You are on page 1of 9

Colloids and Surfaces A: Physicochem. Eng.

Aspects 440 (2014) 161169

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Joint development of insight into colloid stability and surface conduction


Johannes Lyklema
Department of Physical Chemistry and Colloid Science, Wageningen University, Dreijenplein 6 6703, HB Wageningen, Netherlands

h i g h l i g h t s

g r a p h i c a l

a b s t r a c t

 The valence dependence of the coagulations concentration is lower than


according DLVO theory.
 Monovalent ions in the Stern layer
are laterally mobile.
 Interpretation of streaming potentials mostly requires insight into
surface conduction.
 All three double layer parts participate in charge regulation upon
interaction.

a r t i c l e

i n f o

Article history:
Received 2 August 2012
Received in revised form
14 September 2012
Accepted 16 September 2012
Available online 23 September 2012

a b s t r a c t
This paper presents a historical overview of the parallels between the developments of colloid stability and surface conductivity. Starting from the situation during the Second World War, the interaction
between the developments of these two branches of science appeared mutually benecial. In particular,
the properties of the non-diffuse parts of the double layers drew much attention. Implementations in the
direction of future developments are given.
2012 Elsevier B.V. All rights reserved.

Keywords:
Colloid stability history
Surface conduction history
Stern layers
SchulzeHardy rule
Charge regulation

1. Introduction
The rst week of September 1939 was one of the most dramatic periods of the previous century. That week saw the outbreak
the Second World War in Europe which later expanded to become
world-wide. This war not only meant a break in the development of
civilization, but also in that of colloid science. Specically, the contacts between colloid scientists in the Eastern and Western parts of
Europe were severed during a sensitive period where major steps
were being made in developing insight into colloid stability.
In view of the signicant progress during the 1930s in understanding colloid stability in general, and electric double layers in

E-mail address: Hans.Lyklema@wur.nl


0927-7757/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfa.2012.09.013

particular, the British Faraday Society planned a Formal Discussion


on The Electrical Double Layer, to be held in Cambridge (UK),
in September 1939. A host of scientists from all over the world
were invited to contribute. In line with the style of the Faraday Discussions, these participants submitted their contributions of which
printed versions were circulated with the aim to start the formal
discussions in Cambridge on the basis of these papers. The outbreak
of the war prevented these Discussions from taking place, but the
Faraday Society decided to publish the preprints as a special issue
of their Transactions [1].
For the sake of history of colloid science, this publication contains a bounty of information, more or less reviewing the state of the
art at the beginning of the war. Reading all of this is addicting for historians because this literature relays how much was known about
that time, which problems were formulated and where scientists
were looking for solutions.

162

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

The aborted Faraday Discussion is also historically challenging


theme because a situation arose where, as a consequence of the
war, scientic exchange between different countries was severely
impeded, if not completely suppressed, so that local developments
started to take place without interaction with ambient countries.
Contacts across military frontiers were virtually non-existent, so
that the opportunities for research depended on which side of the
frontier between the belligerent nations you found yourself. The
interesting point is that these independent developments in different countries started from the same premises, viz. the printed
version [1]. Another point is that the facilities for doing basic scientic research were also directly challenged by the war conditions,
partly because of growing logistic problems because of the pressure
for carrying out military research.
Of particular interest for the present topic, and familiar to the
older generation of colloid scientists, was the independent development of theory for the stability of lyophobic colloids by two
Russians, Deryagin and Landau, and two Dutchmen, Verwey and
Overbeek. It took till several years after the war before it transpired
that the two parties had developed essentially the same model. This
theory is now known as DLVO theory [2,3]. It is less known that
during the same period also important steps forward have been
made in the domain of electrokinetics. In particular, the role of surface conduction became clearer. Also in this case there were two
players: one a Russian from England (Bikerman) and the other a
Dutchman from Belgium (Rutgers). Both showed that incorporation of surface conduction into electrokinetic equations eventually
leads to a better establishment of electrokinetic or -potentials.
In turn, this development was benecial for the interpretation of
stability experiments.
In the following paper the juxtaposition of the stability and electrokinetic developments over that period will be discussed. History
will not be only considered for its own sake; an extrapolation to the
present time will also be made.
In August 1939 several aspects of colloid stability were
well-established. The phenomenological distinction between
hydrophobic and hydrophilic colloids was generally accepted, but
the role of polymers was confusing as some polymers could stabilize or destabilize the same hydrophobic sol, depending in the
way of mixing. For hydrophobic sols reasonably reproducible coagulation concentrations (ccr ) were available. Modern scientists may
nd this surprising since no interaction forces measurements were
available and because the sols of that time were heterodisperse. The
fact is that painstaking coagulation series were carried out: series of
test tubes, containing sols with increasing salt concentrations were
examined to nd out above which salt concentration the sol coagulated. Methods, involving shaking procedures were developed to
improve establishing the difference between stable and unstable
sols. Nowadays we recognize these steps as a typical combination
of perikinetic and orthokinetic coagulation, but at that time the
development was just phenomenological, the result of painstaking
trial and error, without insight into the theoretical background. In
1939 it was an established fact that ccr decreased very rapidly with
the valence of the counterion. This rule dates back to the turn of the
19th20th century and is now known as the SchulzeHardy rule,
after [4,5]. However, the reason for this very strong dependence
was unclear and a topic of intense dispute. Lyotropic, or Hofmeister, series for the specic ionic effects on ccr were also known.
Under dispute was the surface charge-determining mechanism,
although it was generally felt that a strong chemical afnity of
certain ions to a surface was needed to create a non-zero surface charge. In 1939 charge reversal, or overcharging, was also
a familiar phenomenon, not in the least thanks to Freundlichs
research.
In this paper we shall emphasize the developments in the interpretation of the stability of hydrophobic colloids and collateral

matters, in particular the issue of dening and measuring interfacial


potentials.
The present paper is a sequel to a historical paper by Vincent
[6], emphasizing pre-DLVO studies of particle aggregation and one
on electrokinetics by Wall [7].

2. The aborted Faraday Discussion. Interfacial potentials


and charges
The contents of [1] offer pabulum of food historians. The collection contains almost 40 papers related to colloid stability, divided
into three parts: I. Experiments, II. Theory and III. Applications, consecutively. Part I dealt with electrophoresis, streaming potentials,
surface conductance and electrocapillarity, part II with electric double layers, colloid stability and adsorption potentials and part III
with biological and technical applications. Given the emphasis of
the present paper this last section will not be considered, but it
must be noted that already at that time scholars were aware of
applications of colloid science in domains as diverse as virus stability, rubber technology, deposition, wetting and otation. At that
time part of the search was for the spontaneous creation of potential
differences across interfaces, including those at liquidliquid and
airliquid phase boundaries. The notion of adsorption potentials
has become more or less obsolete, but the basic problem is as much
alive as it was in the older times, because it deals with the issue
of spontaneously creating potentials as the result of adsorption of
charged species.

2.1. Denition and measurement of interfacial potentials and


charges
The present understanding of interfacial potentials and charges,
their denition and measurabilitys, can be found in the textbooks [8]. Measurement of absolute potentials in condensed
systems is thermodynamically inoperable. This is a basic tenet
that has not changed over time. On the other hand, measurements of changes in potential can be experimentally carried out. The
measurement of surface charges ( o ) is possible, once the chargedetermining mechanism is known. Nowadays determination of
surface charges by titration is an established technique but for dispersed systems its systematic development had to wait till after the
war.
In 1939 only for two systems was the potential across a phase
boundary dened and understood, namely for those between
mercury and silver iodide-in contact with (aqueous) electrolyte
solutions. These two categories represented more or less the two
ends of the spectrum of potential-determining mechanisms: that
between polarizable and reversible interfaces. In the former case
(Hg) no charge transfer takes place through the interface so that
a potential across the interface can be externally applied. In the
second case, adsorption from solutions containing Ag+ and/or I
ions is responsible for the creation of a potential difference across
the interface. In the latter case, Nernsts law gives a direct relation
between the change in surface potential, d o and the change of the
negative logarithm of the Ag+ activity, leading to the well-known
58 mV per unit of pAg. This law can be derived thermodynamically
if the chemical potentials of the charge-determining ions inside the
solid are independent of the surface charge. On this basis, Verwey in
his thesis [9] introduced the notion of potential-determining ions.
Absolute values for o cannot be obtained in this way. A reference
point is needed, and for that the point of zero charge p.z.c. is usually
selected. The basic uncertainty is that the zero point of charge is not
necessarily identical to the zero point of potential [8]. It took several
decades before it became clear that also for materials like oxides

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

a similar mechanism has to exist, although Nernst-type behaviour


was already known since long for glass electrodes.
The denition and measurement of surface charges ( o ) does
not involve basic problems. For mercury, the differential capacitance C = (d o /d o ) can be measured by a Wheatstone bridge. Upon
integration it gives  o as a function of o but again a reference point
is needed. For this, the capillary maximum is the obvious choice. For
AgI,  o can be measured by titration of colloidal suspensions with
charge-determining electrolyte, i.e. with AgNO3 and/or KI. Because
of the better denition of charges it is nowadays preferred to talk
about charge-determining ions, rather than potential-determining
ions.
In this connection it is of interest to note that reversible interfaces can be made polarizable if the interfacial transfer process of
charge-determining ions is suppressed. A typical illustration is that
by Pieper and De Vooys [10], who suppressed the Faradaic current by working with AgI electrodes at very high frequencies. The
resulting  o (pAg) curves are almost identical to those obtained by
titration [11].
Returning to the Faraday meeting, it is interesting to consider now what can be found in Ref. [1] about this matter
three contributions address the polarizable interface (mostly the
mercury-electrolyte solution), viz. [1214]. Crawford [12] discusses
the differences between different types of interfaces, including
Langmuir monolayers, and the double layer structures at the solution side. For the latter he uses GouyStern theory. We may
conclude that even as long ago as 1939 experts already knew that
pure diffuse theory applies only under special conditions, viz. very
low surface charges and low electrolyte concentrations. Generally,
GouyStern models are needed. It is remarkable that this observation is nowadays often forgotten. Frumkin [13] discusses the inner
part of the double layer; including electrochemical processes occurring in it. He refers to the Tafel equation for the overpotential and
identies the potential occurring in it as the outer Stern potential d which, in turn, he identies as the electrokinetic potential
. This latter identication is nowadays also often made, but the
potential in the Tafel equation is now rather identied with the
inner Stern potential i . Frumkin was aware that electrokinetics
only involve the (tangentially) movable part of the double layer, a
point of view that he attributed to Freundlich and Smoluchowski.
From double layer capacitance curves he tried to establish the outer
Stern potential, but this procedure is cumbersome because so many
assumptions have to be made. Frumkin did not report own electrokinetic data; in fact for that the mercury system is not suited.
The BarclayButler paper [14] is rather short; its main contribution
is of an experimental origin and essentially supports the observation that the mercury electrode is within limits fully polarizable,
as judged by the absence of any effect of hydrogen production at
the surface. They also gave a set of  o ( o ) curves for mercury,
obtained from electrocapillary curves. In solutions of potassium
salts they found at the positive side, where the anion is the counterion, strong ion specicity, whereas at the negative side, where
potassium is the counterion, all curves coincide. This is an old, but
useful electrochemical example of lyotropic, or Hofmeister series.
So, in summary, in 1939 several potentials were recognized and
identied, the surface potential o , the inner Stern potential i
the outer Stern, or diffuse layer potential d and the electrokinetic potential , with the problem of identication of one of these
potentials with . The ensuing problem is which one to use in particle interaction theory. At issue are both measurability and physical
signicance. At that time distinguishing the corresponding charges
(the surface charge  o , the inner Stern charge  i and the outer Stern
charge  d which might be identied with the electrokinetic charge
 ek or  d , the charge in the diffuse part of the double layer) was
not popular. It is not surprising that this multitude of charges and
potentials gave rise to some confusion.

163

Faced with the complexity of this problem, it was a logical step


to look for alternatives, in particular by considering electrokinetic
potentials.
2.2. Denition and measurement of electrokinetic potentials
As long ago as 1900 Hardy [5] pointed out that there is a relation between colloid stability and . In their introductory paper
to the mentioned Faraday Discussion, Kruyt and Overbeek discussed this matter [15]. They accepted the idea of a stagnant
layer, stating that this layer is kept in place by strong electric
or adsorptive forces, thereby referring to Stern. Their way of
writing suggests that they consider stagnancy as a phenomenon
involving the counterions, not as something originating from uid
dynamics. Had they been right, there would be no stagnant layer
at the p.z.c. Accepting that in many cases Hardys statement is
valid, they also mentioned a number of cases where this rule did
not apply, thereby summing up some problems in experimentally
establishing . First, they were not certain whether it is correct to
substitute bulk values for the dielectric permittivity and viscosity in the HelmholtzSmoluchowski equations. They also mention
some cases where , computed from different methods, say electrophoresis and streaming potentials, are not identical and cases
where the dependence on the indifferent electrolyte concentration is anomalous, i.e. not continually decreasing. They were up
to date by mentioning the work on surface conduction by Gortner
and Rutgers, which removed some of these inconsistencies and to
which I shall return in Section 3.2. However, they underestimated
its impact, stating that the high surface conductivity is only a detail
of the much more important difference between the Helmholtz
theory and their more modern conception of the double layer.
The conclusion is that in 1940 not only the various static double
layer potentials, but also the elecrokinetic potentials were not yet
well-dened. In this connection, it is interesting to re-read the closing sentences of the KruytOverbeek paper where this point is put
into words. It is in an erratum to their paper that, due to the war
conditions was received only just before printing [16]. From this
we quote verbatim From the calculations of Hamaker and Levine it
has become clear that, although  is an important factor in the stability
of colloids, it is not the only determining one and that other quantities (size of the particles, concentration of the solution, van der Waals
attraction) have to be taken into account. These calculations, however, are so very complicated that the problem is still far from being
solved. . .. The intricacy of this problem is most clearly demonstrated
by realizing that after more than 30 years of strenuous research we
only just begin to see our way through this most fundamental problem
of the stability of colloids
Finally, Kruyt and Overbeek already prelude on the fact that,
after all, surface charges may be better dened than surface potentials. They mention some techniques anticipating the more recent
titration methods. One of these, attributed to Pauli, involves measuring the number of counterions that can be released by dialysis.
2.3. Colloid stability in the Faraday Discussions
Not less than seven contributions to the aborted Faraday Discussion dealt directly with colloid stability, viz. Refs. [1723]. It is not
surprising that here, at the dawn of modern colloid science, familiar
names and themes appear. Several illustrations of the now familiar
Gibbs energy-distance G(h) curves with maximums and minimums
can already be seen. Although the quality of many of these curves
does not match those of the post-DLVO era it was at least clear that
interaction is a balance between attractive and repulsive forces,
each of these with its own distance dependence. At issue was how
to quantify them, which required insight into their origins.

164

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

The occurrence of Hamaker as (one of the) author(s) in [17,18]


is proof of the insight that attraction is a general phenomenon
and caused by Van der Waals forces, nowadays rather specied
as dispersion forces. According to Hamaker and De Boer over the
years prior to the Faraday Discussion, these forces can be estimated as additive sums over pairs of London attractions between
atoms or molecules. We now know that these estimations are correct within something like 1020%, depending on the system. For
practical purposes, in 1940 the attractive part of the pair interaction could be considered as semiquantitatively solved. Against this
background several practical issues could already be addressed.
Hamaker and Verwey [17] discussed the formation of electrodeposited layers in non-aqueous solvents and made a number of
observations on the difference between sediments formed from
settling of stable and unstable colloids. In particular, redispersibility received attention. The discussion remained qualitative but
involved G(h) curves containing an additional potential energy contribution caused by the compression of the particles. With this
background the distinction between electrodepositing and sediment formation by gravity could be claried. Hamaker himself [18]
systematically investigated the effect of particle size, concluding
that the nature of the forces does not change with particle radius.
In other words, size effects are of a quantitative, rather than qualitative, nature. For the Van der Waals part Hamaker himself had
already elaborated the theory for spheres of different sizes [24].
The corresponding theory for the electrostatic part remained to be
done. Hamaker also mentioned the difference between, what we
now call, the HelmholtzSmoluchowski and the HckelOnsager
equations for the electrophoretic mobility. The former holds for
small particles (or, more precisely low a) the latter for high a. In
connection with our recent interest in nanocolloids it is interesting
to extrapolate this insight to very small particles, where the nature
of the particle starts to differ essentially from its macroscopic
counterpart [25].
Verwey [19] discussed the stability of emulsions. A large
part of the story deals with the distribution of potentials across
liquidliquid interfaces and his paper contains formulas for double diffuse double layers. He concludes that low molecular weight
electrolytes can never stabilize emulsions; emulsiers are needed.
Nowadays this is obvious.
The most penetrating discussion on colloid stability in [1] came
from Deryagin [20]. (Here we heed the Chem. Abstr. transcription of his name; Deryagin himself mostly went for the French
transcription Derjaguin.) His contribution contains some important elements. In the rst place he recalls his method to obtain
interactions between spherical surfaces on the basis of known
interactions between at plates. Nowadays this method is known
as the Deryagin approximation [26]. He needed this equation
because he wanted to apply his results to the coagulation of sols
of spherical particles. A second important step was that he realized
that for the electric contribution to the interaction the Gibbs energy
was needed rather than the internal energy. The difference is that
in the former case the change of the distribution of the ions in a
double layer is considered. Otherwise stated, in the Gibbs energy
the entropic part is accounted for Deryagins equation (18) in [20],
contains two terms, the former is the Gibbs energy of an electric
eld, which follows from general eld theory, involving Poissons
law and Greens theorem. The latter term represents the change in
the osmotic Gibbs energy of the ions, incurred by the redistributions of the ions upon charging the particles. In their later work,
Verwey and Overbeek [3] also followed the Gibbs energy path.
The distinction between the two types of energy was an important step forward in a dispute of that time because it could exclude
situations of electrostatic attraction between two surfaces of identical charge. Deryagin applied his theory to the rate of coagulation
[27]. He introduced a dimensionless parameter n, determined by

the potential, the dielectric constant, the particle radius a and the
temperature T:
n=

a 2
2kT

(1)

He concluded that for stability n must be at or above 10 and concluded that for sufcient stability  has to be above about 30 mV.
This is one of the rst concrete stability results though still primitive. For one thing, the Van der Waals part is not included. In the
later developments [2,3] that part was included with the height of
the (Gibbs) energy barrier as the stability criterion.
As to the potentials, Deryagin nds interaction at constant
potential more natural than at constant charge (footnote on p. 205
of [20]) and without argument identies o with  (p. 211). This
means that in 1940 he had not yet made up his mind on the composition of double layers and neither on the charge-determining
process. On the other hand, he introduced the application of his
theory to thin liquid lms and to hetero-interaction.
In a sense, the paper by Levine and Dube [21] is more or less
the counterpart of that by Deryagin. These authors do employ full
superpositions of Van der Waals and electrostatic forces, but do so
with simplied formulas. For the interaction they took the energy
instead of the Gibbs energy and they substituted relatively simple
formulas for the repulsion and attraction. As expected, they found
the stability to be determined by  and the Hamaker constant A.
Eilers and Korff [22] essentially empirically derived an expression for the energy Ub to overcome the barrier needed for bringing
two particles together against electrical repulsion:
Ub =

const. 2


(2)

This rule of Eilers and Korff is still sometimes used for a rapid
empirical assessment if interaction forces.
The CheesmanKing paper [23] contains some illustrations of
the inuence of a variety of electrolytes on the stability of emulsions.
Summarizing this part of the state of affairs of colloid stability,
in 1939 already many ingredients for a general and comprehensive
theory were recognized. They were, so to say, waiting for elaboration integration. The relatively least understood aspects were the
origin of the extremely sensitivity of the coagulation concentration
on the valence of the counterion and the distribution of potentials
and charges in the double layer, with the accompanying question
which potential to substitute in stability equations. During the war,
and shortly after it, Verwey and Overbeek [2] and Deryagin and
Landau [3], ignorant of each others work, developed their theories,
which settled the valence problem but not that of the potentials.
3. The aborted Faraday Discussion. Electrokinetics
3.1. General
About 15 contributions in [1] are dedicated to electrokinetics. Some are found under Experimental methods, some under
Streaming Effects and Surface Conduction and some under
Electrokinetic Equations, which, in turn, resides under Theoretical Treatment of the Double Layer. The methodical ones refer
to the classical ways of measuring electrophoretic mobilities of
hydrophilic sols, with some emphasis on biomacromolecules, such
as serum albumin. Given the theme of the present publication
references [2830] are the most relevant. The contribution by Hermans [29] anticipates more advanced theory for the electrophoretic
mobility in which the distortion of double layers around spherical
particles in an external eld is taken into account. As the theme for
his PhD thesis, this theory was developed by Overbeek during the
earlier part of the war (1941). At that time the Netherlands were

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

under German occupation, which allowed publication in a German


journal [31].
3.2. The surface conduction issue
Bikerman [30] makes the point that in many electrokinetic equations a surface conduction contribution must be incorporated and
that this correction may affect the way in which the observed quantity depends on the electrolyte concentration. By way of illustration,
in our present symbols, the classical HelmholtzSmoluchowski
equation for the streaming potential Estr in a capillary reads
Estr =

o p
K L

(3)

where is the dielectric constant of the liquid (assumed to be identical to that of the bulk), o the permittivity of free space, p the
pressure drop,  the (bulk) viscosity and KL the (bulk) conductivity.
When surface conduction K is included, the equation becomes
Estr =

o p
(K L + 2K  /a)

(4)

where a is the capillary radius. It is appreciated that the introduction of the surface conduction term can help to explain
(1) the fact that streaming potentials may depend on the radius of
the capillary and
(2) the sometimes anomalous dependence of the streaming
potential on the electrolyte concentration, because the bulk
conductivity and surface conductivity depend in different ways
on the salt concentration. These two points were made by Bikerman in [30]. It is obvious that incorporation of the surface
conductivity term is a relevant aspect of testing DLVO equations
if for the potential in their relevant equations  is substituted.
During the war this issue was not yet well understood and the
actual incorporation of K had to wait till reliable values were
available. In older papers Bikerman already proposed equations
for K based on the approximation that double layers are fully
diffuse. After some rearrangement and simplication regarding
the mobilities of anions and cations (see [32]) the equation can
be written as

K d =

4F 2 cz 2 D[1 + 3m/z 2 ]
cosh
RT

 zF 
2RT

(5)

The simplication is that the diffusion coefcients of cations and


anions are taken identical and equal to D, that the electrolyte is symmetrical (zz) and that the dimensionless molecular parameter m is
the same for cation and anion (it is about equal to 0.15). For further
information and a graph see [32]. It is obvious that the surface conductivity depends in a complicated way on the salt concentration
and the valence and it may be that this complexity has restrained
earlier investigators to apply this correction to real systems.
It was here that Rutgers contribution [28] must be mentioned,
because he made attempts at estimating the surface conductivity
of glass. He did so by making a virtue of necessity by carrying out
streaming potential experiments in glass capillaries of the same
material but with different radii. Values for K are obtained by
inserting that value that harmonizes the computed  potentials in
capillaries of different radius. This procedure is very laborious and
requires capillaries that are identical with respect to their inner
surfaces. However, thanks to painstaking experimentation Rutgers
did arrive at values that are in line with modern results. He even
presented them as functions of the electrolyte concentration. As at
that time the concepts of the charge-determining processes were
not yet fully crystallized he could not yet interpret his data on a
molecular scale. Typically, he presented data for KCl and HCl as the

165

electrolytes, not realizing that addition of these two electrolytes


had quite different consequences for the surface charge.
In summary, although at that time more research was needed to
quantitatively implement the consequences of surface conduction
on electrokinetic phenomena, it was established that -potentials
are dened characteristics of charged surfaces. Even if in 1939 the
molecular interpretation of the slip process was by no means established, there were convincing reasons to consider -potentials as
characteristics of the charge and potential distribution in double
layers. That was a very important conclusion because for many
types of colloids the surface charge and surface potential were
virtually unknown, whereas electrokinetic potentials were measurable.

4. Colloid stability theory during and after the war


4.1. Development of DLVO theory
The history of the development of the stability theory
for lyophobic colloids during the war, independently by
DeryaginLandau and VerweyOverbeek [2,3], is well enough
known. It was only several years after Verwey and Overbeek had
published their monograph [3] that they became aware of the older
publication of Deryagin and Landau [2]. I myself, in the early 1950s
being a staff member of Overbeek, remember that he made a note
with a paper published in the literature, interpretatie volgens
VO. It meant that at that time neither he, nor the author, was
informed about the progress made in the USSR during the war. The
reason for the fact that it took so long in Western Europe to realize
the existence of [2] and appreciate its contents may have had a
number or reasons. One of these was that the VerweyOverbeek
monograph [3] is very well written and readily available. In this
book, all the steps that had to be taken are well explained, and all
choices are consciously made. These impressions gave the book
an air of nality; it looked like the denitive word, leaving no
reason to look elsewhere. A second point was that the connections
between Western Europe and the Soviet Union after the World
War did not immediately improve; rather it was the beginning of
the Cold War period. A probable third point might have been that
Acta Physicochim. was not so easily accessible in the West and,
nally, paper [2] it is not as well written as [3]. Anyway, as to the
contents the two papers are to a large extent identical.
In 1955 the Faraday Division of the British Chemical Society
organized a follow-up of the aborted Discussion of 1939 [33]. This
meeting had the character of a reunion of the masters and it
was here that Deryagin met for the rst time Verwey and Overbeek. Deryagin was sensitive to the issue of priority and demanded
that Verwey and Overbeek wrote a formal paper in which this was
acknowledged. Eventually, Verwey and Overbeek did so in a brief
note in which they added that they developed their version of the
theory fully independently [34]. Independent proof of this can be
found in a Dutch text from July 1944, which did not make it to
abroad [35] but where Verwey already described the theory. He
stated that his theory is different from all other pre-war attempts,
except for that by Deryagin. Nowadays the acronym DLVO is generally accepted; it was probably proposed by Levine. As a tongue
in cheek anecdote, from that period stems a paper by the couple Robert and Marjory Vold, both colloid scientists at USC in Los
Angeles, in which they refer to the VOLD theory.
In passing it is interesting to digress briey into the meeting in
which Verweys paper [34] appeared because it also suffered from
the war conditions. This meeting took place in Utrecht, in July 1944,
a few weeks after the invasion in Normandy (June) and before the
so-called battle of Arnhem (Sept.), to the younger generation better known from the book and the movie A Bridge too Far. At that

166

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

time trafc in the Netherlands was already disrupted because of


the many bombardments. As a result, of the 200 registered participants only 80 showed up. One of the speakers, Kramers, could
not make it either. There was no way to print the proceedings;
the printed version had to wait till 1946. This booklet is still on
the typical war-quality paper, yellowish and fading in sunlight.
An additional war-aspect was that after the battle of Arnhem the
front line ran between the southern and middle northern parts of
the Netherlands. As a consequence, after September colloid scientists, working in the Philips lab in Eindhoven. (Verwey, Overbeek,
Casimir, etc.) lost all communication with the remainder part of the
Netherlands (including Kruyt). However, they could communicate
with scientists in Germany and have access to German periodicals
like Kolloid Z and Kolloid Beihefte.
4.2. The valency dependence of stability
Let us now return to the science by addressing the explanation
of the SchulzeHardy rule. Both teams were able to give a theoretical background for the very strong inuence of the valence of
the counterion and both arrived at the z6 power dependency for
the case of high potentials. According to [3] the critical coagulation
concentration is given by
ccr =

const.[tanh(ze
A2 z 6

o /kT )]4

(6)

Here A is the Hamaker constant and the constant in the equation


contains parameters like the temperature and dielectric constant of
the medium. For high potentials the hyperbolic tangent approaches
unity so that
ccr =

const 
A2 z 6

(7)

Almost the same equation is found in [2]; see Eq. (1) in their conclusion. The nding of the inverse sixth power law in the valence
dependence was, and is, THE achievement that [2,3] share. It was
seen as the denitive explanation of the SchulzeHardy rule, then
about half a century an open problem. This success was so convincing that the z6 law started to live its own life; for instance Matijevic
and his School, studying the adsorption of complex molecules on
particle surfaces, invoked this law to help establishing the valency,
and hence the composition, of these complexes.
Notwithstanding the deserved applause of the scientic community, the above result requires qualication. The rst question is
why (beyond the convenient truth that the result was so appealing)
the authors decided to consider the situation of high potentials? In
answering this question, V and O are more explicit than D and L.
In [3] the natural question was asked whether application to real
systems should run in terms of o or of , They argued that zeta
potentials were poorly understood and even differed between different investigators and different methods. Too much uncertainty
regarding the data would make a comparison with their theory virtually impossible. So, they settled for interaction determined by .
Having silver iodide sols in their minds, they knew that stability
studies were usually carried out when pAg differed from the p.z.c.
by more than 5 units, yielding absolute Nernst-type values for o
of >290 mV. That is enough to replace the hyperbolic tangent in (6)
by unity and obtain (7).
D and L are less concrete. In the rst place, they are confusing potentials and charges. Even in the title of [2] twice the term
strongly charged occurs, but their analysis emphasizes high potentials. However, a high charge does not yet mean a high potential.
That depends on the capacitance, or, for that matter, on the extent
of screening and on which potential you need: o , i , d or . The
most concrete remark that D and L make on this choice is that they
prefer the term apparent potential denoting the potential which

Fig. 1. GouyStern picture of the electrical double layer.

must be assigned to a plane surface in order that, while decreasing


exponentially with distance from the surface in accordance with the
exact, but approximate DebyeHckel equation and which asymptotically tends to coincide with the potential values which actually
exist in an electrolyte at a large distance from the surface. This is an
interesting statement because it anticipates current usage of quantitatively interpreting the exponential decay often observed in AFM
measurements.
Having thus argued the preference of both teams for the high
potential limit, the next question is whether this choice is correct.
4.3. Post-war DLVO theory
The most prominent feature of DLVO theory after the war is that
gradually the insight grew that not o but d is the proper potential
to be inserted in (6), with to a good approximation  d. For this
insight essentially two arguments can be given.
(1) The rst is the growing insight that only a, relatively small,
part of the outer double layer is diffuse. Fig. 1 illustrates the charge
distribution for the general situation in the absence of overcharging.
In this picture the outer Stern plane and the slip plane are, for
the sake of simplicity, assumed to be identical. That identication
is not an automatism because these two planes have very different origins and are determined by disparate forces. So, justication
is needed, see [35]. However, for quantitative purposes the exact
location of this layer does not matter so much because mostly the
diffuse part of the countercharge is only a small part of the total
countercharge, although the diffuse part plays a major part in particle interaction. The diffuse charge  d rarely exceeds 34 C/cm2
whereas for oxides  o can be as high as 3050 C/cm2 .
Historically it is interesting that D and L did not cite Sterns theory, whereas V and O already included a section on incorporation
of Stern corrections. See Ref. [3], Section VII.5. They realized that
in (6) after all the substitution of d for o would be appropriate and that it would allow interpretation of lyotropic (Hofmeister)
sequences because the potential drop over the Stern layer depends
on the extent of screening of the surface charge by bound counterions. However, at that time they found the measurements of
electrokinetic potentials too uncertain to go one step further and
elaborate systematically the consequences of substituting  for the
potential in (6).
(2) The second argument stems from the fact that in the years
after the war electrokinetic potentials became gradually better
dened, partly because of better theories, including double layer
polarization of particles upon electromigration, theory initiated by
Overbeek himself, but also because the work of Rutgers and his

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

successors showed how incorporation of surface conduction could


quantitatively obviate the seemingly anomalous inuences of concentration and capillary radius.
Implementation of all these improvements took of course its
time although the steps are straightforward and simple. Before
describing these let us briey discuss the mode of interaction
between colloidal particles.

to rephrase the title question of Section 4.3 into: interaction at


constant d  or at constant  d  ek?
Quantitatively speaking, the difference between interaction at
constant potential and at constant charge is smaller than that
between the high and low potential limit of (6). Therefore we shall
rst consider the former.
ccr =

4.4. Interaction at constant potential or at constant charge?


Before proceeding with the implementation, it is useful to say a
few words about this difference. D and L do not discuss this issue,
but V and O do so in some detail. Consider for the moment the
premise of a purely diffuse double layer. The mathematic elaboration of the Gibbs energy change with distance requires as a
boundary condition a reply to the question of what happens with  o
and o upon particle approach. Both D and L, and V and O, carried
out the integration at xed o . As a result, this parameter appears in
the equation for the critical coagulation concentration (6). Unlike
the Russians, V and O explain why that choice was made. Their
argument goes as follows.
Double layer formation is partly of a chemical and partly of
an electric (physical) origin. Spontaneous adsorption of chargedetermining ions must have a chemical origin. No net charge can
be built up without a chemical afnity of ions for the surface that is
high enough to overcome the ensuing electrostatic repulsion. The
driving force for the formation of such double layers has therefore
a chemical origin. The surface potential is related to the properties
of the solution through Nernsts law, meaning that for silver iodide
and oxides o is xed by pAg and pH, respectively. Assuming that
upon interaction the adsorption equilibrium persists, and that these
parameters do not change upon particle approach, it is necessary
that charge-determining ions desorb from the surface in order to
keep o constant. The (repulsive) Gibbs energy upon interaction of
particles of the same charge has therefore a chemical origin. This is
a logical presumption for reversible double layers like those on AgI
and oxides.
On the other hand, interaction at constant charge can take place
if no charges can desorb or adsorb upon particle encounter. This
situation prevails for example for clay particles where the charge is
caused by isomorphic substitution. In this case the Gibbs energy of
interaction is purely electrical. Upon particle approach  o remains
constant but o increases. In this case equations for the coagulation
concentration contain the surface charge.
These considerations have two interesting consequences. The
rst is that the colloid interaction phenomenon acquires a dynamic
nature. Whether or not ions can desorb upon interaction also
depends on the time needed to desorb relative to the time of interaction. We shall not discuss that here, except for referring to [36].
The second consequence is that in real systems most of the
countercharge is bound in the stagnant layer, so that reduction
of the charge upon particle interaction is not necessarily achieved
by desorption of charge-determining charges but can as easily be
obtained by adsorption of more counterions. How in practice this
works out depends on the relative rates of these processes, as compared to the rate of particle approach. Otherwise stated, that again
is a question of interaction dynamics.

167

const   4
A2 z 2

(8)

Here  and d are considered identical. In this limit the concentration dependency of the potential is very strong. For systems in
which stability is studied as a function of  it is indeed found that
there is something like a critical zeta potential, of the order of
2030 mV, depending on the material (through A). On the other
hand, the valence dependence is much weaker than in (7).
In applying (7) or (8) to real systems it must be realized that
(8) does not automatically predict the coagulation concentration to
depend inversely quadratically upon the valency, because  itself is
a decreasing function of the concentration, so that the decrease is
stronger than quadratically and depends on the counterion specicity. The question is then what is observed in practice?
Both D and L and V and O paid attention to the valence dependence in real systems. The former couple referred to older data by
Ostwald (1935). V and O referred to older data by Freundlich (1910,
1912) but they also had more detailed data at their disposal, collected by Klomp during the war and published in German and
in a German journal in 1942 [37]. Later, Overbeek repeated and
extended these data in [38]. Klomp was a dedicated PhD student
of Kruyt, Overbeeks predecessor at Utrecht University. Later she
became the rst female minister in one of the Dutch governments.
For our purpose, her detailed study of coagulation concentrations
for AgI sols has withstood ageing of over more than half a century
and still serves our purpose. For strongly negatively charged sols
the mean values are 142, 2.43, 0.068 and 0.013 mM for counterion
valencies 1, 2, 3 and 4, respectively. Their ratios are 1:0.017:0.0005;
0.0001. If the z6 rule would apply, these ratios ought to be 1:0.
016:0.0013 and 0.00024, respectively. So, between the monovalent
and divalent counterions the high potential law applies perfectly,
but upon increasing z it fails progressively. Nowadays we know
why: for higher z the tendency of the ions to hydrolyze increases.
As the hydrolysis products adsorb strongly, they lower d [39]. All
of this is in fact corroborating DLVO theory provided it is applied
under conditions for which it has been derived.
The story can be further qualied by noting that at given z clear
lyotropic (or Hofmeister) trends are observed. In particular, for
the monovalent counterions the trends are evident, for example
the coagulation concentrations for LiNO3 and RbNO3 are 165 and
126 mM, respectively. This is a substantial specic difference that
comes on top of the DLVO high potential case and it means that
specic adsorption in the Stern layer cannot generally be neglected.
This consideration automatically invites the question of how
to quantify it. With respect to the establishment of the potential
responsible for interaction, this question can be rephrased as how
to optimally specify -potentials. This takes us to the question of
how to incorporate surface conductivity into electrokinetic equations progressed after the war.
5. Surface conductivity during and after the war

4.5. Interaction at low potentials


5.1. Data acquisition
Retracing the historical development, the gradual improvement
in the measurements of -potentials and their functioning in interaction has given rise to revisit the assumption of a high potential
for the expression of the critical coagulation concentration, as
embodied by (7). Particularly under conditions near to instability the low-potential limit is preferred. At the same time one has

Surface conductivity is an excess quantity; it tells by how much


the conductivity is enhanced by the presence of ion excesses in
double layers. In the analysis of some measuring techniques surface
conductivity must be explicitly taken into account, but then the
additionally obtainable information can be enlightening.

168

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169

As to methods for determining K three options with increasing


power can be mentioned.
The rst, and most indirect one, is by comparison of  obtained
by different electrokinetic techniques, one requiring accounting for
surface conduction (like streaming potentials) and one which does
not need that (for example, streaming currents) and then nd by
trial and error for which value of K the resulting electrokinetic
potential is the same.
The second, more systematic, approach goes back to the work
by Gortner and Rutgers, already mentioned above, where K was
more directly measured by using (4) to systematically study the
radius dependence of the streaming potential. The method works
well but is rather time-consuming and it is limited to materials
of which capillaries of well-dened and homogeneous inner radii
could be drawn. Most of the results by Rutgers and his School refer
to glass.
A more general and more foolproof idea was launched recently
by Minor et al. [see 40, 41 for details]. It rests on the fact that the
bulk conductivity and the surface conductivity both depend on the
indifferent electrolyte concentration in bulk. The procedure can be
applied to any material of which plugs can be made. Plugs are percolated by salt solutions and at each salt concentration the plug
conductivity, per unit area, K (plug) is plotted as a function of the
bulk conductivity, KL .The resulting graphs are linear except in the
initial parts. This linearity results from lling of the inner lumen of
the capillaries between the particles with electrolyte of the same
composition as the bulk. Extrapolation to zero bulk concentration
always results in a non-zero intercept. Apparently this reects the
conductivity that is left once the conductivity in the bulk of the
pores is zero; hence it can be attributed to surface conduction. The
procedure allows establishment of K .
5.2. Data and interpretation
Shortly after the war, Overbeek decided to review the best surface conductivity data of that time and check those against the best
theories of that time [42, especially his Table 3]. The best data of his
time stemmed from Rutgers and his group, in addition to older data
by other investigators. The best theory then available was that by
Bikerman, which for simple, but representative conditions leads to
our Eq. (5). Overbeeks conclusion was that most experimental data
exceed Bikermans theoretical predictions by large amounts, even
sometimes by orders of magnitude.
This last disparity persisted till recently when many more data
on surface conduction became available with the Minor technique
and some variants. The modern interpretation is that the origin of
this discrepancy between theory and experiment is not in the data
but in the fact that Bikerman only considers conduction in the diffuse part of the double layer, completely ignoring contributions of
the Stern layer. This insight required a change in the thinking. The
success of PoissonBoltzmann theory was so convincing that investigators tended to see the Stern layer only as a (minor) correction,
whereas with respect to the countercharge in the solution side of
the double layer it can contain it as much as 75%. Even if these
Stern ions would be only half as mobile as those in the diffuse part
of the double layer, the contribution Ki of the inner layer conduction to K is at least comparable to that of Kd . The thinking step
that had to be taken was that even if the inner layer is hydrodynamically stagnant, the ions in it may be electrically mobile. In fact, this
is the situation with many gels, where the liquid is immobilized
but where, processes like diffusion or conduction by solutes is not
much slower than that in the mobile bulk. Historically it is interesting to note that Overbeek [42] stated that Urban et al. [43] added
a contribution to (5) caused by the ions of the Stern layer where the
normal mobility and the normal contribution to the electro-osmosis
of the Stern ions is supposed. Overbeek rejected this assumption,

stating that it does not sound very probable, although it could not be
denied that it worked well in explaining the results of [43].
Nowadays it is commonly accepted that Stern ions are tangentially mobile. In the favourable situation that surface charge
and -potentials are simultaneously experimentally available it is
possible to compute the mobility of the Stern ions by a relatively
straightforward procedure. First, from the -potential the electrokinetic charge is computed using diffuse double layer theory. This, in
turn, is identied as the diffuse charge  d . Subtraction from the
surface charge  o yields the inner charge  i . By the same token,
from the experimental total surface conductivity K the diffuse
(Bikerman) part is subtracted, yielding Ki . Knowing the total number of charges and the total conductivity of the Stern layer, the
average mobility per ion can be computed [44,45]. The outcome is
that monovalent cations move tangentionally almost as fast as in
bulk, bivalent cations at roughly 2/3 of their bulk values and trivalent cations slower than that. These data can be brushed up when
more recent data become available. A possible interpretation of this
valence effect is that for tangentional motion ions hop over the surface from site to site with an activation energy increasing with the
valence.
The general conclusion is that also the development of surface
conduction has not only assisted in improving stability theory but
that it also gave rise to much new physical insight.
6. Future developments
The cross-fertilization of stability and surface conduction studies has over the years given rise to much increased insight and this
development has not yet nished. Better understanding of surface
conduction was protable in two respects: rst it helped increasing
the understanding of the valence effect on stability and second, it
has lead to the interesting intrinsic result regarding the mobility of
ions in the Stern layer.
It is likely that the obtained insight will pay rich rewards in the
interpretation of interaction studies, for instance by AFM and similar techniques, or in phase formation of charged systems. Finding
replies to recurring questions regarding the mode of interaction
(constant charge or constant potential?), charge and/or potential regulation, specic ionic effects and dynamic studies (to what
extent relaxes a double layer upon interaction?) can be facilitated
by the insights presented in this paper.
References
[1] Aut. Div., The Electrical Double Layer, collection of papers offered for discussion at an aborted Faraday Discussion, Trans. Faraday Soc. 30 (1940) 1322.
[2] B.V. Derjaguin, L. Landau, Theory of the stability of strongly charged lyophobic
sols and of the adhesion of strongly charged particles in solutions of electrolytes, Acta Physicochim. 14 (1941) 633661.
[3] E.J.W. Verwey, J. Th.G. Overbeek, Theory of the Stability of Lyophobic Colloids,
Elsevier, Amsterdam, New York, 1948, Also available as a Dover reprint (1999).
[4] H. Schulze, Schwefelarsen im wssriger Lsung, J. Prakt. Chem. 25 (1882) 431;
H. Schulze, Schwefelarsen im wssriger Lsung, J. Prakt. Chem. 27 (1883) 320.
[5] W.B. Hardy, A preliminary investigation of the conditions which determine the
stability of irreversible hydrosols, Proc. Roy. Soc. 66 (1900) 110-125, see also Z.
Phys. Chem. 33 (1900) 3051.
[6] B. Vincent, Early (pre-DLVO) studies of particle aggregation, Adv. Colloid Interface Sci. 170 (2012) 5667.
[7] S. Wall, The early history of electrokinetic phenomena, Curr. Opin. Colloid Interface Sci. 15 (2010) 119125.
[8] See or instance J. Lyklema, Interfacial potentials in Fundamentals of Interface
and Colloid Science (FICS), vol. I, sec. 5.5, Academic Press, London, San Diego,
New York, Boston, Sydney, Tokyo, Toronto, 1993, 2nd print.
[9] E.J.W. Verwey, Ph.D. thesis, State University of Utrecht, 1934.
[10] J.H.A. Pieper, D.A. de Vooys, Direct measurements of the double layer capacity
at the AgI-aqueous solution interface, J. Electroanal. Chem. 53 (1974) 243252.
[11] B.H. Bijsterbosch, J. Lyklema, Interfacial electrochemistry of silver iodide, Adv.
Colloid Interface Sci. 9 (1978) 147251.
[12] S.R. Craxford, On the electrochemistry of simple interphases, with special reference to that between mercury and solutions of electrolytes, Trans. Faraday
Soc. 35 (1940) 85101.

J. Lyklema / Colloids and Surfaces A: Physicochem. Eng. Aspects 440 (2014) 161169
[13] A.N. Frumkin, The study of the double layer at the metal-solution interface
by electrochemical and electrokinetic methods, Trans. Faraday Soc. 35 (1940)
117127.
[14] I.M. Barclay, J.A.V. Butler, Some observations on the double layer capacity at
mercury electrodes, Trans. Faraday Soc. 35 (1940) 128136.
[15] H.R. Kruyt, J. Th.G. Overbeek, Theoretical treatment of the double layer and its
implications: introductory paper, Trans. Faraday Soc. 35 (1940) 110116.
[16] Aut. Div., The Electrical Double Layer, collection of papers offered for discussion at an aborted Faraday Discussion, Trans. Faraday Soc. 30 (1940) 1322
(errata p. 322).
[17] H.C. Hamaker, E.J.W. Verwey, The role of the forces between particles and electrodeposition and other phenomena, Trans. Faraday Soc. 35 (1940) 180185.
[18] H.C. Hamaker, The inuence of particle size on the physical behaviour of colloidal systems, Trans. Faraday Soc. 35 (1940) 186192.
[19] E.J.W. Verwey, Electrical double layer and stability of emulsions, Trans. Faraday
Soc. 35 (1940) 192203.
[20] B. Derjaguin, On the repulsive forces between charged colloid particles and on
the theory of slow coagulation and stability of lyophobic sols, Trans. Faraday
Soc. 35 (1940) 203215.
[21] S. Levine, G.P. Dube, Stability properties in hydrophobic sols: application of the
mutual energy of two particles, Trans. Faraday Soc. 35 (1940) 215229.
[22] H. Eilers, J. Korff, The signicance of the electrical charge on the stability of
hydrophobic dispersions, Trans. Faraday Soc. 35 (1940) 229241.
[23] D.F. Cheesman, A. King, The electrical double layer in relation to the stabilisation
of emulsions with electrolytes, Trans. Faraday Soc. 35 (1940) 241247.
[24] H.C. Hamaker, The London-van der Waals attraction between spherical particles, Physica 4 (1937) 10581072.
[25] J. Lyklema, The bottom size of colloids, Bull. Polish. Acad. Tech. 53 (2005)
317324.
[26] B. Derjaguin, Theorie ber die Reibung und Adhsion, IV. Theorie das Anhaftens
kleiner Teilchen, Kolloid Z. 69 (1934) 155164.
[27] N. Fuchs, On the stability and charging of aerosol, Z. Physik 89 (1934) 736743.
[28] A.J. Rutgers, Streaming potentials and surface conductance, Trans. Faraday Soc.
35 (1940) 6980.
[29] J.J. Hermans, Relaxation effects in the double layer. Cataphoresis; dielectric
constant, Trans. Faraday Soc. 35 (1940) 133139.

169

[30] J.J. Bikerman, Electrokinetic equations and surface conductance. Survey of the
diffuse double layer theory of colloidal solutions, Trans. Faraday Soc. 35 (1940)
154160.
[31] J.Th.G. Overbeek, Theory of the relaxation effect in electrophoresis, Kolloid
Beihefte 54 (1943) 287364.
[32] J. Lyklema, Fundamentals of Interface and Colloid Science, vol. II, Academic
Press, London, San Diego, New York, Boston, Sydney, Tokyo, Toronto, 1995 (sec.
4.3f).
[33] Aut. Div., Coagulation and occulation, Discuss. Faraday Soc. 18 (1955) (in
Shefeld, UK).
[34] E.J.W. Verwey, De Electrochemsche Dubbellaag in Colloidale Systemen, in:
Tweede Symposium over Sterke Electrolyten en de Electrische Dubbellaag,
Dutch Chemical Society, The Hague, 1944.
[35] J. Lyklema, Molecular interpretation of electrokinetic potentials, Curr. Opin.
Colloid Interface Sci. 15 (2010) 125130.
[36] J. Lyklema, Colloid stability as a dynamic phenomenon, Pure Appl. Chem. 52
(1980) 12211227.
[37] H.R. Kruyt, M.A.M. Klomp, Solkonzentration und Flockung beim AgJ-Sol, Kolloid Beihefte 54 (1942) 484553.
[38] J.Th. G. Overbeek, Stability of hydrophobic colloids and emulsions, in: H.R. Kruyt
(Ed.), Colloid Science, vol. I (Chapter VIII in particular p. 307 ff).
[39] J. Lyklema, T. Golub, Electrical double layer on silver iodide and overcharging
in the presence of hydrolyzable cations, Croat. Chem. Acta 80 (2007) 303311.
[40] J. Lyklema, M. Minor, On surface conduction and its role in electrokinetics,
Colloids Surf. A 140 (1998) 3341.
[41] M. Minor, A.J. van der Linde, J. Lyklema, Streaming potentials an conductivities
of latex plugs in indifferent electrolytes, J. Colloid Interface Sci. 203 (1998)
177188.
[42] J.Th. G. Overbeek, On surface conductance, in: H.R. Kruyt (Ed.), Colloid Science,
vol. I, Elsevier, Amsterdam, Houston, New York, London, 1952 (Chapter 5, sec.
12).
[43] F. Urban, H.L. White, E.A. Stassner, Contribution to the theory of surface conductivity at solidliquid interfaces, J. Phys. Chem. 39 (1935) 311330.
[44] J. Lyklema, Tangential motion of uids near charged surfaces: consequences
for wetting, Oil Gas Sci. Tech. Rev. IFP 56 (2001) 4145.
[45] J. Lyklema, Surface conduction, J. Phys. Condens. Matter 13 (2001) 50275034.

You might also like