Você está na página 1de 28

Advances in Colloid and Interface Science.

6 (1976) 1’73~-200 173


@ Elsevier Scientific Rlblishing Company, Amsterdam - Printed in The Netherlands

RHEOLOGY OF CONCENTRATED DISPERSIONS

J. MEWIS
Katholreke Universltert te Leuven, Leuven (Belgium)
A.J.B SPAULL
Brunel Unrwersity, Uxbridge (Gt. Britain)

CONTENTS

A. Introduction ....................... 174


B. Rheological behaviour of concentrated dispersions ........... 175
(i ) Viscous phenomena ................... 175
(ii) Plartlc phenomena .................... 177
(iii) Elastic phenomena ................... 176
C. Parameters in dispersion mechanics ............... 180
(i) Hydrodynamic effects ............. , .. 181
(a) Spherical particles ................. 181
(b) Nonspherical part&es ................ 185
(Ii) Brownian movement ................... l&S
(iii) Attractive forces ................... 185
(iv) Repulsion and stability .................. 186
(a) Electric repulsion ................... 187
(b) Steric repulsion .................... 188
(c) Rheological effects of repulsion .............. 188
D. Structure and rheology .............. 189
(i) Interparticulate bonds in dis~er~on~ : ............. 189
(Ii) Structure ....................... 190
(iii) Elasticity in concentrated dispersions - - - _ _ _ _ _ . _ _ . . 192
(iv) Flow phenomena .................... 194
(v) Relation with other physical properties ............ 195
E. Summary ........................ 195
References ......................... 196

ABSTRACT

Concentrated dispersions constitute an important class of materials, not only in science


but also in technology, yet complete analysis of their rheological behaviour is still lacking.
Available experimental data clearly underline the very general nature of the mechanical
response encountered in eoneentrated dispersions. Thus, dispersions could be useful in
fundamental rheological studies of the various constitutive equations. The analysis of dii-
persion mechanics in terms of underlying phenomena is sufficiep.ly advanced to show a
strong interrelationship between rheology and colloidal statiility. The variety of structcres
found links both. On the one hand, colloidal effects determivie the interpartiele forces
during deformation and consequently the variable structure and the rheological character-
btics. On the other hand, suitable rheological measurements, which may usefully be
supplemented with those obtained using other techniques, can elucidate forces and struc-
tures involved as shown by analyses in terms of variable spectra and other basic visco-
elaalic measurements_

A. INTRODUCTION

This review deals with the mechanical behaviour of heterogeneous systems


containing solid particles having a size or size distribution, at least in one direc-
tion, within the range defined as the colloidal subdivision of matter, i.e.,
between 1 pm and 1 nm [l] _ The terminology recommended by IUPAC [l]
has been followed as far as possible. In view of the present discussion, the fol-
lowing concepts are defined explicitly.
(1) Suspension: a heterogeneous two-phase system, of buoyant particulate
solid, not colloidal, distributed in a continuous liquid.
(2) Dispersion: a two-phase solid in liquid system, the particles of which are
colloidal. In this manner, we discriminate between suspensions and dispersions.
(3) Aggregate: a tight cluster of particles, strongly held together, so that it
cannot be separated in a given rheological esperiment.
(4) Floe or flocculate: a loose, thermodynamically metastable, structure from
colloidal particles. A floe may be destroyed, to a greater or lesser extent, by
the application of stresses in a given rheological experiment and may reform
when these stresses are removed. Some authors discriminate further between
weak, open structures, floccules, and tight but weak structures, agglomerates.
For the present purpose such a distinction is not essential, as all weak struc-
tures can be discussed in common_
The concentrations considered here are such that particle interactions be-
come predominan!, in the resistance to flow and deformation. The lower con-
centration limit will be governed by the particular systems under considera-
tion but can be as low as 1%. Aqueous and nonaqueous systems are dealt with,
but more attention is paid to the latter.
Emphasis is given to the general aspects of the mechanical behaviour of
disperse systems, not just viscosity, and their relationship with the physical
phenomena that govern flow, the models of colloidal stability being included_
In comparison with those recently published, the present review differs in
its--scope and points of importance. On the one hand, Brenner [Z] and Batch-
elor [3] focused attention on the theoretical rheology of dilute suspensions;
emulsions have been dealt with by Barry in this journal [ 41. Mason gives a
bibliography of his work on the microrheology of simple suspensions [ 51;
viscous flow of dispersions has been treated by Goodwin [ 63 _ On the other
hand, Ottewill [7a], Sonntag and Strenge [ 7b], and Lyklema [ 7c ] stress
the developments in the theoretical models of stability_ This review puti
dispersion rheology at the crossing of continuum mechanics and stability
theory, which together provide the essentiaI elements in understanding and
in describing the rheology of such systems_ a
B. RHEOLOGICAL BEHAVIOUR OF CONCENTRATED DISPERSIONS

Before an attempt is made to quantify or explain some data, the available


information about the general nature of the mechanical response in dis-
persions is surveyed. This response does not fit any of the simple classes of
continua. It can be solid-like or liquid-like, viscous, plastic or elastic; a single
sample may even show a combinatiqn of all these distinct responses f 8 J.
Probably all rheological effects known can be produced with dispersions of
Hookean solids in Newtonian liquids, including those often associated with
polymeric components.

(i) Viscous phenomena

When a constant shear rate is applied to a concentrated dispersion, the


resulting shear stress usually tends to a constant value. Hence an apparent
viscosity can be defined for this liquid-like behaviour. In the case of concen-
trated dispersions, there are limitations to the simple test described [9],
but they do not affect the present discussion, in that the stress level is con-
sistently higher than expected from suspension rheology.
This higher level is illustrated in Fig. 1, where an average result for spheri-
cal suspensions [lo] is compared with some selected data for two series of
dispersions containing more or less spherical particles. Under suitable con-
ditions, the relative viscosities even surpass the values represented in Fig. 1
by orders of magnitude If compared at identical concentrations [ 121. In

Fig. 1. Increased stress levels in flowing dispersions as compared with suspensions. 1. Aver-
age curve for suspensions of spheres [lo]; 2. TiOz (0.1 pm) Enstand oil (4 + 0) [ 111; 3. Car-
bon black in mineral oil (4 = 0.7 s-l).
176

each case, the stress levels clearly differ from suspension data. Consequently,
there are sources of resistance to flow in dispersions, the existence of which
indicates the need to treat them as a separate class of materials in rheological
studies. The fact that their rheological properties can be affected by changing
the colloidal characteristics provides additional justification for such an *
approach.
Apart from the magnitude of the shear stress, its change with shear rate is
important. The majority of disperse systems show pronounced non-Newton-
ian behaviour. Only at the lower limit of the concentration range of interest
here can Newtonian fluid properties he found. A cross-section through the
possible patterns is obtained if the particle concentration is systematically
increased. A synthesis of published data on various systems 1s @en in Fig. 2,
and represents qualitatively the average effect of concentration on the
rheogram.
At low concentrations, the viscosity is nearly independent of shear rate.
As the concentration increases, a shear-thinning zone develops, separating
two Newtonian regions, the zone extending to lower shear rates with in-
creasing concentration. The viscosity of the Newtonian region at low shear
rates is more sensitive to concentration than that at high shear rates.
Similar rheograms are encountered with polymers [ 13,141 and dilute sus-
pensions of nonspherical particles [ 14,151. Quite often a region develops

Fig. 2. Changes in the shape of the rheogram at increasing concentration in dispersions.


(Particle concentration increases from 1 to 6.)
,

177

where the viscosity decrease is inversely proportional to the shear rate (Fig. 2,
curves 5 and 6); consequently, the shear stress becomes independent of shear
rate. This plastic region will be discussed below. If the dispersion is a real
liqutd. a Newtonian region must exist at low shear rates. Some materials
behave like solids under small stresses and then the viscosity tends to infinity
at zero shear rate [8].
Starting from a critical concentration, often around 40% for near spherical
particles [ 16-181, a shear-thickening or dilatant region develops at higher
shear rates. It becomes steeper with increasing concentration. When measure-
ments have been extended to sufficiently high shear rates, an upper limit for
dilatancy has been encountered by several investigators [17-211, above which
the material behaves in a shear-thinning or a Newtonian manner again. Vis-
cosity measurements usually become impossible before the upper concentra-
tion limit of closest packing is reached. Particular systems might deviate in
one way or another from the general pattern outhned above. One obvious
group consists of dispersions in non-Newtonian media. The relative viscosity
seems to follow qualitatively the same evolution as in Newtonian media [ 221.
The global viscosity, however, will reflect flow anomalies of the continuous
and the disperse phases, resulting in complex rheograms. Dintenfass has dis-
cussed some of the possible flow curves for such systems [23].
A rheogram represents only the equiltbrium behaviour under shear. One of
the features of concentrated dispersions is the appearance of considerable
transients upon instantaneous changes in the kinematics of the flow. Most
often a sudden application of shear or a sudden increase in shear rate causes
an initial overshoot in stress, followed by a progressive decrease towards the
equilibrium value.
Tne viscosity changes described here are reversible; effects due to ageing
are excluded from the discussion. Reversible decrease of viscosity under shear
is usually called thixotropy [ 241. Slightly different definitions of this concept
are discussed by Reiner and Scott-Blair 1251. Thixotropy is attributed to a
reversible breakdown of structure under shear and is encountered frequently
in dispersions [13,19,23,26,27]. A reversible increase of viscosity with time
under shear, antithixotropy or rheopexy, is less common 1241 and is often
confused with shear thickening. Nevertheless, the physical reality of antithixo-
tropy has been proved beyond doubt [19,24,28]. It could be argued that on
the one hand shear thinning and thixotropy and on the other hand shear
thickening and antithixotropy emanate from the same changes in structure,
and therefore there is no need to introduce new concepts for time dependency.
However, for the moment, it seems useful to introduce the concepts of thixo-
tropy and antithixotropy wherever the emphasis is on time effects.

(ii) Plastic phenomena

The viscous phenomena could be discussed with little reference to idealized


rheological constitutive equations. Plasticity is more closely related to a model
178

description. A material is considered ideally plastic if it does not flow as long


as the shear stress is below a critical value rO, the yield stress. Under flow, the
shear stress remains at the critical value irrespective of the shear rate applied.
This constancy of shear stress under plastic flow is identical with a viscous
flow, where the apparent viscosity decrease is inversely proportional to the
shear rate. It has been mentioned above that various dispersions approach this
kind of behaviour, as seen in the flow curve, over a limited range of shear rates.
In models for non-ideal plastic materials, some change of shear stress with
shear rate is taken into account [29]. It can be concluded that plastic flow is a
special and limiting case of viscous flow.
More specific is the existence of a yield stress. Few investigators have at-
tempted to make measurements under low and constant values of shear stress
in order to prove a no-flow condition. A notable exception is the work of
Rehbinder and his associates [ 81. They proved solid-like behaviour below a
critical stress in real systems. The maximum or overshoot value of the transient
shear stress, after a sudden application of shear, is often assumed to provide
a value for the yield stress. Although helpful in describing the behaviour of the
material, the overshoot technique cannot be used to prove the presence of a
plastic-yield value.
In various papers, 7. is calculated by fitting part of the flow curve to one of
the models for plastic behaviour. Such a value of r. is often called the apparent
yield value. It is only a model parameter of the flow curve and is not a direct
measure of the stress limit for flow, as can be seen by suitable experiments
[8,30]. The different physical interpretations of “absolute” and “apparent”
yield stresses should be kept in mind if one wants to relate r,, to the fundamen-
tal properties of a colloidal system. Some plastic dispersions are thixotropic-
as well. Then the model parameters of the flow curve might depend on the
kinematic history, including ro.
A critical strain corresponds to a critical stress of a plastic material. It is an
important parameter in understanding the deformation mechanism in concen-
trated dispersions. Nevertheless, there are few systematic studies of this param-
eter [ 31). Often, it is vrry small, 10m3 or lower, [8,32-341 which complicates
the measurements. It constitutes an important parameter in industrial dis-
persions.

(iii) Elastic phenomena

From an energetic point of view, viscous phenomena result from dissipative


processes. Elasticity, on the contrary, implies the means to store mechanical
energy reversibly during deformation. The presence of macromolecules is not
necessary to cause dispersions to behave elastically. The more concentrated ones
show a solid-like behaviour at low stresses and strains for which an elasticity or
a shear modulus can be determined [ 8,35,36]. Its value varies considerably
according to the kind and the concentration of the components. In low viscous
media, values between 10 and 10’ NmD2 have been recorded.
-r__ .
179

rog kw)

Fig. 3. Formation of one or two plateau regions in the mo$ulun-frequency curves for dis-
persions in Newtonian (1.) and viscoelastic (2.) media (-G , . ..C ). 1. carbon black in
mineral oil [ 391.2. carbon black in polystyrene solution [-IO].

Transient and oscillatory tests indicate that dispersions combine elastic


and viscous elements and produce a viscoeiastic response. Such behaviour can
be completely described either by a spectrum of relaxation times or the time or
frequency function of the modulus [ 37,383. Mechanical spectral analysis is
extremely useful in analyzing the structure of matter. It has been used extensive.
ly on polymers [ 381, and an increasing number of results on concentrated
emulsions are becoming available [ 4 1. For dispersions, two typical results are
shown in Fig. 3.
Curve 1 represents data taken on a system having a Newtonian fluid as the
medium. The viscoelasticity resides completely within the structure of the
dispersed phase. The curve resembles that of cross-linked polymers [ 38 1, the
reason being that a three-dimensional structure exists in both cases. The
presence of a plateau region has been demonstrated repeatedly in solid-like
disperse systems [3,39-411. In general, and in particular in the plateau region,
the moduli become larger if the concentration of the disperse phase is increased
[ 32,36,42,43]. At higher concentrations, the increase in moGuli becomes steep,
as is the case with viscosity (section (i)). Often, the apparent yield value and the
shear modulus are affected in a similar way by concentration changes [31, 42-
44]. In addition, both increase with decreasing particle size [42,45,46 1.
Curve 2 of Fig. 3 has been obtained on a dispersion in a viscoelastic medium.
Again a plateau develops, owing to the dispersed phase (lower plateau), while
the plateau of the medium also appears in the dispersion (higher plateau),
180

though shifted in level. For non-structurized, liquid-like dispersions, the lower


plateau is missing, but the moduli still increase with concentration. At higher
frequencies, beyond the structural plateau, the difference between the two
kinds of dispersion becomes smaller [ 39,411. Clearly, the effect of network
stticture becomes less important in this frequency region. As a result, the high-
frequency moduli of dispersions in viscoelastic media change less with concen-
tration than the low-frequency moduli [47].
Under steady-state shear flow, elasticity reveals itself by normal stress dif-
ferences. With dispersions, they are often difficult to measure. The few avail-
able data nevertheless provide a picture consistent with oscillatory experiments
[ 11,481. Structurized dispersions show considerable normal stress differences,
which increase faster with concentration than the shear viscosity. At higher
shear rates, where the structure is more broken down, the normal stresses
approach the values for the continuous phase. As soon as the dispersion
structure is completely broken down, the viscosity and normal stress curves
are parallel to those of the medium [ 111.
The structure can also be destroyed by use of oscillations with an amplitude
above the critical value. Destruction of the structure results in nonlinear oscilla-
tory behaviour [39,49,50]. As is to be expected, the moduli decrease with
increasing amplitude, while the loss tangent increases or goes through a
maximum [ 44,491. In some cases, the stress amplitude depends on strain, but
the periodic change remains sinusoidal [51]; in most cases, however, the stress
pattern is strongly distorted. The difference can be explained on the basis of
of structural kinetics, but supporting evidence is still lacking. It is particularly
difficult in these experiments to avoid instrumental artifacts.
Thixotropy and elasticity can exist simultaneously in dispersions. Thixo-
tropic dispersions have been shown to be viscoelastic, at least under certain
conditions [8,40,52]. The viscoelastic properties will then depend on the
kinematic history, as seen from the changing viscoelastic spectrum 140,531
and from the time dependency of nonlinear vihrations [40,50].
It can be concluded that there is a considerable similarity, at least qualita-
tively, between disperse and polymer systems. The quantitative differences
stem from the divergent nature of the structure in both types of material [ 521.
The plateau modulus changes between wider limits in dispersions and depends
more strongly on concentration; structure is also more fragile, increasing the
nonlinearity of the mechanical response. The rates of structural changes are
normally much slower than in polymer systems, thus extending the time scale
over -which transient and thixotropic effects are noticeable.

C. PARAMETERS IN IXSPERSION MECHANICS

Distinction is made between hydrodynamic and colloidal stability effects.


The latter are determined by the physicochemical bulk and surface properties
of the components. Eydrodyn&Gc effects result from the mere presence of
solid particles in the flowing medium. The resulting perturbation of the flow
181

field causes an increase in energy dissipation as expressed by a modification


of the bulk rheological properties. At low shear rates, there is no effect of
absolute size on the hydrodynamic behaviour. Therefore, the results of sus-
pension rheology can be applied to dispersions, without giving, however, a
complete description of their mechanical response. Brownian movement is a
size-dependent, colloidal effect, which is also dealt with, for various reasons,
in suspension rheology.

(i) Hydrodynamic effects

The average stress in a flowing suspension depends on the geometrical and


mechanical characteristics of the components, their relative concentration and
the flow field. The continuous phase affects the flow through its viscosity qm,
the rigid particles through their concentration, size distribution and shape. If
inertia effects have to be included, the densities must be considered. As with
other non-Newtonian systems, the shear viscosities cannot be generalized as
such to other flow fields [Z&3].
The theoretical aspects of suspension rheology have been reviewed recently
by Brenner [2] and by Batchelor [3]. The available analyses deal with lower
concentrations than those of interest here. Nevertheless, they provide informa-
tion about some factors affecting the flow of more concentrated systems.

a. Spherical particles
The spherical shape is more tractable than any other and has been covered
more extensively. For a monodisperse system, the dependence of Newtonian
viscosity on volumetric concentration c is usually presented by a power series
expression [ 54,551 for the relative viscosity TJ,:

t7r = 17hm = 1 + klc + k2c2 + k3c3 + . .. (1)


The factor k, determines the linear concentration range where there is no
interaction between particles. Its value has been calculated by Einstein to be
2.5 1561. The higher-order factors k, are often explained as representing the
additional effect of simultaneous collisions of n particles at increasing concen-
tration. It has been pointed out by Batchelor that there is no theoretical
evidence that concentration effects should be most suitably expressed in such
a way [3].
Detailed calculations are available for the c* term only [ 541. The most
recent and complete analysis by Batchelor and Green [ 571 gives a value of
7.6. It is estimated that truncation at the c? term is valid for concentrations
up to 10%. Experimental values of k2 are usually above the theoretical value
[10,54]. Most factors that could affect this experimental determination would
tend to overestimate k2 , and so the discrepancy is not surprising. As soon as
particle interactions control viscosity, one might expect effects from shear
rate and from the type of flow, i.e., shear, extensional, compression, and
.

182

transient, as they will influence the interactions and consequently the higher-
order terms.
For higher concentrations, attempts have been made to approximate the
complex interactions either by using cell models [58,59] or by analyzing
near-contact hydrodynamics [60]. As their accuracy is as yet unknown, one
must rely cn experiments, from which empirical relationships have also been
derived. The data show considerable scatter, particularly at high values of c.
An average curve is shown in Fig. 4. The two main features are an increasing
sensitivity of qr to c and a tendency for viscosity to rise towards infinity when
c reaches about 0.70. The latter is explained by the existence of a maximum
packing for spheres in that range. The crowding of par%les around that con-
centration results in an interlocking, preventing the suspension from flowing.
Empirical and semi-empirical relations have been suggested in great number.
Rutgers and Sherman have reviewed the earlier work [54,55].
The concentration affects not only the Newtonian viscosity but also causes
non-Newtonian behaviour. Shear thinning owing to volumetric effects in
suspensions of spherical particles is not very pronounced and has not been
studied systematically [ 31. However, at high concentrations, particles are close
enough for weak dispersion effects to become noticeable. Interparticle
forces can give rise to structure formation, even in suspensions, cat;sing the
viscosity to increase [ 3,61,62]. For values of c greater than approximately 0.4-
0.5, shear thickening can appear. Packing and interlocking provide some ex-
planation for it. For dispersions. stability effects seem to dominate [ 16--181.
Except for steady-state shear flow, time-dependent types of flow show some
interesting pecuL,‘-ities. They can be illustrated by means of oscillatory flow.
In concentrated suspensions, the particle interactions might depend on frequen-
cy and amplitude. At high frequencies, where the interaction is minimal, the
viscosity-concentration curve will consequently be flatter than in Fig. 4. Little
experimental infcfmation about this region is available. The data of Hellinckx
and Mewis on dispersions support the picture given [47]. The oscillatory flow
mode can also be used to investigate inertia effects. If the frequency is sys-
tematically increased, the particle movement eventually lags behind that of
the medium and causes a phase shift for the stress. This results in a viscoelastic
response that depends on size [63]. A similar result holds for inertia effects under
steady-state flow [ 641.
In the rheology of monodisperse suspensions containing large spherical
particles, the absolute size enters the calculations only in case of inertia. In con-
trast, in heterodisperse systems the distribution of particle size can be of primary
importance because it determines collisions and packing [65,66]. The possibility
of a much closer packing as compared with monodisperse systems shifts the
critical concentration to higher values and reduces the relative viscosity at fixed
values of c. The drop is particularly pronounced in bimodal mixtures of large
and small particles where the smaller particles fit in be&ween the larger [67-
691. With a known mixture, the relative viscosity can be obtained by multiplying
the relative viscosity of the fractions rjr,i, using the viscosity of the mixture of
183

smaller size fractions as the medium viscosity for the next fraction [ 68,69 ]

77, = nit7r,, (2)

For the continuous size distributions, the result is less pronounced.


A last variable is the medium. The relative viscosity should be independent of
the medium, at least for suspensions in Newtonian media. Viscoelastic media
have been considered in a few investigations. The presence of particles increases
their viscosity and normal stress [ 11,221, qualitatively in a similar manner as in
Newtonian media. Quantitatively, some differences are found, which remain un-
explained [ 6,22 J_

(b) Nonspherical partictes


In addition to size and concentration, shape becomes a factor now. It has an
effect such that geometry and interaction are not completely determined by
the relative position of the particles alone; their orientation must also be
allowed for. Orientation can be expressed by orientational distribution func-
tions. Theoretical work has been mainly confined to the case of dilute suspen-
sions of axisymmetric particles. Jeffrey presented the solution for spheroids [?O]
He concluded,that spheroids rotate around an axis perpendicular to the plane
of shear. The particles spend more time with their largest axis parallel to the
flow direction than perpendicular to it. Anisometry tends to increase alignment
in the flow direction.
Jeffrey’s wvorkhas heen extended and experimentally verified by Anczurowski
and Mason. They included transient behaviour and cstahlished the effect of colli-
sions on orientation distribution functions [ 'il--731.From t!~ calculated orien-
tation distribution function, the viscosity and the normal stress differences can
be computed [74,75]. Increasing the shear improves the alignment in the dlrec-
tion of shear, which reduces the energy dissipation. Hydrodynamic particle
interaction appears at lower concentration with nonspherical particles; it also
seems to promote alignment. Inertial effects work in the opposite direction
[76,77]. They are governed by a particle Reynolds number Re,

Re, = L2?P,/I)m (3)


in which L is the length of the particle and pm the medium density. A value
of 0.01 for this number is sufficient to markedly affect the flow pattern.
Finally, the rheological nature of the medium has been shown to have an
effect on the distribution function and consequently on the viscosity [77].
III extensional flow, interaction becomes pronounced, even at low concen-
trations, if the aspect ratio is large enough, i.e., for fiber-like particles. This
effect was predicted theoretically by Batchelor [78] and verified experi-
mentally by Mewis and Metzner [ 79 1.
(ii) Brownian movement

Small particles distributed in a fluid undergo a statistical movement,


translational and rotational, that effectively acts as a dispersive mechanism.
184

Fig. 4. Average viscosity-concentration curve for spherical suspensions, showing the effect
of spherical inclusions in the absence of other than volumetric effects.

This action is determined by the diffusion coefficients D, and D, for transla-


tion and rotation, respectively [56b]
x--2 =D,C
and (4)
e2 =D,t

where ‘;;* and a2 are the quadratic mean displacement and rotation and t is
time. The diffusion coefficients express the Brownian movement as a balance
between the acting thermal movement and the viscous resistance. Thus for
spheres
Dt = kT/67rq0R; and (5)
D, = kT/8nr),R3
where k is the Boltzmann constant, T the absolute temperature and R the
radius of the spheres.
Brenner S*asrecently discussed Brownian movement effects in translation
and rotation with axisymmetric particles [80]. Results on particles of general
shape will be published [ 811. F rom the point of view of rheology, Brownian
movement introduces particular properties in a dispersion. It entails a ten-
dency towards more random radial and orientational distribution functions,
which in turn will af<ect the stress. The additional angular velocity (Fig. 4)
also makes a direct contribution to the stress [ 51,82;.
One of the main reasons why Brownian movement is considered in suspen-
sion rheology is the fact that it makes the equilibrium distribution functions
185

independent of the initial conditions [2,3]. In dispersion rheology, Brownian


movement is of direct importance due to the smaller diameter of the particles.
During flow, the effects of Brownian movement and shear are superimposed.
Their relative contributions are expressed by translational and rotational
Ptklet numbers, which for spheres are [Z]
(6)

At low shear rates, i.e., low P&let numbers, riiffusion dominates and prevents
any time-average alignment of anisometric particles with respect to the shear
plane. At higher shear rates, the ordering effect of shear becomes systematical-
ly more important. As a result non-Newtonian, shear-thinning, behaviour
appears. Theoretical results are available only for dilute suspensions. Sheraga
has shown the extent of shear thinning that can be expected on this basis for
spheroids of different aspect ratio [ 153. Transient behaviour due to Brownian
movement has been dealt with by Hinch and Leal [ 831. Krieger has made an
attempt to apply some of these principles on more concentrated dispersions of
spheres [ 84 1.
Randomization of the distribution functions due to diffusion is a time-
dependent process. It brings the dispersion back to a random equilibrium con-
dition, and, as such, serves as a “memory” to this reference situation. Memory
corresponds to elasticity In mechamcs. The viscoelastlc response to be expected
from Brownian movement has been calculated in steady-state shear flow, i.e.,
normal stress differences [ 851 and under oscillatiops, i.e., dynamic shear
modulus [ 821.
The physical phenomena described until now suffice to explain qualitatively
the various rheological phenomena encountered m dispersions, with the ex-
ception of yield stress. Available quantitative information is limited to dilute
systems, but, even so, considerable differences between theory and experiment
can be found (see section B). Therefore, additional contributions to the internal
stresses must be considered.

(iii) A t trac tive forces

Particles dispersed in liquids will be in an attractive field as a result of the


van der Waals forces existing between them. The expression is a generic one
describing the total attraction as a result of interaction arising between perma-.
nent, induced and instantaneous dipoles or some combination of these.
In nearly all types of atoms and molecules, the greatest contribution to the
total attraction arises through the interaction of instantaneous dipoles. The
effects of multipolr? interaction other than between dipoles are small or negli-
gible. The attraction between two particles can, in principle, be calculated by
summing up the contributions from all the atoms and molecules constituting
the individual particles, the route from interactions on an atomic and molecu-
186

lar level to their macroscopic effect being known. As the rheological behaviour
of concentrated dispersions depends on the interparticle attraction, the factors
that influence these forces should be considered: the geometry of the particles,
the species adsorbed on their surfaces, the nature of the liquid medium, and
the influence of the surrounding particles [ 7,861.
For a geometry of two spherical particles of equal size in vacuum, the poten-
tial energy of attraction V,, when the inter-particle distance is not large, is
given by

11
=-- x* +2x
f 2 In ( (7)
va 12
4 [ x*+2x
1 + X2 +2X+1
1 -42 +!2x+1
where A, is the Hamaker constant for the particle material and x = d/D,
where d is the distance between the surfaces of the particles [ 871. The general
expression for an arbitrary particle geometry, with fluid surrounding the
particles, can be thrown into the form
V, = AIHg (8)
where Ai is the interaction parameter and H, the geometrical parameter, which
contains the effect of shape and interparticle distance. Apart from the dis-
tance dependency, the attractive force depends on the medium: generally the
following approximation is used .

Ai = (A;/* -AZ*)* (9)


where A, is the Hamaker constant for the medium. In a further extension,
the effect of adsorbed layers could be allowed for [ 881.
7%~ attraction energy between two particles must be influenced by the
presence of surrounding particles in a concentrated dispersion. Little is known
about the changes induced in this manner. Flocculation calculations based
on floe diameter rather than on the diameter of the elementary particle re-
sults in erroneous stability predictions [ 891. Gosiewski attempted a calcula-
tion of interaction effects using a cell model technique [go].
Longuet-Higgins and Widom [91] suggested that in liquids far from the
critical point, the short-range repulsive forces (see below) determine the struc-
ture while the longer-range attraction forces merely provide an almost uni-
form background potential that holds the liquid together. This idea can be
extended to dispersions. The rheological consequences of the London-van
der Waals forces are more efficiently discussed together with repulsive forces.
Suffice it tosay that the attractive forces will oppose deformation; under
shear, the potential energy will increase due to separation of the particles.
These concepts will be used to explain high viscosities, shear thinning,
viscoelasticity and thixo’tropy.
(iv) Repulsion and stability
Interaction betw;?en colloidal particles is composite, consisting of an
attractive and a repulsive part. The original theory of these interactions
187

Fig. 5. Potential mteraction energy between particles as a function of distance, showing two
possible, stable, interparticle distances, one with strong and another with weak interaction
forces.

was given in two publications, one by Derjaguin and Landau 192 ] and the
other by Verwey and Overbeek [ 931, hence the term DLVO theory. The
main development over the years has been to include with the repulsion aris-
ing from electrical interaction that due to steric interaction [7a, 94 1. The
opposing forces can counterbalance each other, thus limiting aggregation or
flocculation and permitting a stable dispersion to exist.
The sum of London-van der Waals and repulsion energy equals the total
potential energy. The shape of the curve representing the total interaction
energy between two particles as a function of distance depends on the values
of the various parameters [7b]. A possible form of particular interest in
rheology is given in Fig. 5. This figure shows a deep primary minimum and a
less pronounced secondary one, separated by an energy barrier. If the two
particles are at a suitable distance, corresponding to the potential wells,
interaction is strong enough to keep them at this distance.

(a) Electric repulsion .


When a particle, having an electrostatic surface charge, or in some instances
no charge, is dispersed in a continuous liquid, an electric double layer may
be formed on the liquid side of the solid-liquid interface, consisting of ions
of different sign. In a first approximation, it consists of two parts, an inner
part close to the surface of the particle and an outer diffuse part, the ions of
which can be exchanged with those of the bulk. The interaction depends on a
pumber of influences, e_.g., the shape of the particles, dielectric constant of
th? bulk liquid and the concentration of electrolyte in the bulk. Methods of
caIculating the force between spheres have been given by Bell and Peterson
[ 951, and their direct experimental determination was reviewed by Ottewill
188

[7a]. The electric repulsive interaction potential V, is given, with qualification,


by
V, = ED@; ln[ 1 + exp(-Kc!)] /4 (10)
where E is the dielectric constant of the medium, $e the surface potential
and K the Debye-Hiickel constant [7b].
There are a number of problems that are of interest in rheology currently
engaging the attention of colloid chemists. Different assumptions about the
boundary conditions at the particle surface have been made for the calculation of
the potential as the distance between the particles is varied; constant surface
charge has been applied instead of constant surface potential [ 96,971. As
with London-van der Waals forces, the surrounding particles must be taken
into account in concentrated dispersions. -L-vine and co-workers are attacking
this problem using a cell model [ 861.

(b) Steric repulsion


Particulate material dispersed in aqueous and nonaqueous media can be
stabilized by the adsorption of polymers and surface-active agents as a result
of what is called steric repulsion. Usually two mechanisms are distinguished.
If the adsorbate concentration is relatively low in the adsorption layer, the
layers can interpenetrate during the mutual approach of two particles, causing
an increase in free energy. In that case repulsion occurs. With high adsorbate
concentration, mixing is less probable; the adsorbate layers undergo elastic
deformation during collision, also causing repulsion. Some recent reviews are
available [7a,94]. Theory is mainly concerned with static situations, i.e.,
theory is not available for the effect of movement of particles on their
steric repulsive interaction. Vincent has questioned the correctness of additivi-
ty of attractive and steric repulsion forces [ 88 1.

(c)_Rheological effects of repulsion


Three electroviscous effects are associated [ 655 ] with electric repulsion.
The primary effect resalts from the distortion of the ionic atmosphere around
the particles under flow. During collision, the repulsive forces keep the parti-
cles further apart, comparable with an apparent increase in particle size; this
is the secondary effect. A similar effect appears with steric repulsion [ 981. The
tertiary effect, arising from changes in size and shape of flexible macromolec-
ular ions, is outside the scope of this review.
The global result of any interparticle potential energy field must be to affect
the interparticle distances. In particular, if there are potential wells (Fig. 5)
these could be preferential distances under suitable energetic conditions. Of
special importance to rheology is the secondary minimum. If it exists, particles
can form flocculates, which can be broken down under the mechanical action
of shear. This reversible mechanism entails a whole range of rheological effects.
The presence of flocculates will necessitate large stresses during flow, increasing
the viscosity. The deflocculation under shear will cause shear thinning, where-
189

as the subsequent reflocculation explains thixotropy. If the floes throughout


the sample unite into one structure, the dispersion becomes solid. Under small
deformations, elastic behaviour is to be expected; for flow to occur, the poten-
tial energy must be overcome, indicating the presence of real yield stresses.
The flocculation-deflocculation situation also will cause viscoelasticity in a
way similar to the combined effect of Brownian motion and shear (section (ii)).
It must be concluded that stability theory provides the basic elements in under-
standing the rheological behaviour of concentrated dispersions. A large body of
experimental work supports this conclusion.
Systematic rheological studies useful in physicochemical research are relative-
ly rare except for clay minerals [99,100]. As an illustration, a few results of
general interest are quoted. The effect of electrolyte concentration has been
found to follow, at least qualitatively, the predictions from potential-energy
considerations [ 18,101,102] ; the same holds for the electrokinetic mobility
[97,103], the dielectric constant [104,105] and the role of adsorbed surfac-
tant and polymers [106-1081.

D. STRUCTURE

(i) In terpartrculate bonds m dispersions

The elements of the previous section can be combined to approach rheologi-


cal behaviour through a physical picture of structure. ip dilute dispersions and
suspensions, the radial and orientational distribution functions can be consld-
ered to describe “structure”. In concentrated systems, the interparticulate inter-
actions thoroughly change the picture. They force the system to a configura-
tion having least potential energy. From stability theory, it is concluded that
this condition corresponds to well-defined interparticle distances. Further,
knowledge of the shape of the potential energy curve permits calculation of
the force or energy necessary to separate the particles to a distance where the
interaction becomes small and thus where the direction of the interparticle
forces can easily be changed by an applied external couple. Hence, the concept
of bond can be used and defined on a physical, potential basis. Although chem-
ical bonds, especially hydrogen bonds, may occur between particles, they are
not considered here, as evidence suggests that the physical bond dominates in
the rheology of dispersions. Some of the evidence suggests that the rheology
can be explained only in terms of a bond having a reversible character with
respect to stress - one that stress destroys, yet which readily reforms when
stress is removed.
It has been mentioned above that two energy minima are distinguished in
colloidal science, the primary and the secondary. Insufficient shielding of the
primary minimum by an energy barrier inevitably leads to irreversible aggrega-
tion. In some dispersions, application of a shear causes the particles to surmount
the barrier and reach the primary minimum [ 1091.
The existence of reversible structures seems to require the presence of a
190

secondary minimum and a suitable barrier to the primary one. In addition,


in order to be physically meaningful the bond must be able to persist for
some time. For this reason, the depth of the energy well must amount to
several kT units of thermal energy. In determining the stability of a dis-
persion, particle concentration is a necessary factor. Only above a critical
concentration is flocculation observed; its value may be used to determine
the depth of the minimum in the potential energy interaction [llO].
Other bonds might esist in the dispersion; even if they are not affected by
the shear stresses, they contribute to structure and rheological behaviour
and interfere with the interpretation of rheological data. In particular, the
solid elements of the dispersed phase often consist, at least partially, of
aggregates of elementary particles strongly bound together, with or even
without liquid between them 11111. This situation could arise as a result of
an incomplete dispersion operation or insufficient stabilization of dispersed
particulate material and is usually referred to as “degree of dispersion” [ 112 1.
It changes the shape, size and properties of the flow units and cons lqently
the global rheological behaviour 1113,114 3. However, the bonding In the
aggregates remains unchanged during flow or deformation. In the following
discussion, particles are understood to be the flow units that cannot be
broken down under stress; they are either elementary particles or aggregates.

(ii) Structure

If particles can form stable physical bonds, the possibility then exists for
the formation of stable structures. Intuitively, and to a frrst approximation,
two possibilities suggest themselves; a linear growth leading to chain-like
structures, as in an outstretched necklace, and a random growth, which gives
rise to a more spherically-shaped cluster of particles. Chains could develop
further into a network. These two shapes are to be considered extreme simpli-
fications of reality. In practice, various intermediate situations might be ex-
pected. At high concentrations, the difference between the two cases must
disappear.
For the existence of flocculated structures, many reported examples of
powerful evidence exist. The most obvious method of detection is that of
direct observation. Four are quoted [115-1181. The changing structure can
also be detected and analyzed by various physical techniques: electric con-
ductivity [115,119], dielectric [116,120] and optical [121] measurements.
For most techniques, a detailed analysis of the data is not yet possible for
concentrated dispersions, but by comparison of rheological and other
physical data it can be concluded that isolated structural elements of variabie
size and shape exist in some cases and continuous elements (networks) in
others.
Structure formation carries a statistical effect, which limits the regularity
to be expected. This lack of regularity complicates the determination of the
spatial arrangement of particles. Flocculation kinetics is somewhat similar
191

to aggregation kinetics, from which it has been derived_ Initial aggregate


size calculations were based on random collision between flow elements
leading to a permanent union, These assumptions were first applied by von
Smoluchowski in his kinetic theory of rapid coagulation [ 1221. The result
is a time-dependent distribution function for size
Iv,(t) = N,(0)(t/&)k-l/(l + t/t,)k+’ (11)
where ATkis the number of aggregates containing I: particles and t, the time
necessary to reduce the total number of floes to half the original number
(N, (0)). Overbeek has discussed some improved versions of this theory
[123].
The problem of particle arrangement within the aggregate has been
tackled by computer simulation [ 1241. Sutherland has produced some inter-
esting results, based on von Smoluchowski’s theory. He has shown the
possibility of chain formation [125], and also that the average shape of
the aggregate is spheroid&l with an aspect ratio of thrse [ 1261. Chain
formation seems to have been underestimated m these simulation studies.
This probably stems from neglecting the effect of the shape of growing
particles on the potential energy.
In order to explain shear thinning and thlxotropy, the influence of shear
rate on flocculation kinetics should be known. Shear will affect the collision
frequency (1271, it will induce breakdown of flocculates [ 128,129], and it
can change the stability condltlons [ 19,130J. First, only the number of
bonds was considered [131--1331. Casson introduced an average length of
rod-like flocculates as a function of shear rate 11341. It should be pointed out
that Casson assumed flocculation to occur as a result of association in the
primary minimum. In later studies, distribution functions under shear were
calculated, mainly in rheological studies [135-1381. Difficulties arise due
to the lack of information about the real structure of the flocculates and
the corresponding susceptibility to rupture under shear. Chain-like structures
have been dealt with by van den Tempel 11351 and by Ruckenstein and Mewis
[136]. Weymann and Mercer [137,138] considered sphere-like flocculates.
Van den Tempel’s treatment considered only the equilibrium structure under
shear; the other two studies contain in addition the transient distribution
functions. For the Ruckenstein-Mewis model, they can be expressed as

Nk(l) = &-l(f) [ l- Y] (12)


where S,(t) is the total number of flocculates and S, the total number of
elementary particles. S,(t) is also a function of shear stress. All mod& are
based on simple assumptions about the sequence in which bonds rupture.
In chains, all bonds are usually given equal probabSlity. A bond strength
distribution could be introduced 11391 or weak spots in the geometrical
structure located [ 1401.
All the models mentioned neglect specific changes in shape, other than
192

those caused by simple changes of size. Bearing in mind the difficulties


in determining floe shape, even without shear, the simplification seems
reasonable. Unfortuxtely, experimental evidence that does not support this
assumption has become available. Mason et al. [118] noticed a rearrange-
ment of chain-like aocculates to a more spherical shape. The chain-like
geometry, iavoured on an interaction-energy basis, is forced into a form that
is more favourable from a hydrodynamic point of view. In addition Mewis et
al. [ 1411 have shown, using rheological and dielectric measurements, that
different kinds of structure are encountered when a dispersion is subject to
snear and when it is subsequently recovering at rest; a phenomenon they
called structural hysteresis.
Whereas the normal effect of shear is to break down structure, it can also
reduce the interparticle repulsion, which would increase particle interaction.
Thus, this decrease in repulsion could lead to shear thickening and antithixo-
tropy. The physical reality of such a mechanism has been supported by di-
electric [ 19 ] and by rheological measurements on systems with variable
colloidal stability [ 17,105].

(iii) Elasticity in concentrated dispersions

An important and as yet unanswered question is the seat of elasticity in


concentrated dispersions that, at least under small deformations, behave as
solids. From a thermodynamic point of view, potential and entropic mecha-
nisms can be distinguished. Neither can be eliminated a priori.
In highly concentrated dispersions of spherical particles the deformation
must be located mainly in the interparticle space. Hence, a potential mecha-
nism as defined from the DLVO theory will describe the elastic response.
Corresponding model calculations are available [ 32,45,142-1441.
Some clay minerals are in the form of flat plates. It has been shown that
the edges and the flat sides of these particles carry opposite charges, with
which cardhouse structures for the flocculated system have been suggested
[ 1453. Using these materials as a model, Rehbinder [ 146 ] and Uchida [ 1473
put forward entropic mechanisms for elasticity, based on orientation of the
plates. Both mechanisms give reasonable values for the modulus. It is difficult
to modify the controlling parameters separately and thus obtain conclusive
evidence.
In principle, L change in temperature could provide a solution. Available
data turn out to be inconclusive [ 33,144,148]. The minor effect of
temperature on the modulus for some aqueous systems could be interpreted
as supporting the potential explanation. A WLF reduction for temperature
[33,149] has been applied to some materials, suggesting similarity with poly-
mer systems [381_ However, since polymers were present, it is impossible
to come to any definite conclusions. Indeed, the fact that the shift factors
are independent af concentration 11491 seems to indicate that elasticity
193

originated in the polymer. Deviations have been reported [150], illustrating


the complexity of the situation.
The case in which a potential elasticity is plausible is important because it
allows for a direct link between rheology and colloidal stability. The basic
model, due to van den Tempel [142], starts from a network structure - a
linear assembly of chains of particles placed close together - which are branched
and interlinked to form a three-dimensional mesh, its voids fiIled with liquid.
Such a structure satisfies two requirements: the particles are close together as a
result of attractive forces, yet the configuration they form extends throughout
the whole of the available space. Using this model of structure, and selecting
a convenient expression for the potential-energy field, van den Tempel calcu-
lated the modulus from the second derivative of the potential energy to par-
ticle distance. For particles having a shape between a sphere and a flat plate,
his relation is
G = ~AcD~‘~/~~xc~-~ (13)
where x0 is the equilibrium particle distance. With spherically-shaped particles,
the prediction is that the modulus is not particularly sensitive to their size
{ 32,143 1. Sherman discussed experimental evidence that proves a clear particle-
size effect, the quantitative expression of which cannot accurately be deter-
mined. By introducing a more complex structure where the number of bonds
changes inversely, Sherman got a similar dependence in the modulus equation.
The interparticle distance enters eqn. (13) through the expression for the
potential energy; its power is about 3 in all models. Similarly, the modulus
is predicted to be proportional to the Hamaker constant and to the concen-
tration. The latter result is not substantiated by the experiments which suggest
a stronger change with c, closer to a logarithmic relation [ 32,42,144,149,
1511. Sherman argued that there are persistent effects of particle size that
obscure the picture 1451. This statement seems difficult to reconcile with the
available evidence. Nevertheless, his model provides the framework for the
inclusion of nonlinear concentration effects. Indeed, if the number of bonds
depends on the particle size in a non-ideal network, it must also depend on
concentration.
As the potential energy is not a second-order function of position, the elas-
tic behaviour is essentially nonlinear. A constant modulus is only to be ex-
pected in the limiting case of infinitely small deformations [ 321. This conclu-
sion agrees with the experimental data [f&32,33]. No effort has been made to
explain the effect quantitatively, the reason being that rupture occurs readily
and changes the pattern. The previous models predict the right order of
magnitude of the modulus. They have been used to calculate the Hamaker
constant, which illustrates the possible uses of rheological measurements to
study structure and colloidal stability in dispersions [144].
Up to now, only undisturbed structures hsve been considered, for which
purely elastic models could be used. In reality, the material response is always
viscoelastic. For cardhouse structures, sliding of the edges on the plates has
194

been suggested as a source of viscosity, leading to a Max-Nell model (146,147].


A change in distance between two particles would necessitate flow in the
pxpressions for flow of this type are available and can be
intermediate liquid. d
used in viscoelastic modelling [ 1431.

(iv) Flow phenomena

A different situation develops if, during the measurement, bonds rupture


and subsequently reform. This reversible change in structure leads to a differ-
ent mechanism for viscoelasticity, first suggested by van den Tempel 11421
and !ater applied by l’akano [ 152 J to calculate a single relaxation time. Harris
has given a phenomenological treatment of viscoelastic behaviour due to
s’,ructural changes j153]. If rupture of bonds is possible, flow can occur. Takano
llas also applied his model to steady-state shear flow [ 152 1. I-Ie obtained a
Bingham equation in which pla&ic viscosity and yield value change with the
square of concentration. Michaels and Bolger used the interaction energy
during collision to calculate the shear viscosity of low concentration kaolinite
dispersions [ 154 1. Their results are rather similar to Takano’s. They could ex-
plain the effect of some physical parameters on flow. An interesting elabora-
tion of this theory which incorporates surface parameters has been published
by Hunter et al. f*‘7,103]. For the apparent yield value they find
_ 12ca ti
70 -@l - erp(-Kcxo ) ] c2 (141
lr202 1 12x0 I
where < is the zeta potential. In general, r. changes with concentration in
much the same way as the shear modulus [ 31,42,155]. At lower concentra-
tions, a region can be selected where the quadratic relation is valid [97,103,
151]. The predicted relation with zeta potential has been verified experi-
mentally. However, the energy required to separate particles after collision
for stable, slowly aggregating sysiems, as calculated from the yield stress,
was higher than calculated from stability theory. Hunter and Friend con-
cluded their data that the repulsive energy is determined by constantsharge
rather than by constant potential conditiom and that the primary minimum
conditions, instead of the secondary minimum, are relevant in high shear cal-
culations.
These flaw models are valid only at relatively low concentrations and high
shear rates. At higher concentrations, plastic viscosity and yield value change
exponentially with concentration [ 42,156]. At lower shear rates the struc-
ture is more complex. Thixotropic theories for this region ha1 e been de-
veloped, but their translation in rheological terms has been phenomenolog-ical
instead of following a structural argument. Van den Tempel presented an
analysis that uses the interparticle forces to calculate the floe size, the in-
creased resistance to flow being uniquely attributed to immobilization of fluid
in the floes [135]. This factor undoubtedly increases the resistance to flow;
whether it overshadows other partic!e interactions has not been proved.
195

(v) i(elation with other physical properties

Besides mechanical routes, a variety of physical techniques are avaiiable


that may be used in translating the previous section into a global picture of
the essentially variable structure of concentrated dispersions; conversely, a
search should be made for using these methods in attacking yet unsolved
problems in colloid science through subtle concerts of techniques. Thus
magnetic [ 1571 and electric fields [ 158,159] can be applied to counteract
the mechanical field. Magnetic [160], electric [115) and optical [121]
measurements can be used to supplement rheological data in detecting struc-
ture. Combination of varicus fields and detection methods gives rise to tech-
niques known as electro-optical, rheo-optic‘al, rheo-electric and the like. In
dilute dispersions, they have been proved useful in analyzing the shape and
orientation of anisometric particles. In concentrated dispersions containing
flocculates and aggregates, technical problems as well as interpretation dif-
ficulties often arise. However, light scattering [161], opacity [162], dc-
conductivity [115,1191 and dielectric [116,157] properties have been shown
to depend on the state of flocculation. u ‘
With conventional equipment and techniques, the use of light scattering is
limited to very low concentrations. The availability of high intensity laser
sources removes the limitation somewhat [161]. From its opacity, the effec-
tive op,t!cal cross section for the dispersion can be calculated, snd if this changes
with flocculation, opacity will be a measure of the variable structure [ 1621.
More detailed information is available through dielectric me‘asurements [ 1631.
As in viscoelasticity, the system is characterized by a spectrum that changes
with structure. No full analysis nor complete explanation is available. The
increase of the low-frequency dielectric constant can be explained partly by
the formation of floes with high-aspect ratios [120,164] because of their effect
on the Maxwell-Wagner dispersion. In addition, surface phenomena take part
in the process 11651. Even in nonaqueous systems, surface conductivity seems
to be an important factor [X66]. DC-conductivity is a direct measure of a
continuous structure of conductive elements in the nonconductive medium;
it provides evidence of the presence of a structure that extends throughout
the :nateriaI. The combination of dielectric and rheological measurements is
a promising tool in the analysis of variable structure; both provide time and
shear-rate dependent spectra [ 39,163 ] but are differently related to structure.
In addition, they provide information about the time scales of the different
relaxation processes. Their dependence on different variables could lead to
substantial progress in colloid science and dispersion rheology.

E.SUMMARY

Concentrated dispersions show complex rheological behaviour; their


response includes viscous, plastic and elastic phenomena. In this respect, a
qualitative similarity with polymer systems becomes evident. Important dif-
196

ferences exist in the level of the plateau region for the dynamic moduli, in
the lineairity limits for deformation, and in the time scale of nonlinear time-
dependent effects.
Whereas the rheological theory for dilute suspensions is fairly well developed,
the theory for concentrated dispersions is still at an early stage. The under-
lying principles are known, but their application in calculating the detailed
structure of dispersions has not yet been developed. Partial results are avail-
able for the flow of rather stable systems and for elastic deformation in
flocculated samples. On a qualitative basis, the effect of several physical and
physicochemical parameters can now be understood.
Theological and other physical techniques have proved useful in analyzing
the structure and the colloidal stability of dispersions. The value of a simul-
taneous combination of more than one method is indicated by available data.
Several investigations of this nature are planned and might provide some in-
formation necessary for further theoretical studies.

REFERENCES

1 D.H. Everett, Pure Applied Chem.. 31 (1972) 579.


2 H. Brenner. in W. Schowalter (Ed.), Progress in Heat and Mass Transfer, Vol. 5, Per-
gamon, Oxford, 1972, p. 89.
3 C.K. Batchelor, Annu. Rev. Fluid Mech., 6 (1974) 227.
4 B.W. Barry. Adv. Colloid Interface Sci.. 5 (1975) 37.
5 S.G. Mason, Rheol. Acta, 13 (1974) 648.
6 J.W. Goodwin, in D.H. Everett (Senior Reporter), Colloid Science, Vol. 2, The Chemical
Society, London, 1975. p. 246.
7 (a) R.H. Ottewill, in D.H. Everett (Senior Reporter), Colloid Science, Vol. 1, The
Chemical Society, London, 1973, p. 173.
(b) H. Sonntag and K. Strenge, Koagulation und StabilitZt Disperser Systeme, Veb.
Deutscher Verlag Der Wissenschaften, Berlin, 1970.
(c) J. Lyklema, Adv. Colloid Interface Sci., 2 (1968) 65.
8 V.A. Fedotova, Kh. Zhodzhaeva and P.A. Rehbmder, Dokl. Akad. Nauk, SSSR, 177
(1967) 165.
9 S.S. Davis,E. Shatton and B. Warburton, J. Pharm. Pharmacol., 20 Suppl. (1968) 157 S.
10 D.G. Thomas, J. Colloid 3ci., 20 (1965) tj7.
11 J. Mewis and R. de Bleyser, Rheol. Acta, 14 (1975) 721.
12 S.A. Dunn, Bull. Amer. Ceramic Sot., 47 (1968) 554.
13 S. Middleman, The Flow of High Polymers, Interscience, New York, 1968.
14 J.J. Hermans (Ed.), Flow Properties of Disperse Systems, North-Holland, Amsterdam,
1953.
15 H.A. Scheraga, J. Chem. Phys., 23 (1955) 1526.
16 A 8. Metzner and W. Whitlock, Trans. Sot. Rheol., 2 (1958) 239.
17 R.J. Morgan, Trans. Sot. Rheol., 12 (1968) 511.
18 R.L. Hoffman, Trans. Sot. Rheol.. 16 (1972) 155.
19 A.A. Trapeznikov, G.G. Petrzhik and T.I. Korotina, Dokl. Akad. Nauk. SSSR, 176
(1967) 378.
20 K. Umeya, in Proc. Int. Congr. Rheol., 5th. Vol. 2, S. Onogi, (Ed.), University of Tokyo
Press, Tokyo, 1970, p. 295.
21 W. Schempp, W. Friesen and J. Schurz, Das Papier, 28 (1974) VI.
22 D.J. Highgate and R.W_ Whorlow. Rheol. Acta. 9 (1970) 569.
197

23 L. f)intenfass, in S. Onogi (Ed.), Proc. Int. Congr. Rhea!.. 5th. Vol. 2. University of
To’ryo Press, Tokyo, 1970, p. 281.
24 W,Ii. Bauer and B.A. Collins in F.R. Eirich (Ed.), Rheology, Theory and Applications.
Vol. 4, Academic, New York, 1967, p. 423.
25 M. Reiner and G.W. Scott-Blair in F.R. Eirich (Ed.), Rheology, Theory and Applications,
Vol. 4. Academic, New York, 1967, p. 461.
26 H. Freundlich. Thixotropy, Hermann. Paris, 1935.
27 M-N. Kruglitshii and N.V. Mlhailov, Rheology of Thixotropic Systems, Nauk. Dum-
ka.. Kiev, USSR, 1972.
28 D-C.-H. Cheng. Nature (London). 245 (1973) 93.
29 A.H.P. Skelland, Non-Newtonian Flow and Heat Transfer, Wdey. New York. 1967.
p. 5.
30 K Umeya. T. Isoda. T. Ishii and K. Sawamura. Powder Techno!.. 3 (1969-70) 59.
31 R. Gotoh and K. Shimizu. J. Sot. Materials Sci. Japan, 15 (1966) 283.
32 C.J. Nederveen, J. Colloid Sci., 18 (1963) 276.
33 R-D. Hoffman and R-R. Myers, in E-H. Lee (Ed.), Proc. Int. Congr. Rheol., 4 th. Vol. 2.
Interscience, New York, 1965, p_ 693.
34 J.M.P. Papenhuijzen, Rheol. Acta, 10 (I 971) 493.
35 N.N. Serb-Serbina and P.k Rehbinder, Kolloidn. Zh., 9 (1947) 381.
36 K. Strenge and H. Sonntag. Coll. PO!. Sci.. 252 (1974) 133.
37 B. Gross, Mathematical Structure of the Theories of Viscoelasticity, Hermann, Paris,
1953.
38 J.D. Ferry, Viscoelastic Properties of Polymers, 2nd ed.. Wiley, New York, 1970.
39 G. Schoukens, A.J.B. Spaull and J. Mewls, to be presented at the 7th Int. Congress on
Rheology, Gothenburg, 1976.
40 S. Onogi, T. Masuda and T. Matsumoto, Trans. Sot. Rheol.. l-1 (1970) 275.
41 A.F. Douglas, G.A. Lewis and A.J.B. Spaull, Rheol. Acta, 10 (1971) 382.
42 S. Onogi, T. Matsumoto and Y. Warashina, Trans. Sot. Rheol., 17 (1973) 175.
43 M. Takano, Bull. Chem. Sot. Jpn., 37 (1964) 78.
44 T. Matsumoto, Y. Segawa, Y. Warashma and S. Onogi, Trans. Sot. Rhea!., 17 (1973) 47.
45 P. Sherman, in S. Onogi, (Ed.), Proc. Int. Congr. Rheol., 5th, Vol. 2, University of
Tokyo Press, 1970, p. 327.
46 H. Kambe and M. Takano, in E.H. Lee (Ed.), Proc. Int. Congr. Rheol., 4th. Vol. 3.
Interscience, New York, 1965. p. 557.
47 L. Hellinckx and J. Mewis. Rheol Acta, 8 (1969) 519.
48 K.M. Beazley, Tech. Assoc. Pulp Pap. Ind., 50 (1967) 151.
49 A.R. Payne and R.E. Whittaker. Rheol. Acta. 9 (1970) 91.
50 N.F. Astbury and F. Moore, Rheol. Acta. 9 (1970) 124.
51 N.S. Parker and G.E. Z-?ibberd, Rheol. Acta, 13 (1974) 910.
52 D.W. de Bruijne. N.J. Prirzhard and J.M.P. Papenhuijzen, Rheol. Acta, 13 (1974) 418.
53 T. Matsumoto, C. Hitormi and S. Onogi. J. Sot. Rheol. Japan, 2 (1974) 12.
54 R. Rutgers, Rheol. Acta. 2 (1962) 305.
55 P. Sherman, Industrial Rheology, Academic, London, 1970.
56 (a) A. Einstein. Ann. Phys. (Leipzig), 19 (1906) 289.
(b) A. Einstein, Investigations on the Theory of the Brownian Movement, Dover
Publications, 1956.
57 G.K. Batchelor and J.T. Green, J. Fluid Mech.. 56 (1972) 401.
58 I. Yaron and B. Gal-Or, Rheol. Acta, 11 (1972) 241.
59 R. Herzynski and I. Pienkowska, Arch. Mech. Eng. (Warsaw), 27 (1975) 201.
60 N.A. Frankel and A. Acrivos. Chem. Eng. Sci., 22 (1967) 847.
61 T. Gillespie, J. Colloid. Sci., 18 (1963) 32.
62 R.F. Fedors, J. Colloid Interface Sci., 46 (1974) 545.
63 E.J. Hinch. Ph.D. Thesis, Cambridge, 1912, p. 1. 21.
64 C.-J. Lin, J.H. Peery and W.R. Schowaiter, J. Fluid Mech., 44 (1970) 1.
198

65 T. Gillespie, J. Colloid Sci., 18 (1963) 562.


66 M. Mooney, J. Colloid Sci., 6 (1951) 162.
67 G.F. Eveson, in C.C. Ml11 (Ed.). Rheology of Disperse Systems, Pergamon, London,
1959, p. 61.
68 R.F. Far&, Trans. Sot. Rheol., 12 (1968) 281.
69 F. Skvara and M. Vancurova. Silikaty (Prague), (1973) 9.
70 G-B. Jeffrey, koc. Roy. Sot. London, Ser. A. 102 (1922) 161.
71 E. Anczurowski and S.G. Mason, J. Colloid Interface Sci., 23 (1967) 522.
72 E. Anczurowski and S-G. Mason, J. Colloid Interface Sci.. 23 (1967) 533.
73 E. Anczurowski, R.G. Cor and S.G. Mason, J. Colloid Interface Sci.. 23 (1967) 547.
7.1 R.G. Cox and H. Brenner, Chem. Eng. Sci.. 26 ( 1971) 65.
75 A. Okagawa, R.G. Cox and S.G. Mason, J. Colloid Interface Sci., 45 (1973) 303.
76 E-Y. Harper and I-Dee Chang, J. Fluid Mech., 33 (1968) 209.
77 F. Gauthier. H.L. Goldsmith and S.G. M‘ason, Kolloid-Z. 2. Polym., 248 (1971) 1000.
78 G.K. Batchelor. J. Fluid Mech.. 46 (1971) 813.
79 J. Mewis and A.B. Metzner. J. Fluid Mech., 62 (1974) 593.
80 (a) H. Brenner, J. Colloid Sci., 23 (1967) 407.
I(b) H. Brenner. Int. J. Multiphase Flow, 1 (1974) 195.
81 J.M. Rallison and E.J. Hinch, in press.
83 J.G. Kirkwood, in P.L. Auer (Ed.), Macromolecules, Gordon and Breach, New York,
1967.
83 (a) E.J. Hinch and L.C. Lea], J. Fluid Mech., 57 (1973) 753.
(b) L.G. Lea1 and E.J. Hinch, Rheol. Acta, 12 (1973) 127.
84 1.M Krieger, Trans. Sot. Rheol.. 7 (1963) 101
85 (a) H. Giesekus. Rheol. Acta. 2 (1962) 50.
(b) H. Giesekus, in M. Reiner and D. Abir (Eds.), Proc. Int. Symp. 2nd Order Effects
in Elasticity, Plasticity and Fluid Dynamrcs, Pergamon, London, 1961, p. 553
86 G.R. Feat and S. Levine, J. Chom. Sot., Faraday Trans. 2.71 (1975) 102.
87 H.C. Hamaker, Physica (Utrecht), 4 (1937) 1058.
88 B. Vincent, Adv. Colloiri Interface Sci., 4 (1974) 193.
89 P. Bagchi and R.D. Void, J. Colloid Interface Sci.. 33 (1970) 405.
96 M. Gosiewski, presented at the European Mechanics Colloquium, Number 49,
Jablonna, Poland, 1974.
.91 H.C. Longuet-Higgins and B. Widom, Mol. Phys., 8 &9f%) 549.
92 B.V. Derjaguin and L. Landau, Acta Physicochim. USSR, 14 (1941) 633.
93 E.J.W. Verwey and J.Th.G. Overbeek, Theory of the Stability of Lyophobic Colloids.
Eisevier, Amsterdam, 1948.
94 D.H. Napper and R.J. Hunter, in M. Kerker (Ed.), Hydrosols, MTP International Rev. of
Science, Physical Chemistry, Ser. 1, Vol. 7, Butterworths, London, 1972, p. 241.
95 G.M. Bell and G.C. Peterson, J. Colloid Interface Sci.. 41 (1972) 542.
96 J.B. Melville and A.L. Smith, J. Chem. Sot., Faraday Trans. 1, 70 (1974) 1551.
97 J.P. Friend and R.J. Hunter, J. Colloid Interface Sci., 37 (1971) 548.
98 A. Doroszkowski and R. Lambourne, J. Colloid Interface Sci.. 27 (1968) 214.
99 M. Schi%r. Rheol. Acta, 9 (1970) 1.
100 J.W. Goodwin, Trans. Brit. Ceram. Sot., 70 (1971) 65.
101 J.G. Brodnyan and E.L. Kelley, J. Colloid Sci.. 20 (1965) 7.
102 A. Packter. Rheol. Acta, 2 (1962) 44.
103 R.J. Hunter and S.K. Nicol, J. Colloid Interface Sci., 28 (1968) 250.
104 M. Takano and H. Kambe, Bull. Chem. Sot. Jpn., 36 (1963) 1424.
105 R.L. Hoffman, J. Colloid Interface Sci., 46 (1974) 491.
106 W. Carr, J. Oil Colour Chem. Assoc., 45 (1962) 28.
107 S.K. Nicol and R.J. Hunter, Aust. J. Chem., 23 (1970) 2177.
108 K. Rehacek and H. Schiitte, Plasto Kautsch., 16 (1969) 773.
109 L.J. Warren, J Colloid lnterface Sci., 50 (1975) 307.
199

110 J.A. Long, D.W J. Osmond and B. Vincent, J. Colloid Interface Sci., 42 (1973) 545.
111 S H. Bell and V.T. Crow& m G.D. Parfitt (Ed.), Dispersron of Powders in Liquids, 2nd
Edition, Applied Sci., London, 1973, p. 267.
112 T.C. Patton, Paint Flow and Pigment Disperion, Interscience, New York, 1964, p-
201.
113 H.D. Jeffries, J. Oil Colour Chem. Assoc., 45 (1962) 681.
114 I.R. Sheppard and G. Cope, Rheol. Acta, 4 (1965) 344.
115 CM. McDowell and F L. Usher, Proc. Roy. Sot. London, Ser. A, 131 (1931) 409.
116 A. Bondi and C.J. Penther, J. Phys. Chem.. 57 (1953) 72.
117 S. Okamoto and S. Hachisu, J. Collord Interface Sci., 43 (1973) 30.
118 E-B. Vadas. H.L. Goldsmith and S.G_ hlason. J. Colloid Interface Sei.. 43 (1973) 630.
119 S. Peter, Kolloid-2. 2 Polym., 113 (1949) 29.
120 A. Voet, J. Phys. Chem.. 51 (1947) 1037.
121 J. Schweitzer and B R. Jennings, J. Colloid Interface Sci., 37 (1971) 443.
122 M. von Smoluch-:gski, Z. Physik. Chem., 92 (1917) 129.
123 J.Th.G. Overbeek, in H.R. Kruyt (Ed.), Colloid Scrence, Vol. 1, Elsevier, Amsterdam,
1952. p. 278.
12-I (a) D-N. Sutherland, J. Colloid Interface Sci.. 25 (1967) 373.
(b) J.C. Ravey, J. Colloid Interface Sci., 50 (1975) 545.
125 D.N. Sutherland, Nature (London), 226 (1970) 1241.
126 1. Goodarz-Nia and D.N. Sutherland, Chem. Eng. Sci., 30 (1975) 407.
127 L.A. Utrackr, J. Colloid Interface Sci , 42 (1973) 185.
128 M.R. Kamal and I. Patterson, Candinn J. Chem. Eng.. 52 (1974) 707.
129 S.V. Kao ant1 S.G. Mason. Nature (London), 253 (2975) 619.
130 K.M. Beazley, Tr.ms Brit. Ceram. Sot., 63 (196-L) -151.
131 CF. Goodeve, ‘IkdnS. Faraday Sot., 35 (1939) 342.
132 3. Peter, Kollo~d-Z. Z. Polym., 11-l (1949) -1-I.
133 T. Gillespie, J. Colloid %I., 15 ( 1960) 219.
134 N. C&son, in C.C. Ml11 (Ed.), Rheology of Disperae Systems, Pergamon, London,
1959, p_ 84.
135 M. van den Tempel. in P. Sherman (Ed.), Rheology of Emulsions, Pergamon, London,
1963. p. 1.
136 E_ Ruckenstein and J. Mewis. J. Colloid Interface Sci.. 4-t (1973) 532.
137 H.D. Weymann. in E.H. Lee (Ed.), Proc. Int. Congr. Rheol.. 4th. Vol. 3. Interscience.
New York, 1965, p_ 573.
138 H.A. Mercer and H.D. Weymann, Trans. Sot. Rheol., 18 (1974) 199.
139 M.P. Volarovich, M.N. Avdeev and A.A. Medvedeva, Kolloidn. Zh., 35 (1973) 148.
140 D. Tomi and D.F. Bagster. Chem. Eng. Sci., 30 (1975) 269.
141 J. Mewis, A.J.B. Spaull and J. Helsen, Nature (London), 253 (1975) 618.
142 M. van den Tempel, J. Colloid Sci., 16 (1961) 284.
143 J.M.P. Papenhuijzen, Rheol. Acta. 11 (1972) 73.
144 K. Strenge and H. Sonntag, Colloid Polym. Sci.. 252 (1971) 133.
145 H. van Olphen, Discuss. Faraday Sot.. 11 (1951) 82.
146 SD. Shchukin and P.A. Rehbinder, Kolloidn. Zh., 33 (1971) 450.
147 I. Uchida and H. Fujimoto. J. Sot. Mater. Sci. Kyoto, 12 (1963) 276.
148 R. Roscoe, in E.H. Lee (Ed.), Proc. Int. Congr. Rheol., 4th. Vol. 3, Interscience, New
York. 1965, p. 593.
149 S. Onogi, T. Masuda and T. Matsumoto, Nippon Kagaku Zasshi. 89 (1968) 464.
150 T. Matsumoto. T. Masuda. K. Tsutsui and S. Onogi. Nippon Kagaku Zasshi, 90 (1969)
360.
151 H. van Olphen, Proc. 6th National Conf. Clays Clay Minerals, U.S. National Res. Council,
Washington DC., 1957. p. 196.
152 M. Takano, Bull. Chem. Sot. Jpn., 36 (1963) 1418.
153 J_ Harris. Rheol. Acta. 6 (1967) 6.
200

164 AS. Michaels and J.C. Bolger, Ind. Eng. Chem., Fundam., 1 (1962) 153.
155 T. Matsumoto, Y. Segawa, Y. Warashina and S. Onogi, Trans. Sot. Rheol., 17 (1973)
47.
156 R.N. Weltmann. in F.R. Eirich (Ed.), Rheology, Theory and Applications, Vol. 3. Aca-
demic, New York, 1960, p. 189.
157 A. Voet and L.R. Suriani, J. Cblloid Sci., 6 (1951) 155.
158 S.T. Demetriades, J. Chem. Phys., 29 (1958) 1054.
159 A. Qkagawa, R.G. Cox and S.G. Mason, J. Colloid Interface Sci.. 47 (1974) 536.
160 Y. Yashuhara and K. Tanaka, J. Jap. Sot. Testing Materials, 11 (1962) 277.
162 R.A. Isaksen, C.R. Williams, J.F. Heaps and R.J. Clark, Ind. Eng. Chem.. Prod. Res.
Develop.. 10 (1971) 298.
162 R.A. Ross, H.D. Weymann and M.C. Chuang. Phys. Fluids, 16 (1973) 784.
163 S.S. Dukhin and V.N. Shilov, Dielectric Phenomena and the Double Layer in Disperse
Systems and Poly-electrolytes, Izdatel’stova Nauk. Dumka., Kiev, USSR, 1972: Engl.
Transl. Keter. Jerusalem. 1974.
164 H. Fricke, J. Phys. Chem.. 57 (1953) 934.
165 C.T.O’Konski. J. Phys. Chem., 64 (1960) 605.
166 k Voet. Amer. Inkmaker. 35 (1957) 34.

Você também pode gostar