving power is increased by increasing the angle between the object and

reference beams, but is limited by the resolution of the photographic material.

beam
splitter
I laser [

mirr
hologram
plate
object

Figure 6.2 Optical arrangement for producing a hologram.

ref e r e n c undiffracted light


beam (zero order)

conjugate
image
\ V virtual
image

Figure 6.3 Reconstruction of the hologram image.

119
d sina
object
bearff

reference
beam
Figure 6.4

6.1.2 Linear Response and Defraction Efficiency of Photographic Materials

So far we have considered so called "amplitude" holograms where the incident light is

diffracted by a pattern of dark lines to form the image wave. A much brighter hologram

image is produced by bleaching the plate during processing. The spatial variation of absorption

is converted to a corresponding variation in either emulsion thickness or refractive index (RI),

or more usually a bit of both, and therefore this is termed a "phase" hologram. The silver in

the emulsion is oxidised, forming a transparent salt which has a different refractive index from

the gelatin. The mathematics for phase holograms is slightly different from that for amplitude

holograms, but the principle is similar; the thickness or RI variation diffracts the

reconstruction beam to produce the image waves. In principle the diffraction efficiency of a

phase hologram is 100 per cent., although in practice phase holograms are not totally

absorptionless and the maximum efficiency is usually about 70 per cent., 98 which is very

much greater than that for amplitude holograms.

The amplitude transmittance of unbleached photographic material is linear with exposure

for only a small exposure range. The beam ratio (relative intensity of the reference and object

beams) and the exposure time need to be adjusted for a linear response. This requires a high

beam ratio (for example 10:1), but this has the penalty of low diffraction efficiency, so

usually a compromise of 3:1 is used. The refractive index of bleached emulsion is also not

linear with exposure, and for phase holograms a beam ratio of 7:1 is recommended. The

exposure time is dependent on the incident light intensity.

6.1.3 Holographic Interferometry

Equation 6.2 gives the transmitted amplitude when the hologram plate is illuminated by

the reference beam only. If the processed plate is replaced exactly in its position during

exposure, the virtual hologram image is superimposed in space on the object. When the object

beam is also incident on the plate, the object and image beams interfere since they are

120
coherent, and the transmitted amplitude is now given by

U = Ta(A0e>* + AreP).

If the complex amplitude of the object beam is changed from the value when the hologram was exposed,
A0el<t>, to Ax e' e , where e' e represents the new phase distribution, then the transmitted amplitude is

u = r 0 ( V 6 + • + * ) •

Substitution for Ta from equation 6.1 gives

2 1 1 l 9
U = K ( a q + A ) + A ^ e ^ + A ^ e - ^ ) [Ax e + A r e ,

Therefore,

C/ =K((A 20 + A?)ALE I9 + A0A?E 1^

+ K(A I + A^ARE^ + V ( * + < H W

I(<FI D
+ KAQAx ARE~ - -^ + KAQA rV'<*-*>.

The first term in equation 6.3 represents the interference between the object and image
waves, the second term is the undiffracted portion of the reference wave and the other terms
are conjugate waves. The intensity, /, of the first term in equation 6.3 is

/ - S ^ L + A ^ A ^ + AEALE'^AL + A ^ A ^ + ACALE'*) ,

i.e.
/ = K2 (^AI + A I) 2 A \ + A \A 4 + I{A I + A 1)a0AX A 2 cos(0 - 0) j . (6.4)

Equation 6.4 shows that the resultant image is modulated in intensity by the cosine term. A dark fringe
is seen when

0-</> = 2 7 r ( m + i ) , (6.5)

where m = 0, + 1 , + 2 . . . . When cos(0 — 0) = —1, equation 6.4 becomes

/ =K 2(^AL + A 2^AX-AQA 2)J .

I = 0 if Ar » A0 and therefore the dark fringe is black.

The fringe pattern can thus be interpreted in terms of a change in the object wave, since the
reference image wave is constant. For a transparent, stationary object such as a micromodel,
the object wave is altered by changing the refractive index of the fluid within the pores; this
changes the optical pathlength through the pores.

121
If the pore depth is d and the fluid refractive index (RI) is /*, the optical pathlength
through the fluid is n-d. An RI change of An results in a pathlength change of An-d and

(6.6)
A

Equations 6.5 and 6.6 give


^ 2kAnd . / . lX
0-0 = — = 2K (M
A

and
A«J=(w+i)A, (6.7)

for a dark fringe.

The refractive index, n, is the average value over the pore depth and therefore refractive
index gradients within the depth direction are not resolved by this technique. To calculate the
value of the refractive index, the pore depth must be a known function of position; the *
simplest case being uniform depth. The fringe pattern corresponds directly in time to the state
of the object, hence the term "real time" holographic interferometry. If the hologram is
exposed twice (not developed in between), once before and once after changing the object, two
images are stored in the hologram. When the hologram images are reconstructed using the
reference beam only, a fringe pattern is seen superimposed on ihe image corresponding to the
interference pattern between the two coherent image waves. This method is termed "frozen
fringe" interferometry because the interference pattern is frozen into the hologram image.
Frozen fringe holography is easier to conduct than real time interferometry, but it is less
versatile. 4

6.1.4 Reference Fringes

The object wave can be changed by slightly altering the angle of the light incident on
the transparent object. This does not change the position of the object, only the phase, and
produces a pattern of parallel fringes apparently superimposed on the object during real time
interferometry. Thus

where e is the angle of rotation and the x-axis is perpendicular to both the axis of rotation and the
original beam direction, as shown in figure 6.5. Substitution for 0 in equation 6.4 from equation 6.8
gives

l = K 2 ( ( A l + A y A 2 l + A 2 0 A t + 2(A 2 0 + A l ) A 0 A l A j c o s ^ f ^ - ) . (6.9)

122
For small e, sine e, and the image intensity is modulated by the cosine term. If Ar » A0 and Ax «

A0, equation 6.9 gives

/ = 2K2A4rA + cos •

The fringe spacing, Ax, is given by equations 6.5 and 6.8, since

0 — 0 = = 2x(m + 0 »
a

xe = (M +i)\ ,

and

A x = * . (6.10)

This set of parallel fringes is termed the reference, or auxiliary, fringe set. Subsequent

changes in the object by variation of RI, will perturb this fringe pattern. At a point on a

dark fringe the phase change d—<j> is given by equations 6.5, 6.6 and 6.8

^ , 2irAnd . lirxe „ , . >


0 - 0 = — + — — = 27r(m+i),
A A

therefore

And + x e = (m + ±)X.

If the mth fringe is initially at x 0 , a change in refractive index of An will cause the fringe to shift to x

and

And + xe = *0e ,

i.e.

An = (*0 — x)e ,

:
0 ~'
And = \ - ^ X = jX (6.11)
A*

from equation 6.10. Thus An at point x can be calculated from the fringe shift, provided that d is a

known function. This is illustrated in figure 6.6.

\
object

Figure 6.5 Rotation of the object beam direction.

123
— x0-X

Ax —

m-2 m-1 m m*1 m*2

(a) (b)

Figure 6.6 Fringe shift.

6.1.5 Rotatable Plate Method for Producing Reference Fringes

In this work, a rotatable optical flat method, similar to Durou's,93 was used to change

the angle of the object beam and produce the reference fringe set. The principle is illustrated

in figure 6.7. From the geometry shown in figures 6.7a and 6.7b

1
tan(90 ° - /) =
e (tan i — tan r) tan /

tanr
7 = ^ 1 " tan i

sinr
sin/

(6.12)
V '

from Snell's law, where ng is the RI of the transparent plate. Figure 6.7c shows that

b\ = cb?7[ s i n ( / - 0 - r J ,

and on expanding the sine term,

bx = e(sin(i -6) — cos(/' — 6) tan r^J .

124
Since,
sin(/ - 0) tan(i - 0)
tan/-! « sin^ = — H » w
g g

Substitution from equation 6.12 gives


bx =szsin(i -d). (6.13)

Figure 6.7d shows that

bx = (s z tan/ —sx) cos i. (6.14)

Equations 6.13 and 6.14 give

sin(/—0)
= s.\ tanz —
cosz

= s z ( t a n / ( l - cos 0) + sin0^ . (6.15)

The angle of rotation of the collimated rays is

6= j, (6.16)

where / is the focal length of the lens. Equations 6.12, 6.15 and 6.16 give

e = j (1 - ^ ) (tan i (1 - cos 0) + sin 0) . (6.17)

For small 0 equation 6.17 becomes

Letting
/ = 200mm
e = 20mm ,

0 = 30°,

equation 6.17 gives e = 0 • 017 radians for both / = 0 and i = 3°, and substitution for e in equation
6.10 gives Ax = 31pm for A = 632 • 8nm. Letting 0 = 45 gives Ax = 26pm.

ft

125
lens
(a) (b)

(c) (d)

Figure 6.7 Rotatable plate method for producing reference fringes.

6.1.6 Laser Speckle

Objects illuminated by coherent light, for instance from a laser, appear to have a
granular texture, which is termed speckle. Speckle can be explained by considering the object
as an array of point scatterers. The scattered coherent rays interfere randomly, producing the
speckle pattern. The size of the speckle grains depends on the resolution of the observer's
eye or optical system. This is easily demonstrated by placing an iris in the pathway. Closing
down the iris decreases the resolution and increases the speckle size. The diffraction limited
resolving power is given by
A
D
O = (6.18)
sin*/

where 2u is the maximum angle subtended at the object by the observer's eye, and d0 is the
size of the smallest element that will be resolved. For a simple lens, sinu=.2)/2z where 2> is
the lens diameter and z is the object distance. Since the speckle size is equal to d0, the
speckle does not reduce the image quality, but is in fact a useful measure of the resolution of
the lens system. Interference fringes will therefore only be resolved when the fringe spacing is
greater than twice the speckle size.

126
6.2 Method of Holographic Interferometry

Interferometric techniques are very sensitive to vibration, indeed they can be useful tools
for detecting vibration, and therefore it was essential to damp vibrations from the optical
table. An in situ hologram plate processing method was used to ensure that the object and
hologram images were exactly superimposed; a change of 0-2pm can be detected. The
apparatus was housed in a constant temperature environment in order to reduce any systematic
changes in optical pathlengths due to thermal expansion.

6.2.1 Apparatus

The optical table consisted of a 1-535 X 0-93m, 0-5 inch thick steel plate mounted on
k a steel cabinet. The plate was floated on six inflated car inner tubes to damp out any
vibrations transmitted to the cabinet from the floor. Initially a box, constructed from
aluminium sheet, covered the apparatus to eliminate any stray laser light. However this was
found to set up vibrations in the apparatus, because it behaved like a eardrum, transmitting
^ airborne sound waves to the base plate. Consequently, a cover made from corrugated

cardboard was used.

A Spectra-Physics 120s 6mW He/Ne laser was used together with Agfa-Gevaert 10E75
Holotest plates, which have a resolving power of 3000 lines/mm. The mirrors were front
silvered microscope slides,and the laser beam was split by a dialectric coated glass plate beam
splitter. The two beams were expanded by spatial filters, each consisting of a X 2 0 microscope
objective with a 25pm diameter pinhole at the focus. The object beam was collimated using a
200mm focal length plano-convex lens. The mirrors and lenses were mounted on standard
* optical equipment, but the laser, phase plate, object and hologram plate mountings were

custom built. The elements were secured magnetically to the base plate, as shown in figure
6.8. The hologram plate holder/in situ processing tank is shown in figure 6.9.

The reference fringes were produced by the rotatable disc method of section 6.1.5, and
the apparatus is shown in figure 6.10. The disc was optically flat and parallel and made from
borosilicate glass ("Pyrex") which has a refractive index, nD20= 1-475. The disc was 0-75
inches thick and 1-5 inches in diameter. Angles of rotation up to 30° (corresponding to a
fringe spacing of 42pm) were possible by this arrangement.

127
The interfering object and hologram images were projected onto a Bronica 120 film

camera (lens removed from camera), such that the whole network region of interest was

recorded at each exposure. The arrangement used is shown in figure 6.11. Here a television

camera was mounted on a travelling stage (movement in two orthogonal directions) above the

Bronica camera. The TV camera monitored the fringe pattern in detail.

Figure 6.8 Micromodel holder and spatial filter.

128
Figure 6.9 Hologram plate holder/z>z situ processing tank.

129
Figure 6.10 Reference fringe apparatus.

130
\

Figure 6.11 Bronica and TV cameras

6.2.2 Experimental Procedure

The optical arrangement is shown in figure 6.12. The spacial filters, lenses and

micromodel were aligned normal to the beam by ensuring that the light reflected from glass

surfaces was co-linear with the incident light. The optical pathlengths of the reference and

object beams were approximately equal to ensure coherence. The angle between the beams was

nominally 30°, which corresponds to a fringe spacing at the hologram plate of around 1-2/um

(830 lines/mm). The beam splitter was selected to give a beam ratio of about 7:1.

The hologram plate was pre-soaked in distilled water for one hour before being placed in
the plate holder/tank. The hologram was exposed for two seconds with the tank full of water,
termed wet-gate, and then processed in situ by flowing the various chemicals into and out of
the tank, as shown in figure 6.13. A safe light was not used during exposure and processing
because of the broad wavelength sensitivity of the hologram material.

131
Figure 6.12 Optical arrangement for holographic interferometry.

Ccf 0 tf tf

Waste Waste
HOLOGRAM
TANK

Waste

Figure 6.13 In situ processing arrangement.

132
The processing procedure was:

1. Develop in Agfa-Gevaert G3P (now obsolete) for four minutes;

2. Fix in Ilford Hypam (one part plus four parts water) for two minutes;

3. Wash in running water for five minutes;

4. Bleach, continue for one minute after apparent clarity is achieved, using bleach formula

of Phillips and Porter;98

5. Wash in running water for fifteen minutes.

All fluids were maintained at 20°C. The bleach constituents (stock solution) were:

20g glycerol

500ml propan-2-ol

500ml distilled water

300mg phenosafranine

150g ferric nitrate

33g potassium bromide.

The stock solution was diluted one part plus four parts water for use. This bleach oxidises the

silver, converting the absorption distribution to a corresponding refractive index distribution,

producing a phase hologram. This process is capable of producing holograms with a diffraction

efficiency of 70 per cent.98

The hologram image was reconstructed under wet-gate conditions. The wet-gate method

reduces the effect of emulsion shrinkage, which would change the image position, reduces

image noise, which results from the non-uniform emulsion, because the RI of water is closer

to that of gelatin (1-53) than air is, and eliminates the need to dry the plate and tank.

The interference fringe patterns were recorded on Ilford Pan-F film, exposure times

varying between l/500s and 1 /60s depending on the image magnification. The film was

developed in Ilford PQ Universal, diluted one part with nine parts water, for five minutes, to

give high image contrast.

The fringe pattern images were printed onto grade four or five photographic paper, at

identical magnification for each pore. The reference fringes for the first exposure, t = 0 , were

traced onto clear acetate sheet, which was used as an overlay on the subsequent prints, as

shown in figure 6.14. The points where two fringes crossed were marked onto a second

overlay. The isocomposition profiles were produced by joining the points corresponding to

identical fringe shifts, as shown in figures 6.14c and 6.14d.

133
1= 0 t>0

(a)

7/ L
/ u/u
hI (c)

t >0
with overlay from t = 0 L i n e s of c o n s t a n t R I

Figure 6.14 Method for determining lines of constant refractive index.

6.3 Preliminary Experiments

The experimental method was tested using

• wedge model to check hologram exposure time, beam ratio, speckle size, and frozen

fringe and real time interferometry;

• cover slip model to check the fringe analysis for a uniform depth model.

6.3.1 Wedge Model

The wedge models were constructed from glass microscope slides and sealed with

'Araldite" adhesive and "Loctite" glass glue. The design is shown in figure 6.15.

134
(a) (b)

(c)

Figure 6.15 Wedge model.

Frozen fringe technique

The model was filled with fluid A of refractive index, nA, and the hologram exposed for
half the normal time with a ground glass screen behind the wedge. Without processing or
moving the hologram plate, the fluid in the wedge was replaced by fluid B, RI=n B . The
hologram was re-exposed and then processed. The hologram image was reconstructed using the
reference beam only, and examined. A series of parallel fringes (parallel to the x-axis in
figure 6.15) were observed, frozen into the hologram image. The speckle size and fringe
spacing were measured and compared with the calculated values. Refering to figure 6.15c, the
optical pathlength through the fluid wedge is
nFd = nFz tanw, (6.19)

where nF is nA or nB. The change in pathlength is therefore,

And = (nB — nA)z tan< (6.20)

Equation 6.7 states that a dark fringe will occur when

And = [m +i)\

Substitution for An d gives

(inB -fl 4 ) z t a n a , = ( m " K M >

135
and the fringe spacing is
Az = . (6.21)
[ n B - n A jtano;

Several fluid pairs and two values of w, 0-5° and 3°, were tested. The measured fringe J
spacing correlated with equation 6.21. The measured speckle size was comparable to the
estimated resolving power of the viewing system.

The refractive indices of the fluids were measured at the sodium doublet wavelength,
\

589 -3nm, but the laser wavelength was 632-8nm. Most fluids exhibit spectral dispersion, and
therefore the fluid RI would be different for each wavelength. However, these experiments
showed that although the absolute values may be different, the difference in refractive index
between two fluids was identical at both wavelengths, at least to within experimental error.
This meant that the refractive index properties of fluid systems could be measured at the %
sodium doublet wavelength and used to characterise the behaviour at the laser wavelength.

The wedge models were then used to test the in situ processing tank method for real
time interferometry.

6.3.2 Cover Slip Model

to
A microscope cover slip model was constructed to test the experimental procedure for
measuring mass transfer in dead-end pores, especially the interference fringe analysis. The
model was made by sticking circular glass cover slips, 13mm in diameter and 150Mm thick,
onto a glass microscope slide with "Loctite" glass glue (set by exposure to ultra violet light).
Inlet and outlet holes were drilled into a second glass slide with an ultrasonic drill, and then ^
this was glued onto the tops of the cover slips. The glue was allowed to imbibe into the gaps
between the discs and the slides from the slide edges, via the contact points of the glass
discs. This model had uniform depth, which simplifies the fringe analysis and also helps to
ensure that the composition field is truly two-dimensional, but the pores were relatively large
and their shape was not easily controlled. The interferometry showed that the glass glue was
also affected by organic solvents, for instance decane and propanol.

The test models showed that holographic interferometry can be applied to micromodel
experiments. An etched cover slip model was constructed, which had more realistic pore sizes m
and uniform depth. The mass transfer results from this model will be described in chapter
seven.

136
CHAPTER S E V E N

STUDY OF M A S S TRANSFER INTO DEAD-END PORES

Micromodels were used to study mass transfer into dead-end pores. Both a totally miscible

and a partially miscible system were investigated:

• Totally miscible case — decane and propan-l-ol. Holographic interferometry was used to

determine the composition field, and mass transfer was modelled by a mass transfer

coefficient equation. The fluid density of the decane-propan-l-ol system is not constant

and buoyancy forces were found to affect both the isocomposition profiles and the rate of

mass transfer at different pore orientations. However in both the experiments, mass

transfer through the pore throats was diffusional.

• Partially miscible case — water and butan-l-ol. Holographic interferometry was not used

quantitatively in this study, because there was insufficient range in refractive index within

the two phases. The rate of mass transfer was determined by measurements of the

relative phase volumes within the dead-end pores, and again modelled using a mass

transfer coefficient equation. The density of this system is also dependent on composition

and gravity forces were found to affect the rate of mass transfer. However, the systems

did not become hydrodynamically unstable under adverse gravity conditions, because of

the stabilising influence of the fluid-fluid interface due to capillary pressure.

A uniform depth micromodel was used in all the experiments in order to ensure that the

composition field was truly two-dimensional, which was assumed in the measurements and

calculations, and also to simplify the analysis.

7.1 Experimental Method

7.1.1 The Micromodel

The micromodel network shown in figure 7.1 was etched into a glass cover slip (115mn

thick) through to an Araldite backing, as described in section 3.2.2. The completed

micromodel, figure 7.2, had a uniform pore depth, although Araldite encroached into the pores

and channels during sealing. The pore throats were designed to be 500nm long and 250/um

wide, because 200Aim was found to be the maximum pore width that could be etched to a

depth of lOO^m. The pore bodies were nominally 1000 and 2000/xm wide in order to

incorporate a high pore aspect ratio.

137
Figure 7.1 Dead-end pore micromodel design.

The complete pore dimensions of the three pores used, are given in table 7.1. The cross-sectional
areas, AT and AB are average values given by

where VT and VB are the throat and body volumes, and £ r and £6 are the throat and body lengths

respectively.
ft

Table 7.1

VT/d AT/d kr vB/d AB/d kB vp/d


Pore
w (nm) (nm) (nm2) Gm) (m) (urn)
Small 105 750 235 450 157 500 445 355 263 250

Medium 138 440 265 525 740250 990 750 878690

Large 142 800 255 560 3630000 2000 1815 3 772 800

7.1.2 Fluid Properties

The refractive index and density properties of the miscible, decane-propan-l-ol, and
partially miscible, water-butan-l-ol, systems are given in figures 7.3 and 7.4. The difference ^

between the refractive indices of decane and propan-l-ol is 0-0268, and from equation 6.22 a
depth of 115jttm corresponds to a shift of about five fringes. Unfortunately there is insufficient
range in refractive index within the two phases of the water-butanol system for quantitative
evaluation of the concentration field.

138
J 5 0 0 (im

Figure 7.2 Dead-end pore micromodel, showing reference fringes.

139
Mass f r a c t i o n of propan-1-ol

Figure 7.3 RI and density for the decane-propan-l-ol system.

Mass f r a c t i o n of water

Figure 7.4 RI and density for the water-butan-l-ol system.

140
7.1.3 Experimental Procedure

The fluids were passed into the micromodel as shown in figure 7.5. The model was
evacuated and saturated with the first fluid through the two inlet tubes, with the outlet valve
closed. The first fluid was decane for the miscible case, and butanol for the partially miscible
system. The saturated micromodel was secured in place on the optical table (chapter six) such
that the network was in a vertical plane, i.e. the y-direction in figure 7.1 was horizontal. The
orientation of the x- and z-axes was adjusted before each experiment.

• O EL

waste" ^ ' f f ^ j T w Oa s t e
W
13 c 3c

teflon tubing

fluid 1 fluid 2

waste

Figure 7.5 Valve arrangement.

At this stage the hologram was exposed and processed, as described in chapter six, and
the phase plate rotated to produce the set of parallel reference fringes. A photograph was
taken to record the state of the model at time, / = 0 . With the outlet valve closed, the second
fluid was injected through the inlet tubing and out through the bypass valve. After sufficient
fluid had been injected to displace all the initial fluid in the tubing, the flow rate was set to
0-8cm 3 /hr (10_3ms_1 in each channel), the outlet valve opened and the bypass valve
closed; time t=t 0 . The fluid flow maintained the boundary condition of W a 0 = l , t>0.

7.2 The Mass Transfer Equations

7.2.1 Miscible Fluids

Fick's first law of binary diffusion, relative to stationary coordinates, for components a and b, is

(7.1)

141
where j is the mass flux ( M L - 2 T - 1 ) , w is mass fraction, /3 is the total fluid density (ML - 3 )
and d is the mutual diffusivity ( L - 2 T - 1 ) . The second term in equation 7.1, wa ( j a + j j ) ,
represents the mass flow of species a resulting from bulk motion (convection) of the fluid.
The mass concentrations of components a and b are given by

w
Pa = aP >
(7.2)
= w
Pb bP'

noting that
+ w 1
b = >
=
Pa + Pb P;

dw„ dw
-i- i-a = n (7-3)
dt ^ dt u'
d d
Pg • Pb _ dp
dt ^ dt dt'

For mass transport into the dead-end pore space, conservation of mass gives

A J
A T = v
V
= v V
a'r>
T aP p d t p d t
(7.4)
a j -vd<p
V
b> V
_vd«wb>p)
T" bO p d t p d t

where AT is the throat cross-sectional area, Vp is the volume of the pore, Ja0 and J60 are the mass fluxes
at the pore throat entrance, <p> and < w ) are volume averages given by

J
<P> = ^
pdV

• (7.5)
1 '
dV

Letting Vwa = ^f and substituting equations 7.3 and 7.4, equation 7.1 becomes

v
_ v P d ( p a > _ / f a p d ( P y
(7.6)

where wa0 is the mass fraction at the pore entrance.

If the density remains constant, the last term in equation 7.5 becomes zero. We can define a mass
transfer coefficient in a manner similar to equation 5.20, by

^==^jpao~<pa»- (7-7)

Equations 7.3, 7.4 and 7.7 yield

142
where p0 is the density at the pore entrance. If there is no volume of mixing effect,

P= WaPa\ + WbPbl ' (7-9)

where pal and pbl are the densities of pure fluid a and pure fluid b respectively. Substitution from equa-

tion 7.9 for p and wb from equation 7.3 into equation 7.8 gives

VP
=
K
m(WaO-<Wa>) ^ ' c7'10)

Rearrangement of equation 7.10 and integration gives

In ^—r = — ^ t, (7.11)
vv
oo - < m , a > v
p

when <w f l ) = 0 at t = 0 and waQ is constant for t > 0. For the boundary condition w ^ = 1 for t > 0,

equation 7.11 becomes


1 t
, n T ( 7 ' 1 2 )

7.2.2 Partially Miscible Fluids

Let the average mass fraction of component a in phase 1 equal the value at the fluid-

fluid interface, wx (dropping the subscript a for convenience) and similarly for phase 2,

<w a >=w 2 . Equation 7.5 becomes

<P> = y ( ^ p d V + (7.13)

where Vl and V2 are the volumes of the two phases in the pore. Equations 7.2, 7.4 and 7.7 are still ap-

propriate and with equation 7.13 give

ATJaQ = ATKjpM-WlPi) = V ff + 1 2 /, 2 2; • (7.14)

But
dw j _ dw2
~~dT = ~dT = 0'

d d
P± = P l = r ,
dt dt

and
dV_i _ _dV2
~dt "5T'

therefore equation 7.14 becomes

At Ja0 = AT Km{pa0 - w l P l ) = - ( w , P l - W2p2) . (7.15)

143
Rearrangement of equation 7.15 and integration gives

V V
2 ~ ti _ KmAT PaO-WjPi
Vp Vp VV1P1-W2P2 ( ' - ' / ) . (7-16)

where Vt is the volume of phase 2 in the pore at time t = t{.

7.3 Mass Transfer Results — Miscible Fluids

The decane-propan-l-ol experiments were conducted at two channel orientations:

• horizontal channel, z-direction vertically upwards, x- and y-axes horizontal (figure 7.1);

• 45° inclined channel, y-axis horizontal, angles subtended between the z-axis and the
upward vertical, and between the x-axis and the upward vertical were both 45° (i.e.
outlet at a higher elevation than the inlet).

In both of these experiments the mass transport through the pore throats was diffusional; i.e.
gravity stable, and therefore no convective transport through the throats.

In the first experiment — the horizontal channel — the fringe shift between pure decane
and pure propan-l-ol was 5*3. Substituting j=5-3, A«=0-02683 and X=0-6328 into equation
6.11, the pore depth is

= 5 - 3 X 0 - 6328
0-02683 '
d = 125pm .

The refractive index of each isocomposition profile is given by substituting for d in equation 6.11, and
the initial fluid refractive index:
, 0-02683 .
rij = l • 41215 J-j—J • (7-17)

The fluid composition corresponding to each tij (J = 0 to 5 and 5 • 3) are given in table 7.2 from the
data shown in figure 7.3.

The fringe patterns at long times showed that the propanol was affecting the Araldite
slightly; the alcohol was probably being absorbed, or was dissolving components from the
Araidite. The micromodel was therefore heat treated at 50°C for one hour to make the resin
more inert. This treatment reduced the pore depth, and in the second experiment the
maximum shift was five fringes, corresponding to a depth of 118Pm. The composition values
for each rij were therefore different for the experiment where the channels were inclined at
45°, and are given in table 7.3.

144
Table 7.2 0°. Table 7.3 45°.

j RI. w j RI, w

1 1•40709 0 168 1 1•40678 0 •180

2 1 •40203 0 348 2 1-40142 0 •373

3 1 •39696 0 545 3 1 • 39605 0 •580

4 1-39190 0 •733 4 1•39069 0 •783

5 1 •38684 0 •938 5 1-38532 1 •000

Example fringe patterns are shown in figure 7.6. The isocomposition curves, Cjt in the

three pores at successive times, are shown in figures 7.7 and 7.8 for the horizontal and 45°

inclination experiments respectively. The density of the fluid within the pore necks was greater

than that within the bodies, and therefore the fluid just inside the body was more dense than

the surrounding fluid. The resultant buoyancy forces caused the fluid to flow sideways and

produced the characteristic straight, horizontal profiles shown in figures 7.7 and 7.8.

j 250 [Im

Figure 7.6a Time, t=0.

145
Figure 7.6b Time, t>0.

146
Figure 7.6c Fringe pattern from figure 7.6a superimposed on figure 7.6b.

Figure 7.6 Fringe patterns.

147
M

Figure 7.7a Small pore, 0°, isocomposition lines.

148
Figure 7.8b Medium pore, 45°, isocomposition lines.

149
fe • 3 mins
tQ • 8 mins

fe + 14 mins
fe • 25mins 30s

Figure 7.7c Large pore, 0°, isocomposition lines.

150
t0 + 1 hr 20 mins
t0 + 2hrs 19 mins 30s

t"0 + 3hrs 10 mins

Figure 7.8b Medium pore, 45°, isocomposition lines.

151
15s

Figure 7.8a Small pore, 45°, isocomposition lines.

152
t0 + 2 mins 15s
to + 5 mins 40 s
C2.s Qs

to • 15s

t 0 + 10mins 3 0 s t0+28mins
t0 + 16 mins 45 s

Figure 7.8b Medium pore, 45°, isocomposition lines.

153
Co

t 0 • 15 s

t0 + 5 mins 40s ^

Figure 7.8b Medium pore, 45°, isocomposition lines.

154
to* 10 mins 30 s
f,, + 16 mins 45 s

f 0 + 28 mins
t0* 1 hr 20 mins

Figure 7.8b Medium pore, 45°, isocomposition lines.

155
\
\
r
J K
to + 2hrs 17 mins 30 s
t0 + 3hrs 24 mins

Figure 7.8c Large pore, 45°, isocomposition lines.

7.3.1 Mass Transfer Time Constant

Equation 7.2 gives a simple relationship between the volume average mass fraction of

alcohol in the pore and time:

where the time constant, r, is a constant provided the mass transfer coefficient is a constant

for the particular pore and fluid system. <w> was calculated by a numerical integration

technique. The pore volume was divided into elements of volume Vp each associated with a

mass fraction, w , as shown in figure 7.9. The average composition is

(7.18)
p

where Vp is the volume of the pore. Since the depth was uniform, the elemental volumes

were calculated by measuring the area of each element on the diagram using a Ushikata Digi-

Plan 220L electronic digital planimeter. <>v> is given as a function of time in tables 7.4 and

7.5 for the horizontal and 45° inclination experiments.

156
Figure 7.9 Integration method.

Table 7.4 0°.

Time <w> <w> o >

(s) Small Medium Large

180 0-494 0-156 0-035

480 0-751 0-270 0-048

840 0-921 0-425 0-067

1530 0-959 0-593 0-137

2400 - 0-760 -

4 800 - - 0-391

8 370 - - 0-587

11400 - 0-947 0-696

Table 7.5 45°.

Time <w> <w> <w>

(s) Small Medium Large

15 0-356 0-014 0-022

135 0-706 0-287 0-076

340 0-900 0-450 0-166

630 - 0-584 0-234

1005 - 0-715 0-282

1680 . - 0-856 0-365

4 800 - - 0-656

8 250 - - 0-836
12 240 - - 0-856

157
l n ( l / l — <w>) is plotted against time in figures 7.10, 7.11 and 7.12. For figure 7.12 the
composition was averaged over the pore body volume, VB. The time constant r in equation
7.12 is given by the gradient of each straight line in these figures, and is given in tables 7.6
and 7.7, where r 0 , r 45 and r 0 * correspond to figures 7.10, 7.11 and 7.12 respectively. These
data show that mass transfer between decane and propan-l-ol can be characterised by a mass
transfer coefficient, which is however, dependent on both pore geometry and pore orientation.
Inclining the pore channels markedly increased the rate of mass transfer.

Table 7.6 Table 7.7

Ui
Pore To r
45 T
o
Pore v t
o
Small 336 111 0-51 Small 339 0 51

Medium 1638 997 0-61 Medium 1754 0 57

Large 9515 5022 0-53 Large 9 575 0 52

I n ( r ^ )

0 2500 5000 7500 10000 f (s)

Figure 7.10 0°, w averaged over V .


Figure 7.11 45°, w averaged over Vp.

Figure 7.12 0°, w averaged over Vt


7.3.2 Mass Transfer Coefficient Correlation

Equation 7.12 gives


Vp

Table 7.8 shows that Km is not a constant for the three pores. The solution for diffusion into a plane
sheet, equation 5.17, gave equation 5.22:

m
~ 71"'

where 2£ is the thickness of the sheet. Table 7.8 shows that Km does not correlate with the total pore
length, £ r + If steady state diffusion through the pore neck is assumed" then the mass flux into
the pore is
j _ D <P.0- < P „ » (? 19)

This
is equivalent to setting Km — j- in equations 7.7 and 7.12, however <(pfl)> and (wa)> are averaged
over the pore body volume, VB, hence figure 7.12. Equation 7.12 becomes
1 D At t
ln
i / \ g v~rt =
- * -
1 - <w> v
B r*

Table 7.8 shows that equation 7.20 gives the best correlation for the diffusivity, D, with an
average value of 0 « 8 5 X 1 0 - 9 m 2 s - 1 . This is a composition average value since the diffusivity
is probably composition dependent. Independent data have not been found for the diffusivity
of the decane-propan-l-ol system, however there are several equations in the literature to
predict liquid diffusivity at low concentration, i.e. wfl=>0 and vva=>l. Wilke and Chang100

give
15 v
d =7-4x10 l b v0 b 6 , (7.21)
_ m * a

where \f/b is an association parameter (ypb= 1 for b=decane, and ^ = 1 - 1 for b=propan-l-ol),
Va is the molar volume of a at the normal boiling point (cm 3 g-mole - 1 ), p is the solution
viscosity (Pa.s), Mb is the molecular weight of b (g), T is absolute temperature (°K), a is the
solute and b is the solvent. This equation is reputed to predict the diffusivity to within ±10
per cent. Equation 7.21 predicts the diffusivity at low alcohol concentration to be 2-1 X
10~ 9 ms _ 1 , and 0 - 3 3 X 10~ 9 ms _ 1 for high alcohol concentration, both for 20°C.

Table 7.8

VB ^ ( I r + fc) V
PTR Vb(It+SB) VB VB
Pore
Ap TQ Aj TQ Aj TQ Aj TQ AJ TQ* AJ TQ* AJTQ
(X10"6MJ-') (XI0~4/MJ-1) (XKR'MV) (XLO-'M2*"1) (X10-9 MJI_L) (XLO-'M2!-1) (X10 -9M2J-')

Small 3-329 1-977 2-979 1-498 1-769 0-890 0-896


Medium 2-024 1-593 3-066 1-063 2-413 0-836 0-895
Large 1-555 1 -484 3-981 0-871 3-807 0-833 0-838

160
7.3.3 Effect of Orientation on the Time Constant

Equation 7.20 gives

T DArr b •

Figure 7.13 shows that inclining the channel changes both the effective area for mass transfer and the

effective throat length, due to the buoyancy effect, and

_ v b i T « * e
To j ,

Vo ->
T,= ^ c o s 2 0 , (7.22)

^r — cos 2 6 , (7.23)
To

where 0 is as shown in figure 7.13. For d = 45 ° equation 7.23 gives

Lr = cos 2 45 ° = 0 - 5 ,
o

which is in good agreement with the data in table 7.7.

Equations 7.20 and 7.22 assume that mass transfer through the neck is diffusional, and therefore

only apply for 0 < tan 1 -jgj-

vertical

Figure 7.13 Cross-sectional area and diffusion distance for an inclined pore.

161
7.4 Mass Transfer Results — Partially Miscible Fluids

The water-butan-l-ol experiments were conducted in the same micromodel as the miscible
studies and under conditions of both gravity stability and gravity instability, i.e.:

• horizontal channel, pores vertically above channel and initially saturated with butanol (the
less dense phase), x- and y-axes horizontal (figure 7.1), z-direction vertically upwards;

• horizontal channel, pores vertically below channel and initially saturated with butanol, x-
and y-axes horizontal, z-direction vertically downwards.

In the first experiment mass transport within the pore throats was diffusional. However in the
second case, the fluid within the pore was less dense than that in the channel and convective
transport contributed to the total mass transfer. The alcohol phase did not become
hydrodynamically unstable because of the resisting forces of capillary pressure. Example
photographs are given in figure 7.14.

The positions of the fluid-fluid interface within the three pores at successive times are
shown in figures 7.15 and 7.16 for the gravity stable and unstable cases respectively. The
volume of the alcohol rich phase was calculated by planimeter measurements (see section
7.2.1) since the depth was uniform, and given in tables 7.9 and 7.10. The alcohol phase
volume is plotted against time in figures 7.17 and 7.18 for the two experiments. Equation 7.16
gives

V V
2 ~ t t _ KmAT PgQ-^P! , y
l
VP vp vvlpl-w2p2u i)-

If phase 2 is the alcohol rich phase and component a is water, then the data shown in figure 7.4 give

PgQ-*>\P\ = 0 • 9983 - 0 • 922 x 0 • 9874 = 0 . n 8 7 = 1


wlpl-w2p2 0-922x0-9874-0-8480x0-2 ' 8-426'

Figures 7.17 and 7.18 show that the mass transfer coefficient is initially high while the
fluid-fluid interface is in the pore neck, but reaches a constant but lower value when V2<VB.
The gradients of the lines for V2<VB were used with equation 7.16 to calculate the mass
transfer coefficients, Km, which are given in table 7.11. The assumption of steady state
diffusion through the throat used in section 7.2.2 for the miscible case, gives D=Km-i; for
the gravity stable case, which is also given in table 7.11. The low value for the large pore
implies that this is not valid.

162
Figure 7.14 Two-phase mass transfer.

163
10 620

Figure 7.15 Position of the fluid-fluid interface at successive times; gravity stable.

164
72 000

19 380

13 3 2 0 H •9120

7320
11 7 0 0 ' '6300
3 960 2910
2 040 5100
660
120' 195
60-
•15

Figure 7.16 Position of the fluid-fluid interface at successive times; gravity unstable.

165
Table 7.9 Gravity stable.

VJVp

(») Small Medium Large

30 0 88 - 1-00

70 0 91 1 00 0-99

135 0 78 0 97 0-98

270 0 65 0 93 0-98

495 0 54 0 88 0-97

660 0 49 0 86 0-97

750 0 46 0 86 -

840 0 42 - -

1065 0 39 0 84 0-97

1320 0 32 0 83 -

1680 0 26 0 80 0-96

2010 0 15 0 78 0-96

2400 0 08 0 74 0-96

2580 0 01 0 71 -

3120 0 67 0-95

3 660 0 62 0-95

4320 0 55 0-94

5 220 0 45 0-93

6540 0 36 0-92

7 800 0 28 0-91

9180 0 15 0-89

10620 0 07 0-87

12105 0 01 0-84

166
Table 7.10 Gravity unstable.

VJVp

(S) Small Medium Large


15 0 •85 - 0-99
60 0 •71 0 93 0-98

120 0 •64 0 91 0-98


195 0 •52 0 89 0-98

300 0 •41 0 88 0-97

435 0 •34 0 85 -

660 0 •23 0 83 0-97

900 0 • 12 0 80 -

1080 0 •08 0 77 0-97

1230 0 •01 0 75 -

1500 - 0 70 0-96

2040 - 0 61 0-96

2910 - 0 48 0-95

3 960 - 0 34 0-94

5100 - 0 20 0-91

6 300 - 0 08 0-87

7 320 0 - 01 0-83

9120 - 0-75

11700 - - 0-63

13 320 - - 0-56

14820 - - 0-51

16020 - - 0-46

17430 - - 0-41

19380 - - 0-35

72000 - - 0-02

167
Figure 7.17 Gravity stable.

Figure 7.18 Gravity unstable.

Equation 7.21 gives £>=1 -OX 1 0 - 9 m 2 s - 1 in the aqueous phase at 20°C. Published
data101 for the diffusivity in the alcohol rich phase show a composition dependence with the
coefficient varying between 1 - 2 4 X 1 0 " 9 and 0-27X 1 0 - 9 m 2 s - 1 at 30°C; equation 7.21 gives
D= 1-48X 1 0 - 9 m 2 s - 1 at 30°C.

168
Table 7.11

Km

{stable)
Pore Grav. stable Grav. unstable Kj?table)ST
Kjunstable)
(X10"6m5-1) ( X 1 0 ~ 6 m j -1 ) (xl0~9m2
Small 2-26 4-49 1-018 0-50
Medium 2-30 4-17 1-209 0-55

Large 1-13 5-15 0-633 0-22

Table 7.11 shows that the rate of mass transfer for the gravity unstable case is between
two and five times the value for the gravity stable experiment. If the mass transfer were
controlled by convective transport through the throat, the coefficient would probably be larger
than the values in table 7.11. For this case, the rate of mass transfer may be limited by
diffusion across the interface. It is, however, interesting to note that the rate for the large
pore is initially equal to the rate for the gravity stable case, before increasing to the value at
larger times. This seems to imply that as the interface recedes, convective transport becomes
more important.

The diffusive flux across the fluid-fluid interface is controlled by the composition gradient
at the interface in both phases. Convective transport within each phase due to buoyancy
forces, increases the gradients at the interface and therefore increases the diffusive flux across
it. Thus, if the rate of mass transfer is controlled by diffusion across the fluid-fluid interface,
the mass transfer coefficient is larger for the gravity unstable case.

7.5 Summary

Mass transfer into dead-end pores for the two fluid systems studied, miscible (decane-
propan-l-ol) and partially miscible (water-butan-l-ol) can be characterised by mass transfer
coefficient equations. The value of the coefficient for the miscible case is dependent on the
diffusivity, and the throat length, ^ and also the orientation of the pore. For gravity stable
cases, buoyancy forces due to the varying fluid density, tend to make the isocomposition
profiles horizontal, and therefore affect both the cross-sectional area for mass transfer and the
effective diffusion distance. In the case of gravity instability, extra convective mixing is
induced.

For the partially miscible case, the alcohol rich phase in the pores did not become
hydrodynamically unstable under an unfavourable density gradient, because of the resisting
forces of capillary pressure. Convective transport due to buoyancy forces increased the value
of the mass transfer coefficient.
169
CHAPTER EIGHT

SUMMARY AND CONCLUSIONS

In this chapter the major achievements of this work are summarised, and the conclusions are

given.

The physics of fluid transport processes within porous media have been studied at the

microscopic pore scale using micromodels. Computer graphics together with the nylon etching

process have enabled networks to be made with controlled pore shapes and sizes. This is an

advance upon any previous work. Although the micromodel networks are two-dimensional, their

high coordination number and realistic pore morphology retained the necessary conditions of

fluid velocity, viscous drag forces, potential gradients and capillary pressure. The recovery of

oil from bypassed zones by capillary pressure displacement, due to IFT gradients has been

analysed and shown to be a significant recovery mechanism for heterogeneous media.

Holographic interferometry was applied to micromodel processes in order to quantify mass

transfer on the pore scale. This is the first time that this technique has been used on such a

microscopic scale. These mass transfer studies demonstrated that buoyancy forces play a

greater role in the rate of transport than previously thought, and that the dynamic interfacial

forces, particularly capillary pressure, are very important in the remobilsation of "residual oil".

Sections 8.1 to 8.3 discuss these points in greater detail.

8.1 Constant IFT Displacement Processes

8.1.1 Summary

The control over pore and network geometries enabled a detailed experimental study of

the role of pore shape and size on displacement processes. Although influenced by the local

pressure field, the structure of the residual oil was found to be governed principally by the

pore geometry, and the pore aspect ratio was found to be the important factor. The pressure

field is dependent on the flow rate, pore sizes, fluid viscosity and density, and capillary

pressure. The pore geometry strongly influences the shape (and curvature) of the fluid-fluid

interfaces, and thereby the capillary pressure. There is a critical value of the pore aspect ratio

below which no residual oil is formed. This critical value was found to be equal to two for

the pore throat shape used, but is approximately three for a toroidal throat shape. At high

aspect ratios, forward snap-off of oil during drainage displacement (water-wet) was also
8.1.2 Conclusions

It is imperative that the mechanisms of residual oil formation are fully understood before

considering EOR processes, because improved recovery is dependent on the structure of the

r immobile fluids. In heterogeneous media oil can be bypassed in relatively large volumes due to

variations in permeability and capillary pressure. The pore scale analysis has explained some of

the physics behind relationships between average properties, for instance between residual oil

saturation and capillary number. It is clear that these relationships are not quantifiable by one

universal curve, but will depend on the particular rock sample.

8.2 Variable IFT EOR Processes

8.2.1 Experimental Studies — Summary

Decane-water-alcohol systems were used to study aspects of EOR processes on the pore
scale. Although at high alcohol concentrations these systems are miscible, initially mobilisation
fc will occur by reduced interfacial tension. When there were IFT gradients, i.e. the IFT was

non-uniform over the system, oil ganglia were observed to flow in a direction opposite to the
general flow. This is explained in terms of capillary pressure: a greater reduction in IFT, and
thereby capillary pressure at the rear end of the ganglion induces a potential gradient within
the oil, and consequently causes flow in the backward direction. The interfacial curvature is
influenced by the pore shape and therefore as the upstream meniscus moves backward into the
throat, the interfacial curvature and therefore the capillary pressure both increase. If the
threshold capillary pressure at the upstream throat is insufficient to stabilise the ganglion, it
will pass through this throat and continue to move backward, until it reaches a throat
** sufficiently small to entrap it, or another ganglion with which it coalesces. The ganglion may

now be displaced forward along a different, and more favourable pathway.

8.2.2 Conclusions
r

The observed dynamic capillary pressure effects occur because the oil and the
surrounding aqueous phase are not in chemical (thermodynamic) equilibrium. These are
dependent on the pore structure, the ganglion size, the concentration field in the aqueous
6 phase, which can be expressed in terms of dispersion, and mass transport to the interfaces of
the immobile oil and within the stagnant fluid. Under reservoir conditions small residual

171
ganglia will probably be in approximate equilibrium with the aqueous phase. However,
longitudinal and transverse equilibrium may not be a valid approximation for bypassed zones
of oil. If the interfacial tension gradient along the boundary between the bypassed and mobile
zones is sufficient, the oil will become hydrodynamically unstable and consequently flow into
the mobile zone. Capillary pressure displacement has been shown to be an important recovery
mechanism when the bypassed zone length, or ganglion length, is greater than about one tenth
of the mixing zone length.

8.2.3 Theoretical Studies — Summary

A theoretical study of diffusional mass transfer into bypassed layers was conducted in
order to assess the effect of transverse mass transport on the concentration gradient along the
layer boundary, and therefore on the probability of capillary pressure displacement. A finite
difference solution technique was used to compute the concentration field.

8.2.4 Conclusions

The results showed that mass transfer increases the longitudinal concentration gradient, by
a factor dependent on the relative layer thickness and the characteristic times for transverse
equilibrium, Teq, and convective transport in the mobile zone, 7y. When Teq<^Tj the
longitudinal gradient was increased by W/W— £ where 2£ and 2(W— £) are the thicknesses of
the immobile and mobile layers respectively. When req>Ty and transverse equilibrium was not
approximated, the longitudinal gradient was not significantly greater than when req<KTf.
However, when req^±TjL/lm, the longitudinal concentration gradient was at a maximum
(where L is the length of the bypassed zone and lm the injected longitudinal mixing zone
length if no mass transfer occurred). When lm^>L, longitudinal equilibrium is a valid
approximation, and dynamic capillary pressure displacement would not occur.

8.3 Mass Transfer in Porous Media

8.3.1 Summary

In the theoretical model of mass transfer into bypassed layers there were some necessary
assumptions which required qualification, especially that each zone behaved as a fluid
continuum. This in essence ignores the structure of the solid matrix. Transverse mass transfer

172
was assumed diffusional, thus convective transport due to buoyancy forces was ignored. An

experimental study of mass transfer in micromodel networks was therefore conducted to help

qualify these assumptions. Mass transfer into dead-end pores was investigated for both a

miscible system, decane-propan-l-ol, and a partially miscible system, water-butan-l-ol. Real

time holographic interferometry was used to map the composition field for the miscible case.

This technique detects changes in fluid refractive index, which was used to determine fluid

composition. A uniform depth micromodel was used to ensure that the composition fields were

two-dimensional, and also to simplify the analysis. The same apparatus was used to investigate

the partially miscible case, however the interferometry was not used quantitatively, because

there was insufficient range in refractive index. The rate of mass transfer was determined

from measurements of the relative phase volumes.

8.3.2 Conclusions

For both fluid systems the density was dependent on composition, but the isocomposition

lines were also isodensity lines. These lines were found to be horizontal, because of convective

transport due to buoyancy forces. Even when transport through the dead-end pore throat was

diffusional, the rate of transfer was found to be dependent on the orientation of the pore. This

was because the buoyancy effect controls both the cross-sectional area for mass transfer and

the effective diffusion distance. Mass transfer in both fluid systems was modelled using mass

transfer coefficient equations. In the miscible system the mass transfer coefficient correlated

with the pore size parameter, the molecular diffusivity and the pore orientation. For the

partially miscible system, the rate of mass transfer was controlled by the rate of diffusion

across the fluid-fluid interface. This rate increased when there was convective transport within

each phase due to unstable density gradients. Convection increases the concentration gradients

at the interface and therefore increases the diffusive flux. The alcohol rich phase did not

become hydrodynamically unstable under unstable density gradients, because of the resisting

forces of capillary pressure. Dynamic capillary pressure effects were not present because the

interfacial tension was constant.

The experimental studies showed that mass transport into simple dead-end pores can be

characterised by a mass transfer coefficient, which is composed of the molecular diffusivity,

the throat cross-sectional area, the pore volume, and a factor dependent on the orientation of

the dead-end pore with respect to the gravitational field. The tortuosity factor is therefore not

a constant, but is influenced by buoyancy forces. Nevertheless, mass transfer into bypassed

zones may often be characterised by an effective diffusivity, provided that the buoyancy effect

is accounted for when the value of the coefficient is assigned.


8.4 Significance of These Studies

This thesis shows that the microscale mechanisms of oil recovery are strongly dependent

on the structure of the porous medium, especially the pore aspect ratio and the degree of

heterogeneity. Capillary pressure controls fluid snap-off and entrapment, and fluid mobilisation i

by reduced IFT. For real, heterogeneous media, oil can be trapped in relatively large zones.

The dynamic capillary pressure displacement mechanisms described in this work are an

important method for recovery of this bypassed oil. Dynamic capillary pressure effects arise

because the entrapped oil is not in thermodynamic equilibrium with the surrounding fluid.

Equilibrium will only be approximated when the solute mixing zone length is much greater,

i.e. more than ten times, than the bypassed zone length. Correct scaling of mass transfer is

critical to the full description of these mechanisms. The experimental studies described here

have shown that the buoyancy effect increases the rate of mass transfer and therefore, scaling

by simple diffusion will always underestimate the mass transfer coefficient and overestimate

the equilibrium time constant. This will be significant for systems such as C0 2 -oil-water,

where the fluid density depends on the composition. This gravity effect is most pronounced

when the boundary between the bypassed and mobile zones is not horizontal.

8.5 Suggestions for Future Work

1. Micromodel networks could be further developed with a greater variation in throat and

body sizes, perhaps in a random, but known manner. The residual oil distribution could

then be experimentally examined and compared with a theoretical network model

prediction. Larger pore networks could also be made in order to aid the scaling of

displacement processes to larger porous media.

2. The mass transfer, holographic interferometry studies could be extended to micromodels

with dead-end zones of inter-connecting pores. When these zones are inclined to the

horizontal, and the fluid density is not constant, then there will be convective mass

transport due to the buoyancy effect. Angles of orientation other than 0° and 45° could

also be examined for the dead-end pore model. Networks with smaller pores could be ^

made by polishing the glass cover slip down to a thickness less than 100 P m, thereby

reducing the pore depth and also the minimum pore width. Decreasing the pore depth

does, however, reduce the resolution of the interferometric method, and fluids with a

greater range in refractive index would have to be used. Mass transfer at constant

density could be studied by using benzene (density of 0-9297g/cm 3 , and RI of 1-5011)

and ethyl butyrate (density of 0-932g/cm 3 , and RI of 1-3922), which are totally

miscible.

174
3. Further micromodel work on dynamic capillary pressure displacement from bypassed zones

would be valuble to further quantify and scale the effect. Modification of models of core

scale displacements to incorporate dynamic capillary pressure effects must be attemped.

Further micromodel experiments could be conducted at the outlet end of a core to

examine the effluent fluids and their behaviour.

4. The effect of the solute mixing zone length in the displacing phase could be examined

experimentally by introducing a known concentration profile into the micromodel. The

profile can be controlled by displacement through a long capillary tube attached to the

micromodel inlet, thereby exhibiting Taylor dispersion.

5. Further alcohol-oil-water systems could be used to study displacement and mass transfer

effects for different phase behaviour (as characterised by the ternary diagram).

8.6 Final Remarks

Clearly the physics of oil recovery is complicated, but this work has shown that a fuller

description can be obtained by investigating each aspect separately in order to produce an

integrated model. The microscale properties must be adequately described before considering

the problems of extending to higher orders of scale as indicated in figure 1.1. The major

problem of heterogeneity can then be treated by some form of statistical process. The

micromodel networks developed in this work have been valuable in describing the microscopic

mechanisms of fluid processes in porous media. The micromodel studies help to answer

specific questions, such as the role of pore morphology, but also help to pose the questions

that need to be answered in order to correctly describe and scale the physics of oil recovery.
REFERENCES

1. HALDORSEN, H.H. and LAKE, L.W.; "A New Approach to Shale Management in
Field Scale Simulation Models", SPE Ann. Fall Conf., New Orleans, L.A. (September
1982); SPE paper 10976.

2. WEBER, K.J.., EIJPE, R., LEIJNSE, D. and MOENS, C.; "Permeability Distribution
in a Holocene Distributary Channel-Fill Near Leerdam (The Netherlands)", Geologie en
Mijnbouw. 51 (1972) 53-62.

3. WEBER, K.J.; "Influence of Common Sedimentary Structures on Fluid Flow 'in


Reservoir Models", JPT 34 (March 1982) 665-672.

4. VAN-QUY, N., SIMANDOUX, P., and CORTEVILLE, J.; "A Numerical Study of
Diphasic Multicomponent Flow", Soc. Pet. Eng. J. 12 (1972) 171-184.

5. HELFFERICH, F.G.; "General Theory of Multicomponent, Multiphase Displacement in


Porous Media", Soc. Pet. Eng. J. 21 (February 1981) 51-62.

6. HIRASAKI, G.J.; "Application of the Theory of Multicomponent, Multiphase 4


Displacement to Three-Component, Two-Phase Surfactant Flooding", Soc. Pet. Eng. J. 21

(1981) 191-204.

7. DAKE, L.P.; "Fundamentals of Reservoir Engineering", Elsevier (1978).

8. AMYX, J.W., BASS, D.M. and WHITING, R.L.; "Petroleum Reservoir Engineering,
Physical Properties", McGraw-Hill (1960). ft
9. WRIGHT, R.J. and DAWE, R.A.; "Surfactant Slug Displacement Efficiency in
Reservoirs", 1981 European Symp. on Enhanced Oil Recovery, Bournemouth, England
(1981) 161-178, Elsevier.

10. WRIGHT, R.J. and DAWE, R.A.; "Fluid Displacement Efficiency in Layered Porous ^
Media; Mobility Ratio Influence", Revue Inst. Franc, du Petrole (July/Aug. 1983) 455-
474.

11. CRANK, J.; "The Mathematics of Diffusion", 2nd Edition, Oxford University Press
(1975).

12. BIRD, R.B., STEWART, W.E. and LIGHTFOOT, E.N.; "Transport Phenomena", Wiley M
(1960).

13. HERSHEY, D.; "Transport Analysis", Plenum Press (1973).

14. SKELLAND, A.H.P.; "Diffusional Mass Transfer", Wiley (1974).

15. SHERWOOD, J.N.; "Diffusion Processes", Proc. Thomas Graham Memorial Symp., 0
Univ. of Strathclyde, (1971) Gordon and Breach.

16. HANSON, C.; "Recent Advances in Liquid-Liquid Extraction", Pergamon (1971).

17. BEAR, J.; "Dynamics of Fluids in Porous Media", Wiley (1972).

18. BUCKLEY, S.E. and LEVERETT, M.C.; "Mechanism of Fluid Displacement in Sands", *
Trans. AIME 146 (1942) 107-116.

176
19. WELGE, H.J.; " A Simplified Method for Computing Oil Recovery by Gas or Water
Drive", Pet. Trans. AIME 195 (1952) 91-98.

20. WRIGHT, R.J. and DAWE, R.A.; "An Examination of the Multiphase Darcy Model of
Fluid Displacement in Porous Media", Revue de I'Inst. Franc, du Pet. 35 (1980) 1011-
1024.

21. MOHANTY, K.K., DAVIS, H.T. and SCRIVEN, L.E.; "Thin Films and Fluid
Distributions in Porous Media", Surface Phenomena in Enhanced Oil Recovery Ed. D.O.
Shah, Plenum Press (1981) 595-609.

22. WRIGHT, R.J.; "The Structural Mechanism of Drying in Kaolinite films", Ph.D. thesis,
Univ. of Bristol (1974).

23. STEGEMEIER, G.L.; "Mechanisms of Entrapment and Mobilisation of Oil in Porous


Media", Improved Oil Recovery by Surfactant and Polymer Flooding, Ed. D.O. Shah
and R.S. Schecter, Academic Press, Inc., New York (1977) 55-91; 81st Nat. Meeting
AIChE, Kansas City (1976).

24. CHATZIS, I. and MORROW, N.R.; "Correlation of Capillary Number Relationships for
Sandstones", 56th SPE Ann. Fall Conf. San Antonio, Texas (1981); SPE paper 10114.

25. AMAEFULE, J.O. and HANDY, L.L.; "The Effect of Interfacial Tensions on Relative
Oil/Water Permeabilities of Consolidated Porous Media", Soc. Pet. Eng. J. 22 (1982)
371-381.

26. SHAH, D.O.; "Fundamental Aspects of Surfactant-Polymer Flooding Process", Proc.


1981 European Symp. on EOR, Bournemouth, England (September 1981) 1-41, Elsevier.

27. MORROW, N.R. and SONGKRAN, B.; "Effect of Viscous and Buoyancy Forces on
Nonwetting Phase Trapping in Porous Media", Surface Phenomena in Enhanced Oil
Recovery Ed. D.O. Shah, Plenum Press (1981), 387-411.

28. RICHARDSON, J.G., HARRIS, D.G., ROSSEN, R.H. and VAN HEE, G.; "The
Effect of Small, Discontinuous Shales on Oil Recovery", J. Pet. Tech. 30 (1978) 1531-
1537.

29. PRATS, M.; "The Influence of Oriented Arrays of Thin Impermeable Shale Lenses or
of Highly Conductive Natural Fractures on Apparent Permeability Anisotropy", J. Pet.
Tech. 24 (1972) 1219-1221.

30. HIBY, J.W.; "Longitudinal and Transverse Mixing During Single-Phase Flow Through
Granular Beds", Interaction between Fluids and Particles, London, Inst. Chem. Engrs.,
312-325.

31. HELLER, J.P.; "Observations of Mixing and Diffusion in Porous Media", Proc. 2nd
Symp. on Fundamentals of Trasnsport Phenomena in Porous Media Univ. Guelph,
Ontario, Canada (1972).

32. PERKINS, T.K. and JOHNSTON, O.C.; "A Review of Diffusion and Dispersion in
Porous Media", Soc. Pet. Eng. J. 3 (1963) 70-84.

33. NUNGE, R.J. and GILL, W.N.; "Mechanisms Affecting Dispersion and Miscible
Displacement", Ind. and Eng. Chem. 61 no.9 (1969) 33-49.

34. WARREN, J.E. and SKIBA, F.F.; "Macroscopic Dispersion", Soc. Pet. Eng. J. 4 (1964)
215-230.

177
35. TAYLOR, G.I.; "Dispersion of Soluble Matter in Solvent Flowing Slowily Through a
Tube", Proc. Royal Soc. A219 (1953) 186-203.

36. TAYLOR, G.I.; "Conditions Under Which Dispersion of a Solute in a Stream of


Solvent can be Used to Measure Molecular Diffusion", Proc. Royal Soc. A225 (1954)
473-477.

37. ARIS, R.; "On the Dispersion of a Solute in a Fluid Flowing Through a Tube", Proc. ^
Royal Soc. A235 (1956) 67-77.

38. BLACKWELL, R.J.; "Laboratory Studies of Microscopic Dispersion Phenomena", Soc.


Pet. Eng. J. 2 (1962) 1-8.

39. HELLER, J.P.; "The Interpretation of Model Experiments for the Displacement of
Fluids Through Porous Media", A.I.Ch.E. J. 9 no.4 (1.963) 452-459. .

40. SAFFMAN, P.G.; "Dispersion in Flow Through a Network of Capillaries", Chem. Eng.
Sci. 11 (1959) 125-129.

41. SAFFMAN, P.G.; " A Theory of Dispersion in a Porous Medium", J. Fluid Mech. 6
(1959) 321-349. ^

42. SAFFMAN, P.G.; "Dispersion due to Molecular Diffusion and Macroscopic Mixing in
Flow Through a Network of Capillaries", J. Fluid Mech. 7 (1960) 194-208.

43. BLACKWELL, R.J., RAYNE, J.R. and TERRY, W.M.; "Factors Influencing the
Efficiency of Miscible Displacement", Trans. AIME 216 (1959) 1-8. ^

44. LEVENSPIEL, O. and FITZGERALD, T.J.; " A Warning on the Misuse of the
Dispersion Model", Chem. Eng. Sci. 38 no.3 (1983) 489-491.

45. VON ROSENBERG, D.U.; "Mechanics of Steady State Single-Phase Fluid Displacement
from Porous Media", A.I.Ch.E. J. 2 no.l (1956) 55-58.

46. KOONCE, H.T. and BLACKWELL, R.J.; "Idealised Behaviour of Solvent Ranks in
Stratified Reservoirs", Soc. Pet. Eng. J. 5 (1965) 318-328.

47. MARLE, C., SIMANDOUX, P., PACSIRSZKY, J. and GAULIER, C.; "Etude du
Displacement de Fluides Miscibles en Milieu Porous Stratifie", Revue de L'Institut
Francois du Petrole 22 no.2 (1967) 272-294. ^

48. LAKE, L.W. and HIRASAKI, G.J.; "Taylor's Dispersion in Stratified Porous Media",
Soc. Pet. Eng. J. 21 (1981) 459-468.

49. ARONOFSKI, J.S. and HELLER, J.P.; " A Diffusion Model to Explain Mixing of
Flowing Miscible Fluids in Porous Media", Pet. Trans. AIME 210 (1957) 345-349.

50. BRIGHAM, W.E.; "Mixing Equations in Short Laboratory Cores", Soc. Pet. Eng. J. 14
(1974) 91-99.

51. LEVENSPIEL, O. and SMITH, W.K.; "Notes on the Diffusion-Type Model for the
Longitudinal Mixing of Fluids in Flow", Chem. Eng. Sci. 6 (1957) 227-233.

52. TURNER, G.A.; "The Flow-Structure in Packed Beds", Chem. Eng. Sci. 7 (1958) 156- •
165.

53. WAKEMAN, R.J.; "Diffusional Extraction from Hydrodynamically Stagnant Regions in


Porous Media", Chem. Eng. J. 11 (1976) 39-56.

178
54. COATS, K.H. and SMITH, B.D.; "Dead-end Pore Volume and Dispersion in Porous
Media", Soc. Pet. Eng. J. 4 (1964) 73-84.

55. SALTER, S.J. and MOHANTY, K.K.; "Multiphase Flow in Porous Media:
1. Macroscopic Observations and Modelling", SPE Ann. Fall Conf. New Orleans, L.A.,
(1982); SPE paper 11017.

56. STALKUP, F.I.; "Displacement of Oil by Solvent at High Water Saturation", Soc. Pet.
Eng. J. 10 (1970) 337-48.

57. SHEARN, R.B. and WAKEMAN, R.J.; "Theoretical Mass Transfer Models for
Assessing Tertiary Recovery by Miscible Fluid Displacement", 1978 European Symp. on
EOR Edinburgh, Scotland (1978) 253-269.

58. VAN DEEMTER, J.J., ZUIDERWEG, F.J. and KLINKENBERG, A.; "Longitudinal
Diffusion and Resistance to Mass Transfer as Causes of Nonideality in
Chromatography", Chem. Eng. Sci. 5 (1956) 271-289.

59. VAN DER POEL, C.; "Effect of Lateral Diffusivity on Miscible Displacement in
Horizontal Reservoirs", Soc. Pet. Eng. J. 2 (1962) 317-326.

60. HELLER, J.P.; "Onset of Instability Patterns between Miscible Fluids in Porous Media",
J. Applied Physics 37 no.4 (1966) 1566-1579.

61a MAHERS, E.G., WRIGHT, R.J. and DAWE, R.A.; "Visualisation of the Behaviour of
EOR Reagents in Displacements in Porous Media", 1981 European Symp. on Enhanced
Oil Recovery, Bournemouth, England (1981) 511-526, Elsevier.

61b DAWE, R.A. and WRIGHT, R.J.; "Oil Reservoir Behaviour - The Microscale", Royal
School of Mines J. 33 (1983) 25-30.

62. MAHERS, E.G. and DAWE, R.A.; "The Role of Diffusion and Mass Transfer
Phenomena in the Mobilisation of Oil during Miscible Displacement", 1982 European
Symp. on EOR Paris, France (1982) 279-288, Editions Technip.

63. EGBOGAH, E.O. and DAWE, R.A.; "Microvisual Studies of Size Distribution of Oil
Droplets in Porous Media", Bull. Can. Pet. Geol. 28 no.2 (1980) 200-210.

64. SCHECHTER, R.S., LAM, C.S. and WADE, W.H.; "Mobilisation of Residual Oil
Under Equilibrium and Nonequilibrium Conditions", SPE Ann. Fall Conf. San Antonio,
Texas (1981); SPE paper 10198.

65. NG, K.M., DAVIS, H.T. and SCRIVEN, L.E.; "Visualisation of Blob Mechanics in •
Flow Through Porous Media", Chem. Eng. Sci. 33 (1978) 1009-1017.

66. CHATENEVER, A. and CALHOUN, J.C.Jr.; "Visual Examinations of Fluid Behaviour


in Porous Media - Part 1", Pet. Trans. AIME 195 (1952) 149-156.

67. MATTAX, C.C. and KYTE, J.R.; "Ever See a Water Flood?", Oil and Gas J. 59 no.42
(1961) 115-128.

68. MICHAELS, A.S., STANCELL, A. and PORTER, M.C.; "Effect of Chromatographic


Transport in Hexylamine on Displacement of Oil by Water in Porous Media", Soc. Pet.
Eng. J. 4 (1964), 231-9; Pet. Trans. AIME 231 (1964) 231-9.

69. DAVIS, J.A. and JONES, S.C.; "Displacement Mechanisms of Micellar Solutions", J.
Pet. Tech. 20 (1968) 1415-1428.
70. W A R D L A W , N.C.; "The Effects of Pore Structure on Displacement Efficiency in
Reservoir Rocks and in Glass Micromodels", 1st Joint SPE/DOE Symp. on EOR, Tulsa,
Oklahoma (1980) 346-352; SPE paper 8843.

71. W A R D L A W , N.C.; "The Effect of Geometry, Wettability, Viscosity and Interfacial


Tension on Trapping in Single Pore-Throat Pairs", J. Can. Pet. Tech. 21 no.3 (1982) 21-
27.

72. W A R D L A W , N.C. and McKELLAR, M.; "Mercury Porosimetry and the Interpretation
of Pore Geometry in Sedimentary Rocks and Artificial Models", Powder Technology 29
(1981) 127-143.

73. McKELLAR, M. and WARDLAW, N.C.; "A Method of Making Two-Dimensional Glass
Micromodels of Pore Systems", J. Can. Pet. Tech. 21 no.4 (1982) 39-41.

74. CHATZIS, I. and DULLIEN, F.A.L.; "Dynamic Immiscible Displacement Mechanisms


in Pore Doublets: Theory versus Experiment", J. Coll. Int. Sci. 91 (1983) 199-222.

75. BONNET, J. and LENORMAND, R.; "Constructing Micromodels for the Study of
Multiphase Flow in Porous Media", Revue de L'lnst. Francois du Petrole 42 (1977) 477-
480.

76. LENORMAND, R.; "Displacements Polyphasiques en Milieu Poreux sous l'lnfluence des
Forces Capillaires; Etude Experimentale et Modelisation de Type Percolation", L'Institut
National Polytechnique de toToulouse, France (1981), Ph.D. Thesis.

77a ARRIOLA, A., WILLHITE, G.P. and GREEN, D.W.; "Trapping of Oil Drops in a
Noncircular Pore Throat", SPE Conf. Dallas, Texas (1980); SPE paper 9404.

77b ARRIOLA, A., WILLHITE, G.P. and GREEN, D.W.; "Mobilisation of an Oil Drop
Trapped in a Noncircular Pore Throat upon Contact with Surfactants", SPE Ann. Fall
Conf. Dallas, Texas (1980); SPE paper 9405.

78. ARRIOLA, A., WILLHITE, G.P. and GREEN, D.W.; "Trapping of Oil Drops in a
Noncircular Pore Throat and Mobilisation Upon Contact with a Surfactant", Soc. Pet.
Eng. J. 23 (1983) 99-114.

79. PLATEAU, J,; "The Figures of Equilibrium of a Liquid Mass Withdrawn from the
Action of Gravity", Ann. Report, Smithsonian Institute, no. 83 (1864) 228-285.

80. RAYLEIGH, Lord; "On the Instability of Jets", Proc. Lond. Math. Soc. 10 (1879) 4-13.

81. RAYLEIGH, Lord; "On the Instability of a Cylinder of Viscous Liquid under Capillary
Force", Phil. Mag. 34 (1892) 145-154.

82. TAYLOR, G.I.; "The Formation of Emulsions in Definable Fields of Flow", Proc. Royal
Soc. London A146 (1934) 501-23.

83. TOMOTIKA, S.; "On the Instability of a Cylindrical Thread of a Viscous Liquid
Surrounded by Another Viscous Fluid", Proc. Royal Soc. London A150 (1935) 322-337.

84. HAINES, W.B.; "Studies in the Physical Properties of Soil - Part V. The Hysteresis
Effect in Capillary Properties, and the Modes of Moisture Distribution Associated
therein", J. Agric. Sci. 20 (1930) 97.

85. ROOF, J.G.; "Snap-off of Oil Droplets in Water-Wet Pores", Soc. Pet. Eng. J. 10
(1970) 85-90.
86. MOHANTY, K.K.; "Fluids in Porous Media: Two-Phase Distribution and Flow", Univ.
of Minnesota, USA (1981), Ph.D. Thesis.

87. CHATZIS, I., MORROW, N.R. and LIM, H.T.; "Magnitude and Detailed Structure of
Residual Oil Saturation", Soc. Pet. Eng. J. 23 (1983) 311-326.

88. STALKUP, F.I.; "Status of Miscible Displacement", J. Pet. Tech. 35 (1983) 815-826.

89. HIRASAKI, G.J.; "Scaling of Non-Equilibrium Phenomena in Surfactant Flooding", I st


Joint SPE/DOE Symp. on EOR Tulsa, Oklahoma (1980) 323-343; SPE paper 8841.

90. SAWISTOWSKI, H.; "Interfacial Phenomena", Recent Advances in Liquid-Liquid


Extraction, Ed. C. Hanson, Pergamon (1971) 293-366.

91. SORENSEN, T.S.; "Marangoni Instability at a Spherical Interface", J.C.S. Faraday II,
76 (1980) 1170-1195.

92. MORROW, N.R. and CHATZIS, I.; "Measurement and Correlation of Conditions for
Entrapment and Mobilisation of Residual Oil", U.S. DOE Final Report (1981)
DOE/BETC/3251-12.

93. DUROU, C.; "Une Installation d'lnterferometric Holographique Destinee a l'Etude des
. Phenomenes de Diffusion dans les Liquides", J. Phys. E. Sci. Inst. 6 (1973) 1116-1120.

94; GABOR, D.; "A New Microscopic Principle", Nature 161 (1948) 777-778.

95. GABOR, D.; "Microscopy by Reconstructed Wave-fronts", Proc. Royal Soc. A197 (1949)
454-487.

96. GABOR, D.; "Holography, or the 'Whole Picture'", New Scientist 29 (1966) 74-78.

97. LEITH and UPATNIEKS; "Wavefront Reconstruction with Continuous-Tone Objects", J.


Opt. Soc. Am. 53 (1963) 1377-1381.

98. PHILLIPS, N.J. and PORTER, D.; "An Advance in the Processing of Holograms", J.
Phys. E. Sci. Inst. 9 (1976) 631-634.

99. GOODNIGHT, R.C., KLIKOFF, W.A. and FATT, I.; "Non-Steady-State Fluid Flow
and Diffusion in Porous Media Containing Dead-End Pore Volume", J. Phys. Chem. 64
(1960) 1162-1168.

100. WILKE, C.R. and CHANG, P.; "Correlation of Diffusion Coefficients in Dilute
Solutions", A.I.Ch.E. J. 1 (1955) 264-270.

101. JOHNSON, P.A. and BABB, A.L.; "Liquid Diffusion in Non-Electrolytes", Chem. Revs.
56 (1956) 387-453.

181

S-ar putea să vă placă și