Sie sind auf Seite 1von 11

TECHNICAL REVIEW

Colloid Stability
Donald H. Napper
Department o f Physical Chemistry, Uniuersity o f Sydney,
Sydney, N.S.W., 2006, Australia
3 principles that govern colloid stability are of manifold importance to many different
technologies. The origin of the London disper!;ion forces, which cause naked, uncharged
idol particles to flocculate, and how this fl occulation may be prevented are discussed.
; are a t least two general mechaniswIS whereby colloid stability is imparted:
.ostotic stabilization and steric stobilizotic~ n .The Deryagin-Landau-Verwey-Overbeek
. ..
y provides a quantitative explanation at manv Dhenomena invowin4- electrostatic
lization. No comparable theory for steric stabilization exists, although some general
,
iples which appear to govern this type of stabilization have been delineated.
These give considerable insight into the vener able phenomenon of the protective action
of certain adsorbed polymers. An understancling of the general principles that govern
colloid stability also permits conclusions as to the best procedures by which stable
dispersions may be induced to flocculate.

This Technical Review updates, by analytical assessment,


current views on one of the most stubborn and necessary
processing techniques encountered by the industrial chemist.
It is t h e .d i c v. of this iournal to invite manuscriots which
treat concepts CIS well CIS products.
The Editor

D. H. NAPPERis currently
a Queen Elizabeth I1 Fellow
at the Uniuersity of Sydney.
He graduated B.Sc., with
First Class Honours and the
Uniuersity Medal, from that
uniuersitv in 1959, and receiued the M.S. in 1961. Subsequently, he was awarded
a Ph.D. at the Uniuersity of Cambridge. He has since
held positions in both industry and the academy. He
has published papers on colloid stability, thermodynamics, polymerization kinetics, light scattering, dissolution kinetics, tooth decay, and protein chemistry.

r
LOLLOID

....

" .

stability impinges on many areas ot technological significance. Some that come to mind readily are
emulsion and suspension polymerizations, detergent action
in washing, the formulation of paints and pharmaceutical
emulsions and dispersions, and the clarification of water
and water residues. Beyond the industrial realm, the
principles that govern colloid stability are also relevant
to many biological phenomena-e.g.,
the properties of
blood, antigen-antibody reactions, and even the nature
of that humble commodity, milk-and
agricultural
practices-e.g., the degree of dispersion of the soil affects
its crop hearing potential.
The questions posed in considering colloid stability may
be discussed under three heads. First, what causes unstable
colloidal particles to flocculate? Second, how can the
occurrence of flocculation he prevented? Third, given stable dispersions, how may these he flocculated? I n what
follows, some of the more recent concepts developed to
answer these questions are presented. This account is
intended to supplement other pertinent reviews (Lyklema,
1967a,b, 1968; Ottewill, 1967; Verwey and Overheek,
1948).
London Attraction

It is a matter of common observation that dispersions


of uncharged, naked colloidal particles flocculate rapidly:
The number of particles is reduced by a factor of 2 in
Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970 467

a matter of a few seconds, a t normal particle number


concentrations. This behavior is a consequence of the
long-range attractive forces operative between the particles.
Gravitational forces are not responsible for this longrange attraction. Calculation of the gravitational potential
energy between two colloidal particles shows that it is
insignificant compared with thermal energy. This is not
too surprising, since gravitational forces are the weakest
of the four classes of interactions between elements of
matter discovered to date-strong, electromagnetic, weak,
and gravitational. T o explain rapid flocculation, electromagnetic forces, which are the second strongest in the
hierarchy, must be invoked. I n particular, this action
a t a distance is attributed to London dispersion forces,
which are primarily of an electrical origin (Kallmann and
Willstatter, 1932).
Dispersion forces are not so called because they are
operative in colloidal dispersions, but rather, because the
electron fluctuations that give rise to optical refractive
dispersion also give rise to these forces. The existence
of dispersion interactions between elements of matter is
most convincingly demonstrated by considering the
liquefaction on cooling of the noble gases-e.g., argon
and helium. Atoms of these gases possess no permanent
dipole moments to give dipole-dipole (Keesom) or dipoleinduced dipole (Debye) interactions. Yet liquids are characterized by strong intermolecular or interatomic interactions. Attractive forces must therefore be operative
between the atoms of a noble gas.
The origin of dispersion forces is firmly rooted in
quantum mechanical considerations. A simplified, nonquantized view of their origin is presented in Figure 1.
An electron in its orbital movement about the nucleus
of a rare gas atom is instantaneously stopped in the zone
marked Z (Figure 1,a);this zone is diffuse as a consequence
of the uncertainty principle. Suppose, now, that a second
atom is placed in the immediate neighborhood of the
first. I t is easy to see that the simplest calculations must
imply that the position of an electron in the second atom
is determined solely by the position of its associated
nucleus. More refined second (and higher) order calculations, however, show that the position of an electron in
the second atom is correlated with the position of the
electron in the first atom. Specifically, the charge
configuration illustrated in Figure l , a , is adopted. The

.//
+ -

Figure 1. Schematic representation of origin of dispersion


forces

468

Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970

presence of instantaneous dipoles in the two atoms is


such that attraction ensues. When the electron a t Z has
moved to its new position at Z , the second electron
has also moved, so that the polarity of the two instantaneous dipoles is reversed. Attraction still occurs. When the
instantaneous dipoles are time-averaged over all positions,
no permanent dipole moment results. But the time-average
attractive force is not zero, since counterbalancing repulsive forces are negligible, a t least a t large separations.
From the foregoing discussion, the approximate
additivity of dispersion forces should not appear too
improbable. This additivity is important because the dispersion forces operative between noble gas atoms are relatively short-range, extending over only a few Angstroms.
Because each atom or molecule in one colloid particle
attracts every atom or molecule in another, even small
colloidal particles containing, say, lo3 atoms would have
their attractive interactions magnified by a factor of ca.
(lo3 x lOj2) = 5 x lo, as a result of additivity. Not
surprisingly, the London attraction between two colloidal
particles is a long-range effect, operative over tens of
Angstroms.
The attractive potential energy ( V4) between two
spheres, each of radius a and with a distance of closest
approach of H,, is calculated from the microscopic dipole
approximation outlined above to be (Hamaker, 1937)

Vq= - ( A / 1 2 ) { ( x 2 + 2 x ) - + (x+ 2 x + 1)

where A = effective Hamaker constant and x = (H,/


2a). For close approach of the spheres, x << 1 and so
we obtain

V , = -A/24x = -Am/lSH,

(2)

Other geometrical situations may be similarly considerede.g., flat plates of infinite thickness-but colloidal dispersions often approximate closely to sphere-sphere interactions.
The effective Hamaker constant, A , can be calculated
from the vacuum Hamaker constant. The latter is a
measure of the intrinsic strength of the dispersion attraction between unit volume elements of the material a t
unit separational distance. The vacuum Hamaker constant
is obviously calculable from optical refractive dispersion
data (Ottewill and Wilkins, 1962). However, measurements of surface tension (Fowkes, 19641, diamagnetic susceptibility (Watillon and Joseph-Petit, 1966), and ionization potentials (Ottewill and Shaw, 1966) also permit the
Hamaker constant in vacuo to be calculated.
T o calculate A from the vacuum value, corrections
must be applied to allow for the presence of the dispersion
medium (Hamaker, 1937); the presence of any adsorbed
material (Vold, 1961); the retardation effect (Casimir and
Polder, 1948); and the dielectric constant of the dispersion
medium (Schenkel and Kitchener, 1960).
The dispersion medium correction is really an elementary consequence of Archimedes principle (Napper, 1967);
it accounts for the fact that two gas bubbles in a liquid
attract one another with the same force that exists between
two liquid drops of the same size and a t the same distance
of separation in a gas (Lyklema, 1967b).
The retardation correction becomes increasingly important as the particles become increasingly separated.
A finite time is required for the propagation of an electro-

magnetic wave from an electron in position 2 (Figure


1,a) to the electron in the second atom and back again.
If this time interval is sufficiently long, the position of
the electron originally a t Z will have altered appreciably
on return of the electromagnetic wave. An out-of-phase
component then arises, which attenuates the dispersion
attraction. This occurs when the distance between the
colloidal particles is comparable to the London characteristic dispersion wavelength of the material (London,
1930).
Values of' the effective Hamaker constant calculated
to 10-l" erg.
in this way usually lie in the range 10
Thus the force of at traction between two colloidal particles
is characteristically of the order of 10~'dyne, or, roughly,
the weight exerted by 10 gram. Such small forces are
not yet readily measurable. But the London dispersion
attraction between larger, macroscopic bodies has been
measured directly (Deryagin and Abrikossova, 1954; Deryagin et ai.. 1956; Kitchener and Prosser, 1957). These
early experiments measured retarded dispersion attractions, since the minimum separational distance was usually
greater than 1000 A, which is of the order of the characteristic London wavelength. More recently, elegant experiments devised by Tabor and Winterton (1968, 1969) have
permitted the direct measurement of both the retarded
and unretarded dispersion attraction between ultrasmooth
surfaces of mica. The approach distances between the
surfaces were measured by multiple-beam interferometry;
the attractive forces were monitored by the deflection
of a cantilever beam. Separations as small as 50 A were
studied. A relatively sharp transition from normal t o
retarded forces was evident at ea. 150 A , as expected
theoretically.
The experimental results of Tabor and Winterton, as
well as those of previous investigators, relevant to the
retarded force domain are in good agreement with the
Lifshitz macroscopic theory of dispersion forces (Deryagin
et al., 1956). The Lifshitz theory is superior to the
Hamaker approach because multiple interactions between
the constituent species are taken into account. The numerical evaluation of the Lifshitz formula requires detailed
knowledge of the macroscopic dielectric constant and
dielectric loss factor over the complete frequency range
of dispersion. This demanding requirement has prevented
extensive application of the theory to colloidal dispersions,
except perhaps in the calculation of retarded forces when
only the static dielectric constant-Le., the square of the
refractive index extrapolated to long wavelengths-need
be known (Overbeek, 1966).
Repulsion

A potential energy diagram for the interaction of two


colloidal particles is presented in Figure 2. The total potential energy (V.1) curve normally required for stability is
characterized by the presence of a potential energy maximum ( V.ll). Essentially, this functions as an activation
energy. The Maxwell-Boltzmann distribution of thermal
energies is the sole source of energy which enables the
two particles t o surmount the potential energy hump and
to enter the primary potential energy minimum. If V.rr
is very much larger than thermal energy-Le.,
V N >>
hT-the frequency of transitions across the energy barrier
is very small. Dispersions of such particles exhibit longterm stability, albeit of a kinetic type. Incidentally, the
primary potential energy minimum arises from the

/
+(

\,

ELECTROSTATIC

R E P u L sioN,
STERIC

vT

SECONDARY MINIMUM

XPRIMARY MINIMUM

I
I

I VA 1

I LONDON ATTRACTION

Figure 2 . Total potential energy (VT) diagrom for two colloidal particles

occurrence of the short-range overlap (or Born) repulsive


forces. Essentially, these reflect the operation of the Pauli
exclusion principle, which ensures that the density of matter is finite.
T o achieve a potential energy maximum, it is necessary
to put in a repulsive potential energy between the particles.
This can be achieved in a t least two ways at present:
electrostatically and sterically. We consider these two
possibilities in turn.
Electrostatic Stabilization

The origin of the repulsive potential energy in the case


of electmstatic stabilization is most readily understood
by considering first a charged aerosol-e.g., a fog or a
smog. Close approach of the aerosol particles is opposed
by the Coulombic repulsion operative between the particles. If the aerosol particles were then dispersed in a
liquid continuous phase, the interparticle repulsion would
be expected t o be decreased somewhat. In fact. the
attenuation is commonly rather greater than expected,
because each particle is surrounded by an ionic atmosphere
(Figure 3 ) , as a consequence of the principle of electroneutrality .
The atmosphere of counterions and the surface charge
constitute the electrical double layer. The stability of
the dispersion is governed inter alia by the properties
of the double layer. Indeed, the repulsive potential energy
can be considered to arise from the work that must be
done to remove the counterions from between the particles
so that the particles can approach one another. We must,
however, adopt this view with some caution, because sta-

+
+
n
2
-

- + +

4.

-+

A-

+,,----:

A -

++

+ - - I

+;

* - \

/ +*

- t

Figure 3. Collision of two electrostatically stabilized particles

Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970 469

bility is evident in aerosols when counterions are absent;


moreover, flocculation may occur when a large number
of counterions are not removed from between the particles.
Deryagin and Landau (1941), and later Verwey and
Overbeek (1948), calculated the repulsive potential energy
in the case of electrostatic stabilization. When combined
with the attractive potential energy discussed previously,
these calculations enable a quantitative theory of the stability of lyophobic colloids to be formulated. This theory
is commonly termed the DLVO theory, in recognition
of its originators. I t is considered by many investigators
to be the definitive theory of electrostatic stabilization.
The DLVO theory assumes that the ions in the electrical
double layer are point charges and that specific ion adsorption a t the particle surface is absent. This latter assumption is often violated in practical systems, except perhaps
at elevated temperatures (Lyklema, 1966). But the theory
is saved somewhat by the fact that the outer, rather
than the inner, regions of the double layer ordinarily
are important in stability. I t is further assumed in the
DLVO theory that, as the particles approach, the electroBtatic potential at the surfaces is constant-i.e.,
some
discharge of potential determining ions occurs.
The surface potential of electrostatically stabilized dispersions may arise in a variety of ways-e.g., by the
adsorption of potential determining ions, such as anionic
or cationic surfactants; by the ionization of ionogenic
groups, such as carboxylic acid or sulfate half-ester groups;
and by the presence of lattice defects in the disperse
phase, such as in clays. This surface potential tends to attract
the counterions to the particle surface, a process opposed
by thermal agitation. Consequently, most of the counterions exist in a diffuse layer surrounding the particle
(Figure 3).
Because of the assumptions involved in the DLVO
theory, the distribution of ions in the diffuse double
layer corresponds to a Poisson-Boltzmann distribution
(Chapman, 1913; Gouy, 1910, 1917). This is the same
distribution function that occurs in the well-known DebyeHuckel ionic atmosphere theory of strong electrolytes.
However, the Gouy-Chapman theory for the distribution
of ions a t a charged, planar interface was formulated
long before the Debye-Huckel theory. The thickness of
the diffuse double layer ( l / ~ is
) arbitrarily taken as the
distance from the surface at which the potential falls
to l / e t h (= 0.37) of its surface value, e being the base
of the natural logarithms. At room temperature, the thickness (in Angstroms) of the double layer associated with
a planar surface may be calculated from

where = dielectric constant of the dispersion medium


(assumed constant in the double layer) and I = ionic
strength. In water, ( 1 / ~ is
) simply equal to 311.
The repulsive potential energy (V,) for two charged
spheres as calculated by the DLVO theory is rather complicated (Verwey and Overbeek, 1948) and we confine
ourselves here t o two simpler, but approximate, limiting
expressions. First, if the surface potential $ is small, as
is h a ,

V R= { ea$/ ( [ H , / a ]+ 2) }exp(-hHn)
while, if h a is large and $. < 50 mv.

vK= (,a$12)1n[1 + e x p ( - ~ H , ) ]
470

Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970

(4)

(5)

The term (ta$/2) occurs approximately in both expressions and corresponds to the effective Coulombic repulsion
between the particles in the absence of double layers.
The diffuse atmospheres of counterions screen the surface
charges from each other, while being themselves mutually
repulsive. The additional exponential and logarithmic
terms allow for this attenuation of the Coulombic repulsion.
To apply the above expressions, it is necessary to determine $. If specific adsorption of ions occurs, the value
of $ required is that which determines the diffuse double
layer potential. This is not the surface potential but rather
the Stern potential, $ ,, which occurs a t the plane where
specific adsorption effects cease and the diffuse layer
begins. Commonly G o , is equated with the electrokinetic
or {-potential, which is determined by the plane of shear
in electrokinetic movement.
The total potential energy (Vr) in the DLVO theory
is obtained by summing the attractive potential energy
( Vq) and the repulsive potential energy (Vx)-i.e.,

vr = VK + v.4
What phenomena can the DLVO theory explain? First,
it accounts for the sensitivity of electrostatically stabilized
dispersions to the ionic strength of the dispersion medium.
According to Equation 3, the thickness of the double
layer decreases with increasing ionic strength. If this is
reduced sufficiently, the particles approach one another
to the point where the dispersion attraction takes over
and flocculation is observed.
Second, it explains the sensitivity of the stability to
the valency of the counterion, embodied in the SchulzeHardy rule. This states that the comparative flocculation
effectiveness of different counterions is proportional to
the sixth power of their valencies. The sixth power dependence can be derived explicitly, using the DLVO theory
for flat plates.
Third, the DLVO theory can account, a t least qualitatively, for the rate of slow coagulation of colloidal
dispersions-Le., coagulation slower than the rapid coagulation of uncharged particles. Slow coagulation occurs
when the total interaction potential energy still assumes
positive values because the surface potential, and thus
the repulsive potential energy, is not zero. I t is possible
from measurements of slow coagulation to determine
values for the effective Hamaker constant. These usually
agree in order of magnitude with those calculated by
the various microscopic procedures outlined previously
(Lyklema, 1967a). The experimental values, however, are
often smaller than those theoretically calculated, sometimes by a factor of 3 or 4. Moreover, the dependence
of the rate of slow coagulation upon the particle radius
is not apparently adequately described by the DLVO
theory (Ottewill and Shaw, 1966), although this view has
recently been strongly challenged (Wiese and Healy, 1970).
Fourth, the coalescence of polarized mercury droplets
is in reasonable agreement with the predictions of the
DLVO theory (Usui et al., 1967; Watanabe and Gotoh,
1963).
Fifth, the DLVO theory can account for the equilibrium
thicknesses of certain thin soap films. The forces that
determine film thicknesses are essentially the same as
those involved in colloid stability. Studies of thin films
are important because the distance dependence of the
forces which determine the equilibrium thickness can be

evaluated (Lyklema, 1967h). Flocculation studies of incipiently unstable dispersions are incapable of yielding
this information directly. The chief disadvantage of soap
films resides in the fact that
cannot he measured readily
hut must be estimated empirically. As yet, very precise
agreement between theory and experiment is still lacking
in soap film studies: Values of the experimental Hamaker
constants are larger by a factor of from 3 to 5 than
those estimated theoretically (cf. the reverse situation
in flocculation studies).
Sixth, certain phenomena, such as Schiller layers, tactoid~
formation, and the flocculation of large, electrostatically
stabilized polystyrene latex particles, can he explained
in terms of secondary minima (Overheek, 1952; Schenkel
and Kitchener, 1960). The secondary minimum (Figure
2) arises because the falloff in repulsive potential energy
is more rapid than that of the attractive potential energy.
The depth of the secondary minimum is often relatively
small and so flocculation in it should he easily reversede.g., by dilution.
The DLVO theory, however, is not without its
deficiencies. The usual assumption of constant surface
potential corresponds to the maintenance of equilibrium
during Brownian collisions of particles. This seems unlikely
to occur, not only in dispersion studies hut also in soap
film investigations (Jones, 1969). The extreme disequilibrium situation might be taken as the maintenance I f
constant surface charge density-Le., no discharge of the
potential determining ions occurs on close approach.
Fortunately, both the constant surface potential and constant surface charge density assumptions give reasonably
similar results for flocculation of colloidal dispersions.
The DLVO theory in its simplest form is also unable
to cope with the specific differences observed between
the flocculation effectiveness of counterions of equal
charge. The antagonism of one ion for the flocculation
properties of another ion also goes unexplained. However,
allowance for specific ion adsorption may resolve these
anomalies.

negatively chnrgau p u ~ y x ~ y r e ~Ll e~ L L C ~ 111


S
LLXS
way
(Gregory, 1969). The higher molecular weight cationics
appear to function primarily by a bridging mechanism.
Steric Stabilization
Uncharged colloidal particles which are prevented from
flocculating by the presence of nonionic polymer molecules,
are said to he sterically stabilized (Heller and Pugh, 1954).
The generic term steric is used in this context with
a broad thermodynamic connotation, rather than with
the restricted meaning its use entails in organic
chemistry-viz.,
the operation of the Pauli exclusion
principle. The mechanism whereby polymer molecules
protect colloidal particles against flocculation is one of
the oldest unsolved problems in colloid science. The phenomenology of steric Stabilization is a t present being
actively catalogued. Some general principles which govern
this type of stabilization are slowly emerging.
The origin of the repulsive potential energy in steric
Stabilization is illustrated in Figure 4. Consider a Brownian
collision between two colloidal particles, completely surrounded by polymer chains securely attached to their
surfaces. The particle surfaces can approach one another
only if interpenetration and/or compression of the stabilizing chains occur. Qualitatively, the processes of macromolecular interpenetration and macromolecular compression are very similar; only interpenetration is therefore
considered in detail he1
When the interpenet
(Fieure 5). there is a chanee in tilbbs free enerev l A t i n l .

Bridging Flocculation
The flocculation of electrostatically stabilized particles
by the addition of strong electrolytes has been mentioned.
A second method for inducing flocculation is to add a
small concentration-e.g., IO ppm-of a high molecular
weight polymer-e.g., 2 1 x lo6. The use of polymers
for flocculating dispersed particles is becoming increasingly
prevalent in the fields of water treatment and mineral
processing.
Considerable evidence has now accrued to suggest that
these polymeric flocculants function by adsorbing onto
the surfaces of more than one particle a t a time and
forming bridges between the particles (LaMer and Healy,
1963; Ruehnvein and Ward, 1952).
Suitable polymeric flocculants may he nonionic-e.g.,
polyacrylamide is nominally so over a certain pH rangecationic-e.g., polyethylenimine-or anionic-e.g.,
polyacrylamide with 30% of the amide residues hydrolyzed.
Where the surfaces of the particles and the polymer are
oppositely charged, flocculation may occur by a t least
two different mechanisms: by particle bridging, or by
the macromolecules reducing the surface charge density
of the particles through charge neutralization. Recent
experimental results suggest that relatively low molecular
weight (ea. 1 x IO4) cationic polyelectrolytes flocculate

Figure 4. Collision of two sterically stabilized particles

Figure 5. Interpenetration of two polymer chains


Ind. Eng. Chem. Prod. Res. Develop., Vol. 9,No. 4, 1970 471

Table I. Three Ways of Obtaining


a Positive AGE for Stabilization
AHn

si

+
+

+
-

I IH~I / I r s i 1
<1
>1
$1

Stability Type

Entropic
Enthalpic
Combined enthalpic-entropic

This is given by the usual Gibbs-Helmholtz relationship


1G.q = AHR - TASK

(7)

AGR may assume values which are zero, positive, or negative. If AGR is negative, the particles are sensitized to
flocculation by the polymer chains. If AGR is zero,

flocculation proceeds (to a first approximation only) as


if the macromolecules were absent. Stabilization may be
observed if AGR is suitably positive.
Positive values of AGHmay be obtained in three different
ways (Table I). I n the first case, both the entropy (ASH)
and enthalpy (A)
of interpenetration are negative, but
the entropy contribution to AGR outweighs that of the
enthalpy. We may term this entropic stabilization (van
der Waarden, 1950), as entropic effects promote stability
while enthalpic interactions oppose it.
The second possibility reverses the signs of both AHR
and ASR,and the relative magnitudes of their contributions
to AGR. Enthalpic interactions are stabilizing and outweigh
the destabilizing entropic effects. This is termed, by
analogy, enthalpic stabilization.
I n the third situation, A H R is positive and ASR is negative. Both the enthalpy and entropy of interpenetration
contribute to stability and so we have combined enthalpicentropic stabilization.
The generic term steric stabilization thus encompasses
entropic, enthalpic, and combined enthalpic-entropic stabilization.
To distinguish between entropic and enthalpic stabilization experimentally is sometimes simple. For an
entropically stabilized dispersion, 1 TASR1 is greater than
I A H ~ I ,but this usually becomes decreasingly so as T
is reduced. Entropically stabilized dispersions thus often
flocculate on cooling (Napper, 1968b). In contradistinction, enthalpically stabilized systems normally flocculate
on heating. The T A S R term is less than A H R in this
instance, but the difference between them decreases as
T is raised. One well-known example of this type of behavior is the flocculation on heating of aqueous emulsions
and dispersions which are stabilized by poly(ethy1ene
oxide) adducts (Tuapper, 1969a).
The simple temperature change criterion is not
definitive, however, because both ASR and A H H may be
temperature-dependent. Moreover, certain dispersions
may flocculate both on heating and cooling. Otherse.g., those stabilized by combined enthalpic-entropic
effects-do not in principle flocculate a t any accessible
temperature. An unambiguous method for determining
the type of stabilization is clearly required.
The best procedure to date is to determine the signs
of AHK and l S R , upon which the original classifications
were based. The simplest method converts the process
of intermolecular interpenetration, shown in Figure 5 , into
one of intramolecular interpenetration. This is easily
achieved by joining the polymer chains end-to-end at
472

Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970

one point (say, A in Figure 5 ) . Intramolecular


interpenetration influences the conformational properties
of the macromolecule and these we can measure readilye.g., by viscometry (Moore, 1967) or light scattering
(Kerker, 1969). Measurement of the size of a macromolecule in a given solvent enables the signs of A H 8
and ASK in that solvent to be determined unequivocally,
using the Flory theory (Flory, 1953). Unfortunately, the
self-excluded volume problem is. in general, still unsolved
for macromolecules; we must treat the magnitudes of AHH
and ASH derived in this fashion with some caution (Flory,
1953).
The measurement of one other thermodynamic parameter, the theta (0)-temperature, is necessary to permit
the determination of the signs of AH,? and ASK. The
@condition is of central importance to the theory of steric
stabilization. We therefore consider this somewhat unusual
notion.
Theta-Conditions

Flory (1953, 1969) has suggested a useful, though by


no means complete, gas analogy to the &condition. The
equation of state for n moles of an ideal gas is essentially
Boyles law:

P V = n RT

(8)

This equation would be obeyed by an ideal system of


noninteracting point molecules. Real gas molecules may
be made to approach ideal behavior by reducing the pressure sufficiently. Under normal conditions, however, moderate deviations from ideality appear (Figure 6).
Van der Waals proposed a semiempirical modification
to the 1.h.s. of the ideal equation, to allow for such deviations:

( P + a [ n ?V ] ) ( V - n b ) = n R T

(9)

The b-term allows for the finite volume occupied by each


gas molecule, creating a spatial volume which is unavailable t o other molecules (hence the negative sign). This
geometrical excluded volume term is thermodynamically
an entropic correction factor. The a-term allows for the
attractive forces between the molecules, often predominantly London dispersion forces. This interactional term
is therefore an enthalpy correction factor. Both correction
terms arise because each gas molecule can see, in a
thermodynamic sense, neighboring gas molecules.
The van der Waals equation of state reduces to the
2 r

REAL GAS

//\

0
0

100

200

300 L O O 500 600


P R E S S U R E (atrn.)

Figure 6. Nonideal behavior of real gas molecules

Figure 7. Hypothetical joining of gas molecules


to form a polymer chain

0
A

ideal equation a t very low pressures, when V is large.


This reduction can also occur at high pressures-e.g., 500
atm-if the negative excluded volume term exactly counterbalances the positive interactional term (Figure 6). As
Boyle's law is then obeyed, this is referred to as the
Boyle point. At the Boyle point, gas molecules behave
ideally as they cannot see one another, because of the
chance cancelling of the entropic and enthalpic effects.
At first sight there might appear little to relate the
behavior of the minimolecules of a gas with that of polymeric macromolecules. However, if a large number of gas
molecules are joined as shown in Figure 7 , a pearl necklace
model of a polymer chain results. I n reality, this might
be the polymerization of ethylene or ethylene oxide, for
example. Qualitatively, the familiar ideas outlined for gas
molecules can be applied to considerations of the thermodynamics of polymer solutions. Due allowance must, of
course, be made for the presence of solvent molecules.
Dilute solutions of small molecules obey the van't Hoff
analog of the ideal gas equation:

T V = nRT or

=/e-,
= RTIM,

(10)

where K is the osmotic pressure and C? is the weight


concentration of solute of molecular weight Mr.
Macromolecules in solution would not be expected to obey
the ideal van't Hoff equation. In good solvents, the
reduced osmotic pressure of polymer molecules shows
strong positive deviations from ideality as a consequence
of macromolecular repulsion (Figure 8). I n bad solvents,
negative deviations ensue. I n a 0-solvent, macromolecules
behave ideally, even a t relatively high concentrationse.g.. several per cent.
Theta-conditions in a binary polymer-solvent mixture
occur a t a particular temperature, the @temperature.
which is independent of molecular weight. Thetaconditions can in principle also be obtained in a ternary
polymer-solvent-nonsolvent mixture. In practice, however,
we must exercise some caution in considering 0 compositions (Dondos and Benoit, 1968). Dondos and
Benoit (1969) have even suggested that in ternary systems,
two slightly different ()-conditions may exist, one relating
to intermolecular interpenetration and the other to
intramolecular interpenetration. I t is the former which
is relevant to steric stabilization.
Ideal behavior in @solvents has several important corollaries. We need stress here only the corollary relating
to the macromolecules behaving as if they were noninteracting point molecules. Each macromolecule does not
see its neighbors thermodynamically, in analogy to the
behavior of a gas a t the Boyle point. Macromolecules
can telescope one another: without net segmental interaction. Thus AGK for interpenetration is zero, due to the
exact equality of AH,?and TASK (cf. the Boyle point).
The traditional methods for determining ()-conditions
include the estimation of solvents in which the macromolecules behave ideally, as determined by osmometry,
light scattering, or the ultracentrifuge (Elias et al., 1966).
Viscometry and phase separational procedures can also
be employed (Flory, 1953).

The primary dissatisfaction with these older methods


stems from the time (often a week or more) and sheer
physical effort needed to prepare suitable samples and
perform the requisite measurements. Recently several
rapid turbidimetric procedures have been devised (Cornet
and van Ballegooijen, 1966; Napper, 1968a, 1969e; Suh
and Clarke, 1968; Talamini and Vidotto, 1967), all of
which are developments of the pioneering studies of Elias
(1959, 1961). Measurements are simply made of the conditions required for incipient phase separation at different
volume fractions of polymer ( 5 lee). These enable
0-conditions to be found rapidly-e.g., in less than one
hour-often
using unfractionated material. The results
obtained are comparable to those determined by the older
methods. These procedures are not applicable, however,
to systems where the polymer separates as a semicrystalline solid (Flory, 1953).
Model Dispersions

Dispersions which have proved suitable for investigating


steric stabilization contain particles such as the one shown
schematically in Figure 9. The disperse phase is a polymer,
insoluble in the dispersion medium. The surfactants which
impart steric stabilization are amphipathic, nonionic

@-SOLVENT

B A D SOLVENT

11""
4 1
0

POLYMER

CONCEN TRATlON

I%)

Figure 8. Reduced osmotic pressure in different types of


solvents

.-----.I

ANCHOR POLYMER

DISPERSE
PHASE

MOIETIES
DISPERSION
MEDIUM

Figure 9. Typical particles in sterically stabilized dispersions


Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970

473

/-

/
/
/

Figure 10. Schematic representation of entropic


stabilization

macromolecules: One constituent, termed the anchor polymer, is nominally insoluble in the continuous phase; the
other, termed the stabilizing moiety, is soluble. The anchor
polymer attaches the stabilizing moieties to the surfaces
of the particles. The precise nature of the anchor polymer
is not critically important to steric stabilization. Its omission, however, leads to a dramatic reduction in stability.
The stabilizing moieties usually surround the particles
completely. I t is the solution properties of these moieties,
and not those of the entire surfactant molecule, which
govern the stability of these dispersions (Napper, 196813,
1969~).
A short hand notation for representing these dispersions
is to write them as follows: disperse phaseianchor
polymer/ stabilizing moieties/ dispersion medium.
Entropic Stabilization. A typical dispersion which is
entropically stabilized is (Osmond et al., 1966a,b):
poly(methy1 methacrylate) /poly (methyl methacrylate) /
poly(l2-hydroxystearic acid) /n-heptane.
The origin of the repulsive potential energy in the case
of entropic stabilization is illustrated in Figure 10. The
gas analogy may be invoked: Work must be performed
on a gas to compress it. This is in accord with the second
law of thermodynamics. Compression of a gas decreases
its configurational entropy as the atoms or molecules
become more ordered. By analogy, interpenetration and
compression of polymer chains also lead to a decrease
in segmental entropy. By themselves these processes are
not therefore spontaneous; work must be expended to
interpenetrate and compress any polymer chains existing
between two colloidal particles. This work is a reflection
of the repulsive potential energy.
Of course, the above discussion has ignored the van
der Waals attractive forces operative between the gas
molecules. These assist the compressional process. The
analogous enthalpy effects would be expected to promote
the flocculation of sterically stabilized dispersions.
The foregoing concepts are a deliberate oversimplification designed to highlight the principles involved.
A more complete entropy discussion allows for the presence
of the solvent: Entropic repulsion originates in the loss
of the entropy of mixing of polymer segments with the
molecules of the dispersion medium when interpenetration
and/or compression occurs. Only the excess or nonideal
entropy of mixing is important and, t o a first approximation, only the first-order terms of the excess function
are relevant (Napper, 196813). Comparable restrictions
apply to the excess enthalpy of mixing function.
474

Ind. Eng. Chern. Prod. Res. Develop., Vol. 9 , No. 4, 1970

I n @solvents macromolecules behave as if they are


noninteracting point molecules. The Gibbs free energy
of interpenetration is zero and so sterically stabilized dispersions ought to be unstable in dispersion media which
are @-solventsfor the stabilizing moieties (Fischer, 1958;
Napper, 1968b; Ottewill, 1967). Entropically stabilized
dispersions may be flocculated by a t least two different
methods: by reducing the 1 TASRI term-e.g., by coolingor by increasing the I~HKI
term, if T is held constant.
As discussed in a previous plenary article (Hansen, 19691,
the latter can be achieved readily by adding a nonsolvent
for the stabilizing moieties to the dispersion medium.
The catastrophic flocculation of an entropically stabilized dispersion, on addition t o the continuous phase of
a nonsolvent for the stabilizing moieties, is illustrated
in Figure 11. The dramatic increase in the turbidity of
the dispersion when flocculation occurs enables us to define
a critical flocculation volume (c.f.v.). This is the minimum
volume fraction (5%) of nonsolvent which generates incipient flocculation.
A comparison is presented in Table I1 of some c.f.v.'s
with their corresponding &compositions. The results
illustrate the strong correlation existing between these
two parameters. Entropic stabilization persists until the
attainment of @conditions in the dispersion medium for
the stabilizing moieties. Flocculation then ensues (Napper,
196813).

1.0

r-iI

Figure 11. Catastrophic onset of flocculation on addition


of nonsolvent for stabilizing moieties (PSA)

Table II. Comparison of 0-Compositions and C.f.v.s.


in n-Heptane Using Ethanol as Flocculant
Stabilizing
Moieties

0-Composition,
Temp., ' K

C.f.v., Yo

O h

PSA

274
297
313

39.5
50.5
58

39
51
59

P1,iLI'

249
274
297
313

31
38
44
51

33
39
45
51

PSA = poly(l2-hvdroxvstearic acid). ' P L M = poly(laury1


methacrylate).
"

Enthalpic Stabilization

A dispersion which has been shown t o be enthalpically


stabilized is: poly(viny1 acetate) /poly(vinyl acetate) /
poly(ethy1ene oxide) /water.
One possible origin of the stabilizing enthalpic interaction for this dispersion is illustrated in Figure 12. There
is circumstantial evidence to suggest that poly(ethy1ene
oxide) is hydrated in aqueous solution (Rosch, 1956). This
means that water molecules are associated with the
poly(ethy1ene oxide) in such a way as to lose some degrees
of freedom-e.g.,
rotational degrees of freedom. This
specific interaction may be a consequence of hydrogen
bonding between the water molecules and the ether
oxygens of the poly(ethy1ene oxide) chains.
When interpenetration and compression of poly(ethy1ene
oxide) chains occur in a Brownian collision, there is an
increased probability of ethylene oxide-ethylene oxide contacts in the interaction zone. Consequently, some of the
hydrated water molecules are excluded from this region.
These released water molecules have greater degrees of
freedom than those in the hydrated state. The equipartition of energy principle implies that the new degrees of
freedom must be supplied with energy-viz., 0.5 kT per
mole per mode. It is this energy requirement which provides a positive enthalpy of interpenetration, capable of
imparting stability (Napper, 1 9 6 9 ~ ) .

\'

' /+

n H-0

-, k I

Figure 12. Schematic representation of enthalpic stabilization

Table Ill. Comparison of C.f.t.'s and ()-Temperatures


for Aqueous Dispersions Stabilized by
Poly( ethylene Oxide)
Electrolyte

2M LiCl
2 M NH Cl
2 M XaC1
0.45M KZSO,

C.f.t.,

359
349
332
305

0-Temperature, ' K

363
349
333
307

The origin of the destabilizing entropic effects is also


obvious from this model. There is a decrease in segmental
entropy in the interaction zone, as in entropic stabilization.
But this is overridden by the increase in configurational
entropy of the excluded water molecules. The latter, incidentally, does not occur in the case of combined enthalpicentropic stabilization.
Aqueous dispersions stabilized by poly(ethy1ene oxide)
chains are stable on boiling, if all electrolyte is removed.
But the presence of even small concentrations of electrolyte
in the dispersion medium induces flocculation on heating.
The transition from long-term stability to catastrophic
flocculation occurs over only a 1' to 2'K temperature
range. The critical flocculation temperature (c.f.t.1 is
defined in this case as the minimum temperature required
to induce incipient flocculation. Some c.f.t.'s of poly(ethylene oxide) stabilized dispersions are compared
with the corresponding 0-temperatures for poly(ethy1ene
oxide) in Table 111. As in the case of entropic stabilization,
a strong correlation between these two parameters is evident (h'apper, 196813, 1969a,b,c). Stability is exhibited
until the dispersion medium is transformed into a 0-solvent
for the stabilizing moieties.
I t might be inferred from the above discussion that
the effectiveness of different ions in promoting flocculation
is dependent upon their ability to compete successfully
with the poly(ethy1ene oxide) for the water. Thus highly
hydrated ions, such as lithium, magnesium, and calcium,
would be expected to be very effective flocculants, whereas
poorly hydrated ions-e.g., rubidium and cesium- would
be predicted to be poor flocculants. Just the reverse is
found to be true experimentally: Rubidium and cesium
are among the most effective flocculating ions, while lithium and magnesium ions are among the least effective
(Napper, 1969a). These results imply that flocculation
is not merely a question of dehydrating the poly(ethy1ene
oxide) chains. Of more importance is the influence of
the electrolyte on the association properties of water
(Napper, 1969a). N M R results also suggest that dehydration is not primarily responsible for flocculation induced
by heating (Liu and Parsons, 1969).
Enthalpic stabilization is probably most commonly
encountered in aqueous dispersions. Yet dispersions stabilized by poly(isobuty1ene) in n-pentane should also be
enthalpically stabilized (Eichinger and Flory, 1968). Since
specific interactions are presumably absent in this instance,
some doubt must be cast on the role of specific hydration
effects in enthalpic stabilization. Recent inelastic neutron
scattering results also appear to be a t variance with the
orthodox notions of hydration of poly(ethy1ene oxide) in
aqueous solution (Assarsson et al., 1969). A proper understanding of enthalpic stabilization poses one of the most
challenging problems in steric stabilization. The answer
seems to lie in the free volume dissimilarity approach
Ind. Eng. Chem. Prod. Res. Develop., Vol. 9, No. 4, 1970

475

to polymer solution thermodynamics (Patterson, 1969).


This approach may also be applied to aqueous dispersions
stabilized by poly(ethy1ene oxide) chains of molecular
weight less than lo3, when even the qualitative application
of Flory-Huggins theory must be in doubt. However, conventional Gibbsian, large systems thermodynamics may
no longer be applicable and resort to small systems thermodynamics (Hill, 1964) would then seem inevitable.
Combined Enthalpic-Entropic Stabilization

Examples of combined enthalpic-entropic stabilization


are rare. This is not too surprising, since the requirement
for this type of stabilization implies a negative
@temperature. Negative @temperatures are thermodynamically permissible; but they are the exception rather
than the rule.
Dieu (1954) has shown that for poly(viny1 alcohol) in
water, the thermodynamic parameters correspond to those
required for enthalpic stabilization. Addition of 33% dioxane to the water, however, changes the signs of these
parameters so that entropic stabilization would be expected. At lower dioxane concentrations a domain corresponding to combined enthalpic-entropic stabilization is
anticipated. Experiments with latex particles stabilized
solely by poly(viny1 alcohol) moieties tend to support
these predictions (Napper, 1969d). I t seems likely that
transformations from enthalpic to entropic stabilization,
or vice versa, often pass through a region of combined
enthalpic-entropic stabilization.
Technological Implications of Steric Stabilization

Several problems arise with aqueous electrostatically


stabilized dispersions in practical situations. The first is
their sensitivity to electrolytes. Often ionic strengths as
small as 0.01 are sufficient to induce flocculation of electrostatically stabilized dispersions. Dispersions stabilized sterically by poly(ethy1ene oxide) a t room temperature, for
example, permit the addition of electrolyte a t concentrations two orders of magnitude larger (say, 1 t o 10M)
than do electrostatically stabilized dispersions. Steric stabilization may thus be gainfully employed if the dispersion
medium has a high ionic strength.
A second difficulty associated with electrostatic stabilization derives from the electroviscous effect. When a charge
stabilized latex is made to flow, the dispersed particles
move relative to one another. The presence of electrical
double layers means that extra energy must be expended
to move the counterion atmospheres, one past the other.
This additional energy loss is reflected in an increase in
viscosity. This increase depends upon the solids content
of the dispersion. But it sets a practical limit, sometimes
as small as 20% solids, to the concentration of dispersions
which may easily be pumped around a factory. With
a proper choice of stabilizer, sterically stabilized dispersions
with a high solids-e.g., >50%-content may be prepared
in water with convenient rheological properties. The preparation of the base polymer for certain emulsion paints,
for example, utilizes this principle. although often an
electrostatic contribution to stability is also included.
I n nonaqueous dispersions, yet another difficulty with
electrostatic stabilization arises-for example, the ionic
strength in a n-heptane dispersion is so small that the
double layer thickness is very large. This means that
the falloff in repulsive potential energy, and hence in
the total potential energy, is very slow (Figure 2 ) . Con476

Ind. Eng. Chern. Prod. Res. Develop., Vol. 9, No. 4, 1970

sequently, the physical expenditure of energy associated


with the mere act of preparing a high solids dispersion
is such as to overcome most of the repulsive potential
energy between the particles. The barrier to flocculation
is then insufficient to prevent its subsequent occurrence.
Electrostatically stabilized dispersions in n-heptane are
often stable to only CCI. l C csolids. High solids-e.g.,
> 60C;--dispersions in n-heptane are easily prepared if
steric stabilization is employed.
Sterically stabilized dispersions often undergo reversible
flocculation. This is important because it imparts good
freeze-thaw stability to such dispersions.
One disadvantage of steric stabilization probably
appears in the emulsion polymerization of certain
monomers-e.g., vinyl acetate. Seed particles, which are
sterically stabilized, have been shown to grow a t a rate
only half of that for electrostatically stabilized particles.
The stabilizing chains apparently inhibit the entry of oligomeric free radicals into the seed particles. Sterically
stabilized systems are thus more prone to new nucleation
(Netschey et al , 1969).
literature Cited

Assarsson, P. G., Leung, P. S., afford, G. H., Abstracts


of Papers, Division of Polymer Chemistry, 158th
Meeting, ACS, New York, N. Y., September 1969.
Casimir, H. B. G., Polder, D., Phys. Rev. 73, 360 (1948).
Chapman, D. L., Phil. M a g . 25, 475 (1913).
Cornet, C. F., van Ballegooijen, H., Polymer 7, 293 (1966).
Deryagin, B. V., Abrikossova, I. I., Discuss. Faraday Soc.
18, 24 (1954).
Deryagin, B. V., Abrikossova, I. I., Lifshitz, E. M., Quart.
R e v . ( L o n d o n ) 10, 295 (1956).
Deryagin, B. V.. Landau. L.. Acta Physicochirn. GRSS
14, 633 (1941).
Dieu, H. A., J . Pol3,rn. Sci. 12, 417 (1954).
Dondos, A., Benoit, H., Europ. Poly. J . 4, 561 (1968).
Dondos, A,, Benoit, H., J . Polym. Sci. B7, 335 (1969).
Eichinger, B. E., Flory, P. J., T r a n s . Faraday Soc. 64,
2066 (1968).
Elias, H.-G., Makromol. Chem. 33, 140 (1959).
Elias, H.-G., Makromol. Chem. 50, 1 (1961).
Elias, H.-G., Adank, G., Dietschy, H.. Etter, O., Gruber.
L.,Ibrahim, F. W., Theta-Solvents in Polymer
Handbook, p. IV-163, Interscience, New York, 1966.
Fischer, E. W., Kolloid-2. 160, 120 (1958).
Flory, P. J., Principles of Polymer Chemistry,! p. 576,
Cornell University Press, Ithaca, PI;. Y., 1953.
Flory. P. J., Statistical Mechanics of Chain Molecules,
p. 34, Interscience, New York, 1969.
Fowkes, F. M., I n d . Eng. Chern. 56, 40 (1964).
Gouy, G., Ann. Physik 7, 129 (1917).
Gouy, G., J . Phys. 9, 457 (1910).
Gregory, J., T r a n s . Faraday SOC.65, 2260 (1969).
Hamaker, H. C., Physica 4, 1058 (1937).
Hansen, C. M., IND. ENG. CHEM.PROD.RES. DEVELOP.
8, 2 (1969).
Heller, W., Pugh, T. L., J . Chem. Phys. 22, 1778 (1954).
Hill, T . L., Thermodynamics of Small Systems, R.
A. Benjamin, S e w York, 1964.
Jones, M. N.,
Kolioid-2.2. Pol?m. 230, 236 (1969).
Kallmann, H., Willstatter, M., l\raturuissenschajen. 20,
952 (1932).
Kerker, M., Scattering of Light and Other Electromagnetic Radiation, p. 433. Academic Press, Kew
York, 1969.
Kitchener, J . A,, Prosser, A. P., Proc. Roy. SOC.(London)
Ser. A 242, 403 (1957).

La Mer, V. K.. Healy, T. W., Reu. Pure Appl. Chem.


13. 112 (1963).
Liu, K.-J., Parsons, J. L., Macromolecules 2, 529 (1969).
London. F., 2. Physih. 63, 245 (1930).
Lyklema, *J., Adcan. Colloid Interface Sei. 2, 65 (1968).
Lyklema, ,J., Uiscuss. Faraday Soc. 42, 81 (1966).
Lyklema, J . , Pontif. Acad. Sei. Scripta Varia 31, I (1967a).
Lyklema, .J.. Pontif. Acad. Sei. Scripta Varia 31, I1
(1967b).
Moore, W. R.. Progr. Polym. Sci. 1, 1 (1967).
Napper. D. H., Abstracts of Papers, Division of Colloid
and Surface Chemistry, 138th Meeting, ACS, Kew York,
N. Y.. September 1969a.
Xapper, D. H., J . Colloid Interface Sei. 29, 168 (196913).
Napper. D. H.. J . Colloid Interface Sei., in press, 1969c.
Sapper, D. H., Kolloid-2. Z . Polym. 234, 1149 (1969d).
Napper. D. H.. Makromol. Chem. 120, 231 (1968a).
Napper, D. H., Polymer 10, 181 (1969e).
Napper. D. H., Sci. Progr. (Oxford) 55, 91 (1967).
Napper. D. H., Trans. Faraday Soc. 64, 1701 (196813).
Xetschey, A., Napper, D. H., Alexander, A. E., J . Polym.
Sci. R7, 829 (1969).
Osmond, D. W. J., Waite, F. A,, Walbridge, D. J. (to
Imperial Chemical Industries Ltd.) , Belg. Patent 667,784
(1966a): CA 65, 7388c.
Osmond, D. W. J., Waite, F. A., Walbridge, D. J. ( t o
Imperial Chemical Industries Ltd.) , Belg. Patent 667,785
(196613); CA 65, 7403a.
Ottewill, R. H.. Effect of Xonionic Surfactants on the
Stability of Dispersions in Nonionic Surfactants, Vol.
1, p. 627, Marcel Dekker, Xew York, 1967.
Ottewill, R. H., Shaw, J. N., Discuss. Faraday Soc. 42,
154 (1966).
Ottewill, R . H., Wilkins, D. J., Trans. Faraday Soc. 58,
608 (1962).

Overbeek, J. Th. G., Discuss. Faraday Soc. 42, 7 (1966).


Overbeek, J. Th. G., Interaction between Colloidal
Particles, in Colloid Science, Vol. 1, Elsevier,
Amsterdam, 1952.
Patterson, D.: Macromolecules 2,672 (1969).
Rosch, M., Koiloid-2. 147, 78 (19563.
Ruehrwein, R. A., Ward, D . W., Soil Sei. 73, 485 (1952).
Schenkel, J. H., Kitchener, J. A., Trans. Faraday SOC.
56, 161 (1960).
Suh, K. W., Clarke, D. H., J . Appl. Polym. Sei. 12,
1775 (1968).
Tabor, D., Winterton, R . H. S., Nature 219, 1120 (1968).
Tabor, D., Winterton, R . H. S., Proc. Roy. Soc. (London),
Ser. A, 312, 433 (1969).
Talamini, G., Vidotto, G., Makromol. Chem. 110, 111
(1967).
Usui, S., Yamasaki, T., Shimouzaka, J., J . Phys. Chem.
71, 3194 (1967).
van der Waarden, M., J . Colloid Sei. 5, 317 (1950).
Verwey, E . J. W., Overbeek, J. T h . G., Theory of the
Stability of Hydrophobic Colloids, Elsevier, Amsterdam, 1948.
Vold. M. J., J . Colloid Sei. 16, 1 (1961).
Watanabe. A., Gotoh. R., Kolloid-Z. Polym. 191, 36
(1963).
Watillon, A., Joseph-Petit, A.-M, Discuss. Faraday Soc.
42, 143 (1966).
Wiese, G. R., Healy, T. W., Trans. Faraday Soc. 66,
490 (1970).

RECEIVED
for review February 2, 1970
ACCEPTED
May 29, 1970

Ind. Eng. Chem. Prod. Res. Develop., Vol. 9,No. 4, 1970

477

Das könnte Ihnen auch gefallen