Vous êtes sur la page 1sur 7

Current Opinion in Colloid & Interface Science 9 (2004) 178 184

www.elsevier.com/locate/cocis

Review

Bubble coalescence and specific-ion effects


Vincent S.J. Craig *
Department of Applied Mathematics, Research School of Physical Sciences and Engineering, Australian National University,
Canberra ACT 0200, Australia

Abstract
Major recent advances: Recent advances that contribute to our understanding of specific-ion effects in bubble coalescence include new
theoretical and simulation efforts to determine the arrangement of ions at interfaces and a clearer recognition that specific-ion effects in
bubble coalescence are related to many other phenomena that exhibit ion specificity.
D 2004 Elsevier Ltd. All rights reserved.
Keywords: Bubble coalescence; Specific-ion effects; Electrolytes; Thin film stability; Hydrophobic; Hofmeister; Nanobubbles

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . .
2. The importance of ion specificity . . . . . . . . . . . . .
3. Bubble bubble coalescence in the presence of surfactant .
4. Films between two bubbles in the absence of surfactant .
5. Films formed between bubbles and solid surfaces . . . . .
6. Specific-ion effects and bubble stability . . . . . . . . . .
7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . .
References. . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

178
178
179
180
180
182
183
183
183

1. Introduction

2. The importance of ion specificity

Perhaps, one of the most surprising and easily observed examples of a macroscopic effect of ion specificity
is the influence that dissolved electrolytes have on bubble
coalescence in aqueous solutions. Above a certain concentration, which is dependent on the particular salt
chosen (typically f 0.1 M), many common electrolytes
inhibit the coalescence of bubbles, whereas others have no
influence on bubble coalescence [1,2..]. This simple
observation is not understood. Here, we will attempt to
determine if the ions influence coalescence by action at
the interface or in bulk and investigate their possible
influences on film rupture.

The Derjaguin, Landau, Verwey and Overbeek (DLVO)


theory of colloidal interactions has been the dominant
paradigm for 50 years. This theory generally works well
under the circumstances for which it was intended, low salt,
inert surfaces and interactions at separations greater than a
few nanometers. Under these conditions, the charge on an
ion, rather than the particular type of ion, is important.
Therefore, the theory explicitly excludes specific-ion effects
and can largely be seen as a chemistry-free regime.
However, since the work of Hofmeister in the 19th century,
it has been known that the nature of an ion can have a very
great influence on the stability of a colloidal dispersion,
particularly at high salt concentrations. At concentrations
exceeding 0.1 M, the range of the electrostatic interactions is
greatly reduced, and specific-ion effects dominate. Therefore, interactions in biological systems [3.], as well as
complex fluids [4] and slurries [5], often cannot be under-

* Tel.: +61-2-6125-3359; fax: +61-2-6125-0732.


E-mail address: vince.craig@anu.edu.au (V.S.J. Craig).
1359-0294/$ - see front matter D 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.cocis.2004.06.002

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

stood within the DLVO framework, but rather, the specific


nature of the ions is important. Our understanding of ion
specificity has barely progressed from the empirical work of
Hofmeister more than 100 years ago, with the terms ion
size and polarisability giving names to our ignorance, in
the absence of an ion-specific theory. Without theoretical
predictions, molecular recognition, drug design, formulations science and protein crystallization will remain largely
empirical arts.
Whilst there are a great number of systems that
exhibit ion specificity, attempts to investigate these
effects are challenging. Currently, short-range interactions
cannot be adequately measured for most systems, particularly when soft surfaces are involved. Therefore,
efforts necessarily concentrate on indirect methods. Studies in protein solubility, complex fluids, enzyme action,
surface tension, pH and bubble coalescence have provided a great deal of evidence against which theories
can be tested. The challenges for theoreticians are
extensive, requiring recognition of the granularity and
chemical nature of the solvent, the geometry and nature
of the surfaces, inherent nonlinearity and an appropriate
description of the ion, which may include ion size,
polarisability, solvent interactions and ion geometry,
amongst other descriptors. Simulation studies must contend with all this and find means to deal effectively
with a range of length scales that extend from the
classical down to the quantum regime. Underlying all
this is a belief that a theory of ion specificity should be
able to describe all of the very diverse specific-ion
effects, such as Hofmeister, anti-Hofmeister behavior,
strong co-ion influences and competition between ions.
Therefore, a theoretical description that is adequate in one
system may lead to an increased understanding of other
systems, even if they do not exhibit the same ion-specific
effects. If this is true, the apparently simple system of
two bubbles colliding in an aqueous electrolyte solution
may be useful in elucidating the nature of short-range
interactions.
Over a decade ago, we described simple combining rules
that predict the bubble coalescence behavior of electrolytes
based on assigned properties of the ions that make up the
electrolytes [1,2..]. Ions were empirically assigned a type, a
or h and the combination of ions present in the added
electrolyte determined, if the solution would inhibit bubble
coalescence. Bubble coalescence was inhibited relative to
pure water with an aa or hh combination and unchanged
when an ah or ha combination was employed (see Fig. 1).
As yet, no exceptions have been found to these rules,
although it has been claimed that NaClO4 inhibits bubble
coalescence at very high concentrations [6]. However, at
such high concentrations, the level of contaminant ions is
likely to be sufficient to produce coalescence inhibition
independently.
What does the existence of these combining rules indicate? Primarily, it enables us to separate the ions into four

179

classes; a anions, h anions, a cations and h cations. We


should expect a fundamental difference between a and h
anions and, similarly, between the a and h cations. However, the labeling does not necessarily mean that a cations
and a anions or h anions and h cations share a commonality, although this is possible. The combining rules tell us
that it is not the absolute behavior of a single ion that
matters but the combination of ions. In my opinion, this is
an important and significant point that has often been
overlooked. The behavior is inherently nonlinear and specific to both ions present. Therefore, any suitable description
must identify a property or properties that separates the a
from the h ions and describe a means by which this property
or properties, in combination, can switch the coalescence
inhibition on or off. Ideally, in time, our ignorance will pass,
and the terms a and h will be replaced by terms that
describe the appropriate property. Clearly, the combining
rules provide a very strong test of any theory.

3. Bubble bubble coalescence in the presence of


surfactant
At this stage, we still do not have a clear understanding
of the influence of electrolytes on bubble coalescence.
Indeed, our ignorance is deeper than it first appears, as we
do not fully understand the coalescence process that takes
place when two air bubbles collide in pure water. In the
following, I will examine what is known of this process.
During the collision of two bubbles, a film of liquid is
formed between the two gaseous phases. The propensity of
the film to drain and rupture ultimately determines the
outcome of the collision process, and whilst it is obvious
that these two processes are not independent, they are often
treated separately for convenience, with the acknowledgment that once the film has drained below a certain
thickness, then film rupture is possible, and at another
thinner film dimension, film rupture is often assured.
Drainage is often treated within the Reynolds model of
two circular parallel plates separated by a fluid, although
under some conditions, it is known that dimpling of the
interfaces occurs [7. 9..]. Additionally, the influence of
interfacial forces must be taken into account.
The case of a thin film stabilized by a surface active
agent is relatively well understood. For films that are
electrostatically stabilized, the double-layer repulsive surface forces act to provide a disjoining pressure that slows or
prevents film drainage. The film thickness increases with
the range and magnitude of the repulsive force, as does its
stability to rupture. Addition of electrolyte screens the
repulsive force and allows a thinner film to form. As
drainage proceeds, the film becomes sufficiently thin, such
that the rupture of the film becomes likely. Thermal disturbances cause surface waves that can grow under the
action of attractive Van der Waals forces. If these waves
grow sufficiently, the liquid film will be pinched off, and

180

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

Fig. 1. Table depicting the effect of a range of electrolytes on bubble coalescence. A tick indicates that the salt inhibits bubble coalescence. A cross indicates
that no bubble coalescence inhibition has been observed. The properties a and h have been assigned empirically and can be used with the combining rules to
predict the bubble coalescence behavior of an anion cation combination. aa or hh combinations correspond to bubble coalescence inhibition and ah or ha
combinations, no inhibition. No exceptions have been found.

film rupture will occur [10]. The growth of surface waves


can be responsible for film rupture at thicknesses exceeding
100 nm given sufficient time to develop [11.]. For surfactant
films, rupture times exceed 10 s and often exceed 100 s, and
agreement with theory is not complete but is generally good,
supporting a capillary wave rupture mechanism [11.,12].
Films may also rupture by a nucleation mechanism [8..,9..],
where a hole in the film forms spontaneously. The likelihood of nucleation rupture is much greater as the film thins,
hence, film rupture by this mechanism is also closely related
to film drainage.

4. Films between two bubbles in the absence of


surfactant
It is commonly observed that even mildly stable films
cannot be formed in pure water (or any pure fluid); however,
the addition of certain salts in sufficient amounts permits a
metastable aqueous film to be formed. The increased stability of the film is sufficient to prevent the coalescence of
bubbles that collide in bulk or are brought together during
the production process at a frit. Cain and Lee [13..]
investigated the coalescence of captive bubbles in aqueous
KCl solutions and found an increase in film stability with
increasing salt concentration. As the electrolyte concentration increased, the film lifetime increased, but in all cases,
was less than 1 s. Films could not be produced in 0.1 M
KCl; films formed in 0.5 M KCl were found to rupture at
separations of between 75 and 95 nm, and in 1.0 M KCl,
films drained to thicknesses of between 55 and 75 nm before
rupture. In surfactant films, a decrease in the thickness of a
stable film is observed as the electrical double layer is
compressed upon the addition of electrolyte. There are
important differences when surfactant is absent, the film
thickness is not related to the Debye length of the solution

and the rate of film thinning has not been arrested at the
point of rupture. Clearly, the description of film rupture by
capillary waves applied to films formed in the presence of
surfactants cannot be applied here. The lifetimes of these
films is shorter than the characteristic time required for the
Van der Waals forces to rupture the films via capillary waves
[11.]. A description of film rupture in these cases therefore
requires a modification to the capillary wave approach or an
alternative means of film rupture, such as nucleation of
holes.
The absence of surface-active material at the interface
results in an increase in surface mobility. This will lead to a
considerable increase in the drainage rate and film thinning
times comparable with the film lifetimes measured in clean
systems [14.,15.], although the mobility of the interfaces,
important in these calculations, is not truly known. Capillary
waves will also grow more rapidly if the attraction between
the gaseous phases is increased, such as in the presence of a
long-range hydrophobic attraction.

5. Films formed between bubbles and solid surfaces


Studies of bubbles approaching solid surfaces can be
categorized depending on the wettability of the solid material [7.,16. 18]. When the material is hydrophilic, a stable
film is formed, and the film thickness is well described by
DLVO theory. An increase in salt concentration results in a
decrease in equilibrium film thickness, and this thickness is
directly related to the Debye length. However, when a
hydrophobic surface is employed, the film ruptures rapidly
and at large separations [16.,18] (typically >100 nm), and
the addition of salt prolongs the stability of the film, such
that the drainage process proceeds further and rupture
occurs at smaller separations [16.,19..]. This is true even
if the hydrophobic surface bears the same surface charge as

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

the hydrophilic system [19..] does and therefore cannot be


explained by a reduction in electrical double layer repulsion.
The degree of surface hydrophobicity need not be great;
generally, a contact angle in excess of 45j is sufficient to
lead to an unstable film [8..]. Clearly the hydrophobicity of
the surface plays an important role in the film stability,
although a direct correlation with the contact angle has not
been shown [8..].
Films formed between a bubble and a hydrophilic solid
surface act similarly to films formed between two bubbles
formed from surfactant solution. Their behavior is well
understood in both cases. Films formed between a bubble
and a hydrophobic solid surface are unstable, unless
electrolyte is added, as is the case for films formed
between two bubbles in the absence of surface-active
material. In this case, our understanding is poor, but
given the similarity of behavior, it is reasonable to suggest
that the same destabilisation mechanism is operating in
both cases. The evidence suggests that the hydrophobicity
of the substrate is important to film rupture. Previously
[1,2..], we have proposed that an uncontaminated air
water interface is a strongly hydrophobic surface, and this
has been supported in later studies [20]. Therefore, it is
reasonable to assume that a film separating two bubbles
will act similarly to a film confined between a bubble and
a hydrophobic solid surface.
Direct force measurements between hydrophobic surfaces are often characterized by a strong, long-range attractive
interaction. The range of the measured force varies considerably and appears to be strongly dependent upon the
surface preparation method employed rather than the contact
angle at the surface. This suggests that the hydrophobicity of
the interface is not the determining factor in the magnitude
of the interaction. Many theories have been proposed to
explain the interaction, but none are completely satisfactory
in terms of explaining its range and magnitude. An extensive
review by Christenson [21.] is recommended to the interested reader.
In many systems where a hydrophobic attraction is
measured, it can be argued that nanobubbles are present
on the surface and that these nanobubbles play an important
role in the attraction and often give a false impression of the
range of the attraction. The separation is determined experimentally as the solid solid separation distance, when, in
fact, the solid nanobubble or nanobubble nanobubble separation may be considerably less, depending on the size of
the nanobubbles. There is clear evidence of the existence of
nanobubbles on some hydrophobic surfaces [22 24]. In the
opinion of the author, many published measurements can be
attributed to the presence of nanobubbles, but there exist
several measurements between hydrophobic surfaces that
exhibit a hydrophobic attraction in circumstances where
nanobubbles are absent. For example, Ishida et al. [23.]
measured an attraction larger than is expected based on Van
der Waals forces, between hydrophobic surfaces that had
never been exposed to air and where the absence of nano-

181

bubbles was assured by AFM examination. In the absence


of nanobubbles, the long-range attraction measured can be
considered the true long-range hydrophobic attraction.
Mahnke et al. [8..,9..] have elegantly demonstrated that
the film separating a bubble and a hydrophobic surface
ruptures at separations exceeding 40 nm in the absence of
nanobubbles. They attribute this to the action of hydrophobic forces. There is also evidence that the drainage rate is
faster on a hydrophobic surface [7.], suggesting that an
additional attraction may be present, providing further
evidence for a hydrophobic attraction between a bubble
and a hydrophobic surface.
Some time ago, we suggested that the hydrophobic
attraction may be operating between colliding bubbles to
produce coalescence [1,2..]. Let us examine this in more
detail. If a hydrophobic attraction (of unclear origin) that is
considerably larger than the Van der Waals interaction is
acting between bubble surfaces, this may play a role in the
destabilization of the aqueous film. First, we will consider
the capillary wave mechanism. Calculations reveal [11.] that
the range at which films rupture is consistent with capillary
waves, but the measured time of rupture is not sufficient if
the interfaces are considered immobile. However, the presence of an attractive force much larger in magnitude and
range than the Van der Waals interaction will lead to more
rapid growth of capillary waves and a large decrease in
rupture time. Therefore, the capillary wave mechanism of
film rupture driven by the hydrophobic attraction may be
valid in the pure water system.
Let us consider nucleation as a rupture mechanism.
The nucleation of a vapor phase between two nonwetting
surfaces at separations below 500 nm is energetically
favourable, and in the presence of dissolved gas, the
range at which nucleation is possible is extended considerably [25.]. However, there is a large activation barrier
to nucleation of a gas phase between hydrophobic surfaces, unless the surfaces are at a very small separation.
When the hydrophobic surface is itself a bubble, the gas
within the bubble is able to enter the film, and this
should reduce the barrier to nucleation. The amount of
gas entering the liquid film could be increased by a
strongly attractive force; therefore, it is feasible that the
hydrophobic attraction could increase the range and
probability of film rupture through nucleation. Furthermore, the origin of the hydrophobic attraction may be
related to the metastability of the intervening film, but
until the mechanism of the attraction is understood, any
relationship is speculative.
If one accepts a role for the long-range hydrophobic
attraction in bubble coalescence in aqueous systems, a
similar force must be proposed in other pure liquids, as
stable films cannot be produced in any pure liquid. Measurements have revealed a similar, long-range interaction in
other liquids [26]; however, these forces were attributed to
the presence of nanobubbles, and other liquids studied did
not exhibit a long-range attraction.

182

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

In the absence of solute, Gibbs Elasticity will be absent


and Marangoni effects minimised. One could therefore
argue that capillary waves could grow undisturbed or
mechanical disturbances could have catastrophic effects on
film stability; however, the considerable film stability observed between bubbles and hydrophilic substrates in pure
water are strong evidence that this is not the case. We have
already indicated that if the interface is mobile, film thinning
progresses more rapidly and this can lead to short rupture
times. A highly mobile bubble aqueous interface could be
invoked to explain the short rupture times of films between
a bubble and a hydrophobic surface. However, films formed
between a bubble and a hydrophilic surface are stable for
long periods. In both systems, the bubble aqueous interface
is identical and equally mobile, hence, this is clearly not the
dominant influence. Therefore, the rupture of films between
a bubble and solid surface or between two bubbles in pure
water remains unresolved.

6. Specific-ion effects and bubble stability


Any discussion of the role of electrolytes in bubble
stability is going to be speculative given our lack of
understanding of the coalescence process in pure water.
Electrolytes have only a little [27] influence on bubble
coalescence up to the transition concentration (typically
f 0.1 M for a 1:1 salt). Therefore, the electrostatic
double layer is highly compressed at the concentrations
that concern us. The challenge then is to resolve how
the short-range effect of electrolyte can influence the
bubble coalescence process that occurs at separations of
approximately 100 nm. We must determine if the influence of electrolyte is at the interface or if it is a bulk
phenomenon.
In the past, we have proposed that electrolyte may reduce
the range of the hydrophobic attraction between two bubbles
sufficiently to prevent film instability [1,2..]. However,
subsequent measurements of the hydrophobic attraction in
concentrated salt solutions showed that the attraction is not
reduced by the addition of salt [28]. Therefore, we can
abandon this hypothesis. Two other possibilities will be
investigated here: the possible damping of the growth of
capillary waves by electrolyte, thereby preventing film
rupture, and film rupture by nucleation being reduced in
the presence of electrolyte.
Both surface elasticity and surface diffusion [29.] will
oppose the growth of surface waves. Gibbs [30] has defined
the surface elasticity as
E

4C221

dl2
dM2


1
M1

where E is the Gibbs Elasticity, the suffixes 1 and 2 indicate


components of the film, the latter being in excess, C is the
surface excess, l is the chemical potential and M is the total

quantity of material per unit area of the film. The Gibbs


Elasticity is also written in the form [31],

4c
E

dc
dc

2

kB TD

where c is the concentration, c is the surface tension, kB is


Boltzmanns constant, T is temperature and D is a poorly
defined measure of thickness. Note that surface elasticity is
often interpreted as arising from an increase in surface tension
when an expansion of the interface leads to a reduction in the
amount of adsorbed material at the interface in the presence of
a surface-active component. However, it equally applies
when the solute is depleted from the interface, as is the case
with many electrolytes (surface tension increases with concentration); thus, the mechanism giving rise to elasticity is not
a local increase in surface tension, as is often described.
Inspection of Eq. (2) reveals that the magnitude of the surface
tension change with concentration of solute is important, but
not the sign. Indeed, a correlation between the transition
concentration at which bubble coalescence is prevented and
the magnitude of the surface tension gradient has long been
recognized [31 34..], suggesting that the elasticity of the
film during rapid stretching associated with film thinning, or
capillary waves, is crucial in the prevention of bubble
coalescence. However Stoyanov and Benkov [29.] argues
that the Gibbs Elasticity disappears from the equations that
describe the drainage and hydrodynamic stability of thin
films and that surface diffusion dominates the behavior. Like
the Gibbs Elasticity, the drainage velocity determined using
the surface diffusion approach [29.] is also approximately
proportional to (dc/dc)2. Thus, the observed experimental
correlation with (dc/dc)2 may arise from the diffusion of
solute in the thin film rather than from Gibbs Elasticity. The
correlation with (dc/dc)2 also includes electrolytes that have
no influence on bubble coalescence, as they generally have
lower values of (dc/dc)2 [31]. Values of (dc/dc)2 below f 1.0
(mN2 m 2 M 2) indicate no bubble coalescence inhibition.
Large values of (dc/dc)2 may dampen capillary waves and
reduce film thinning and rupture. However, the coalescence
inhibition of some salts is not described by their influence on
surface tension. The tetramethylammonium acetate electrolyte is a strong test of the validity of the combing rules, as it is
a hh salt and bubble coalescence is prevented. However, the
(dc/dc)2 value of 0.25 predicts that it will have no effect on
bubble coalescence. Sodium acetate has a (dc/dc)2 value of
2.1, yet, has no significant influence on bubble coalescence;
thus, the value of this term as a true indicator of the role of
electrolyte must questioned.
Perhaps then, electrolytes inhibit rupture by preventing
nucleation in the film separating bubbles. It is known that
most electrolytes have an electrorestrictive influence on
water, causing a contraction in the volume occupied by
water upon the addition of ions. It is possible that electrorestriction is accompanied by an increase in the cohesiveness

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

of water. However, no reasonable correlation can be found


with electrorestriction. A reduction of gas concentration in
the film could reduce the probability of rupture. Indeed, a
possible influence of dissolved gas on bubble coalescence
has been raised [1,2..,33.,34..]. Ions are known to salt
out dissolved gas, and the electrolytes that more strongly
salt out gas are known to effect bubble coalescence inhibition
at lower concentrations [20,33.,34..]. The diffusivity of the
sparging gas has also been related to the amount of electrolyte required to prevent coalescence [35]. However, sufficient data on a complete series of electrolytes are lacking to
fully test these correlations. These data would be useful, as it
is possible that a reduction in the amount and ease with
which gas enters the aqueous film separating bubbles could
inhibit nucleation and film rupture. A possible mechanism
for rupture in the absence of electrolyte is the migration of
gas molecules into the liquid film under action of an
attractive force. If electrolyte prevents or reduces this migration, film rupture could be arrested.

7. Conclusions
The bubble coalescence process in aqueous electrolyte
solutions remains unresolved. The mechanism of film rupture remains unclear, and the means by which ions influence
coalescence behavior remains elusive. Additionally, the
influence of both ions in combination, as described by the
combining rules, remains a major challenge. However,
recent efforts at understanding the specific-ion interfacial
behavior at high salt concentrations may provide important
clues. A recent suggestion by Marcelja (personal communication and in this issue) focuses on the behavior of ions at
the interface. The recent work by Jungwirth et al. [36.. 38.]
and Ninham et al. [3.,39.,40..], which indicates that some
ions are attracted to the interface despite the image charge
repulsion, has inspired his interpretation. Marcelja proposes
categorizing ions based on their preference for the surface or
the bulk. The beauty of this proposal is that it naturally
incorporates a combining law where the influence of an ion
is dependent upon the nature of the other ions present. Thus,
ions that are both located in the bulk or both at the interface
will have little effect, but ion combinations that are located
in the bulk and at the interface will prevent bubble coalescence. Whilst in its infancy, this is an exciting proposal that
can be tested using mixtures of electrolytes. This permits a
tentative model to be proposed. Ion combinations that are
separated at the interface result in a reduction in the mobility
of the interface. The reduced mobility ensures that thermal
capillary waves grow less quickly, leading to an increase in
the stability of the film. At a sufficient concentration of ions
in bulk, the surface mobility is suppressed to an extent that
bubble coalescence does not take place within the lifetime of
a collision. Another related challenge awaiting experimental
attention is an investigation of electrolyte effects on bubble
coalescence in nonaqueous solvents. I propose that specific-

183

ion effects will again be important, but due to the different


properties of the interface, the influence of ion combinations
may be very different from one system to the next. Such a
study may see ion-specific behavior relevant to biological
systems.

Acknowledgements
Discussions with Barry Ninham, Stjepan Marcelja,
Hakan Wennerstrom and Pavel Jungwirth have been
illuminating. The assistance of Chiara Neto and Mika
Kohonen in the preparation of this manuscript is appreciated. I would like to acknowledge the support of the
Australian Research Council, through the provision of a
research fellowship.

References and recommended reading


[1] Craig VSJ, Ninham BW, Pashley RM. Effect of electrolytes on bubble
coalescence. Nature 1993;364:317 9.
[2] Craig VSJ, Ninham BW, Pashley RM. The effect of electrolytes on
..
bubble coalescence in water. J Phys Chem 1993;97:10192 7.
[3] Bostrom M, Williams DRM, Ninham BW. Specific ion effects: the
.
role of co-ions in biology. Europhys Lett 2003;63:610 5.
[4] Leontidis E. Hofmeister anion effects on surfactant self-assembly and
the formation of mesoporous solids. Curr Opin Colloid Interface Sci
2002;7:81 91.
[5] Franks GV. Zeta potentials and yield stresses of silica suspensions in
concentrated monovalent electrolytes: isoelectric point shift and additional attraction. J Colloid Interface Sci 2002;249:44 51.
[6] Hofmeier U, Yaminsky VV, Christenson HK. Observations of solute effects on bubble formation. J Colloid Interface Sci 1995;174:
199 210.
[7] Fisher LR, Hewitt D, Mitchell EE, Ralston J, Wolfe J. The drainage of
.
an aqueous film between a solid plane and an air bubble. Adv Colloid
Interface Sci 1992;39:397 416.
[8] Mahnke J, Schulze HJ, Stockelhuber KW, Radoev B. Rupture of thin
..
wetting films on hydrophobic surfaces: Part I. Methylated glass surfaces. Colloids Surf, A Physicochem Eng Asp 1999;157:1 9.
[9] Mahnke J, Schulze HJ, Stockelhuber KW, Radoev B. Rupture of thin
..
wetting films on hydrophobic surfaces: Part II. Fatty acid Langmuir
Blodgett layers on glass surfaces. Colloids Surf, A Physicochem Eng
Asp 1999;157:11 20.
[10] Doubliez L. Capillary instability of a fast-draining film. Colloids Surf
1992;68:17 23.
[11] Coons JE, Halley PJ, McGlashan SA, Tran-Cong T. A review of
.
drainage and spontaneous rupture in free standing thin films with
tangentially immobile interfaces. Adv Colloid Interface Sci 2003;
105:3 62.
[12] Ruckenstein E, Jain RK. Spontaneous rupture of thin liquid films. J
Chem Soc, Faraday Trans II 1973;70:132 47.
[13] Cain FW, Lee JC. A technique for studying the drainage and rupture
.
of unstable liquid films formed between two captive bubbles: measurements on KCl solutions. J Colloid Interface Sci 1985;106:70 85.
[14] Jeelani SAK, Hartland S. Effect of interfacial mobility on thin film
.
drainage. J Colloid Interface Sci 1994;164:296 308.
[15] Li D, Lie S. Coalescence between small bubbles or drops in pure
.
liquids. Langmuir 1996;12:5216 20.
[16] Schulze HJ. Einige Untersuchungen uber da Zerreien dunner Flus.
sigkeitsfilme auf Feststoffoberflachen. Colloid Polym Sci 1975;253:
730 7.

184

V.S.J. Craig / Current Opinion in Colloid & Interface Science 9 (2004) 178184

[17] Butt H.-J. A technique for measuring the force between a colloidal
.
particle in water and a bubble. J Colloid Interface Sci 1994;166:
109 19.
[18] Yoon R.-H., Yordan JL. The critical rupture thickness of thin water
films on hydrophobic surfaces. J Colloid Interface Sci 1991;146:
565 72.
[19] Blake TD, Kitchener JA. Stability of aqueous films on hydrophobic
..
methylated silica. J Chem Soc, Faraday Trans I 1972;68:1435 42.
[20] Deschenes LA, Zilaro P, Muller LJ, Fourkas JT, Mohanty U. Quantitative measure of hydrophobicity: experiment and theory. J Phys
Chem, B 1997;101:5777 9.
[21] Christenson HK. M C P: direct measurements of the force between
.
hydrophobic surfaces in water. Adv Colloid Interface Sci 2001;91:
391 436.
[22] Miller JD, Hi Y, Veeramasuneni S, Lu Y. In-situ detection of butane
gas at a hydrophobic silicon surface. Colloids Surf, A Physicochem
Eng Asp 1999;154:137 47.
[23] Ishida N, Inoue T, Miyahara M, Higashitani K. Nano bubbles on a
.
hydrophobic surface in water observed by tapping-mode atomic force
microscopy. Langmuir 2000;16:6377 80.
[24] Steitz R, Gutberlet T, Hauss T, Klosgen B, Krastev R, Schemmel S,
et al. Nanobubbles and their precursor layer at the interface of water
against a hydrophobic substrate. Langmuir 2003;19:2409 18.
[25] Wennerstrom H. Influence of dissolved gas on the interaction between
.
hydrophobic surfaces in water. J Phys Chem, B 2003;107:13772 3.
[26] Considine RF, Drummond CJ. Long-range force of attraction between
solvophobic surfaces in water and organic liquids containing dissolved air. Langmuir 2000;16:631 5.
[27] Deschenes LA, Barrett J, Muller LJ, Fourkas JT, Mohanty U. Inhibition of bubble coalescence in aqueous solutions: 1. Electrolytes.
J Phys Chem, B 1998;102:5115 9.
[28] Craig VSJ, Ninham BW, Pashley RM. Study of the long-range hy-

[29]

[30]
[31]
[32]
[33]
.

[34]
..

[35]
[36]
..

[37]
.

[38]
.

[39]
.

[40]
..

drophobic attraction in concentrated salt solutions and its implications


for electrostatic models. Langmuir 1998;14:3326 32.
Stoyanov SD, Benkov ND. Role of surface diffusion for the drainage
and hydrodynamic stability of thin liquid films. Langmuir 2001;17:
1150 6.
Gibbs JW. The collected works. New York: Longmans, Green and
Co.; 1928. p. 300 15.
Christenson HK, Yamiinsky VV. Solute effects on bubble coalescence. J Phys Chem 1995;99:10420.
Marrucci G, Nicodemo L. Coalescence of gas bubbles in aqueous
solutions of inorganic electrolytes. Chem Eng Sci 1967;22:1257 65.
Weissenborn PK, Pugh RJ. Surface tension and bubble coalescence
phenomena of aqueous solutions of electrolytes. Langmuir 1995;11:
1422 6.
Weissenborn PK, Pugh RJ. Surface tension of aqueous solutions of
electrolytes: relationship with ion hydration, oxygen solubility, and
bubble coalescence. J Colloid Interface Sci 1996;184:550 63.
Pashley RM, Craig VSJ. Effects of electrolytes on bubble coalescence. Langmuir 1997;13:4772 4.
Jungwirth P, Tobias DJ. Ions at the air/water interface. J Phys Chem, B
B 2002;106:6361 73.
Jungwirth P, Curtis JE, Tobias DJ. Polarizability and aqueous solvation of the sulfate dianion. Chem Phys Lett 2003;367:704 10.
Salvador P, Curtis JE, Tobias DJ, Jungwirth P. Polarizability of the
nitrate anion and its solvation at the air/water interface. Phys Chem
Chem Phys 2003;5:3752 7.
Bostrom M, Williams DRM, Ninham BW. Surface tension of electrolytes: specific ion effects explained by dispersion forces. Langmuir
2001;17:4475 8.
Bostrom M, Williams DRM, Ninham BW. Specific ion effects: why
DLVO theory fails for biology and colloidal systems. Phys Rev Lett
2001;87:8103.

Vous aimerez peut-être aussi