7 The Stability/instability of Bubbles and Foams
7 The Stability/instability of Bubbles and Foams
and foams
The paradox is easily explained. Profit-seeking people will take more financial risk when they
believe the coast is clear. By taking bigger chances, however, they unwittingly make the world
unsafe all over again.
Hyman Minsky, US Economist, Paying the Price for the Fed’s Success, The
New York Times, www.nytimes.com/200801/27/opinion/27grant.html
7.1 Overview
All foams are thermodynamically unstable due to their high interfacial free energy, the
decrease of which causes foam decay. It is well known that there are several different
types of mechanisms involved in the stabilization and decay of foams, which has
caused a considerable amount of confusion. In the literature there are many conflicting
explanations frequently caused by experimental anomalies and the incomplete inter-
pretation of foaming experiments. Another aspect to consider is that the lifetime of
a foam can pass through several different stages, and each stage may involve
a different type of mechanism. To explain the overall stability in terms of one
mechanism is almost impossible, and the interplay of different mechanisms needs to
be taken into consideration. During generation, bubbles expand and contract and are
subjected to severe vibrations and dynamic disturbances causing distortion of
the adsorption layer. During this process, the liquid films separating the bubbles are
relatively thick and subject to stretching, and viscous elastic forces play a crucial role.
Possibly the most important mechanisms for the survival of a wet foam during this
stage involves the surface elasticity theories of Gibbs and Marangoni. Gravitational
forces also cause fairly rapid drainage to occur during this preliminary stage, but this
can be retarded by a high bulk viscosity. On entering a secondary stage, capillary
forces come into play causing suction and thinning of the lamellae, and this occurs
at a lower rate. In addition, disproportionation may occur causing the diffusion of
gas between bubbles. As all these processes occur under dynamic conditions, the
equilibrium adsorption coverage is rarely reached.
The process of gas diffusion owes its origins to the difference in pressure, surface
tension and curvature of the bubbles, but the gas diffusion to the atmosphere also needs
to be considered. In addition to diffusive disproportionation theories to explain the
changes in size distribution in bubbles, alternate processes have been considered
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.1 Overview 221
which involve the effect of interfacial rheology on the shrinkage of bubbles. As the
smaller bubbles are further reduced in size, the adsorption layer becomes compressed
during shrinkage, causing an increase in the surface excess (Γ) and a decrease in surface
tension. This process would act to counteract shrinking. Under these circumstances, the
driving force for disproportionation will also be influenced by the surface dilational
modulus of the film. However, bubble shrinkage is usually a fairly slow process, so that
the dilational modulus may have sufficient time to relax. At the moment, the role of the
surface rheological parameters on diffusive disproportionation remains controversial.
At the same time, evaporation, external vibration and shock may occur at the surface of
the foam. However, provided the thick films survive drainage, the foam can proceed
toward the final state of stability described by the dry foam state. In this final stage, an
equilibrium is reached in the formation of the thin lamella films, stabilized by the
“disjoining pressure” theories, as discussed in Chapter 3. For well-drained quasi-static
foam films (< 100 nm), intermolecular forces become dominant and stability must be
discussed in terms of common black films (CBFs) and Newton black films (NBFs).
In this chapter the main stabilization mechanisms are reviewed, and these are summar-
ized in Table 7.1. More recently, other stabilizing theories involving partial hydrophobic
particles have been under intense study and are discussed in Chapter 8.
3. Disproportionation Low solubility gases (low gas diffusion) Bubble soluble gas through perfluorohexane
(Oswald’s ripening) Air, N2 > O2 > CO2 type vapours or use surfactant to control type
Gas diffusion through of film.
liquid phase (Common black or Newton black films)
High Low
press pressure
4. Evaporation, temperature Temperature and humidity control Well packed coherent dense
gradients, humidity packed molecular films
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
222 The stability/instability of bubbles and foams
Clearly, there is a sharp distinction between the rates of decay of different types of
foams and, for convenience, the instability can be classified into the following different
types: (I) unstable or transient foams with lifetimes of seconds or even less; (II)
metastable foams with lifetimes of hours or even days; (III) high stability foams
which exhibit stability of the order of weeks or a few months; (IV) ultra-stable foams
which are stable for at least 6 months and in some cases for years. There is an industrial
demand for high-stability foams, particularly in the manufacture of polymers, materials
and metal foams.
(a) (b)
Fig. 7.1 Transient foams (champagne) are dynamic, with lifetimes in seconds, whereas static foams
such as beer have lifetimes of hours. Drainage and disproportionation occur in both systems,
but the beer foams are stabilized by proteins which induce steric disjoining pressure interactions
between the lamellae.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.3 Reversing the stability of foams 223
In certain industrial processes, the stability of the foam needs to be tuned for specific
performance requirements. For example, there is an initial requirement for
a surfactant with strong foaming action, but, at a later stage in the process, the
foam needs to be reduced or suppressed without the use of defoaming agents. This is
the case in the cleaning of radioactive vessels, where a good foaming agent can be
applied to remove particles from the walls of the vessel, but then the foam needs to
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
224 The stability/instability of bubbles and foams
(a) (b)
60 100 Maximum hydrophobicity
Maximum hydrophobicity
Foamability (%)
50 75
Stability (%)
40 50
20 0
2 4 6 8 10 2 4 6 8 10
pH pH
Fig. 7.2 Foamability (determined by the maximum foam volume after 10 s of gas flow at a constant flow
rate of 5 ml/s) versus pH and (b) foam stability (after 300 s standing) versus pH. From ref (2).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.3 Reversing the stability of foams 225
HO CH3
OH
NH2 NH2
HO HO
ethanolamine hexanolamine
Fig. 7.3 Fatty acid system 12-hydroxyl stearic acid (12-HAS) surfactant which was dispersed in water
with an organic counterion (such as ethanolamine or hexanolamine). From ref (4).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
226 The stability/instability of bubbles and foams
(a)
1000
CO2 N2 CO2 N2
800
Foam volume (mm)
600
400
200
0
0 100 200 300 400 500 600 700
Time (s)
(b) Me Me
CO2 + H2O
R R HCO3–
N NMe2 N2 N + NMe2
H
Fig. 7.4 (a) Foam control by diffusing different types of gases through foaming solution. (a) The
transformation from stable to unstable foam by changing the type of gas. Two cycles of treatment
with CO2 followed by N2 are shown. (b) The change in surfactant structure. Bubbling with
CO2 results in the conversion of imidazoline (HEAIs) type surfactant (non-foaming) to
the 2-alkyl-1-hydroxylethylimidazolinium carbonate cationic (HEAIBs) type surfactant.
From ref (6).
temperature inside the foam to increase, resulting in the transition of tubes into micelles
and leading to rapid destabilization. The approach was extended to magnetic particles,
producing the first foam to exhibit a thermo-photo-magneto-tunable response.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.5 Interfacial rheology 227
The results indicate a lower volume fraction in the second cycles, possibly because
this conversion process was not 100% efficient. This type of triggering mechanism was
thought to be useful for generating and breaking foams in gas wells, but these foam
systems had a disadvantage in that they became unstable at higher temperatures.
Surface tension gradients are essential for a freshly produced foam to survive.
However, the main deficiency in the early studies on Gibbs elasticity (as discussed
in Chapter 1) was that it applies to thin films, and the rate of diffusion of surfactant
from bulk solution to the bubble was neglected. In fact, the Gibbs theory only applies
to a hypothetical equilibrium state (i.e. it is assumed that there is insufficient
surfactant in the film to diffuse to the surface and lower the surface tension). For
thick lamellae, under dynamic conditions, the Marangoni effect becomes important
and operates on both expanding and contracting films. The Marangoni effect tends to
oppose any rapid expansion or contraction of the surface and may provide
a temporary restoring or stabilizing force to thin films, which are susceptible to
rupture. In fact, the Marangoni effect is superimposed on the Gibbs elasticity, so
that the effective restoring force is a function of the rate of extension, as well as the
thickness. The Gibbs–Marangoni effect can also explain the maximum foaming
behavior which occurs at a critical surfactant concentration in transient foams, as
explained schematically in Fig. 7.5.
If the solution is too dilute, then the greatest possible differential surface tension will
only be relatively small and little foaming will occur (Fig. 7.5(a)). In the case of the
solution being too concentrated, the differential tension relaxes too rapidly because of
the supply of surfactant which diffuses to the surface. This causes the restoring force to
have time to counteract the disturbing forces and produce a thinner film (Fig. 7.5(c)),
resulting in poor foaming. It is in the intermediate surfactant concentration range that
maximum foaming occurs (Fig. 7.5(b)). As yet, there is no clear-cut technique to
measure the magnitude of the Marangoni effect, and the theoretical treatment is incom-
plete. The damping of waves (without changing the wavelength of the ripples) in dilute
pure surfactant is believed to be caused by the Marangoni effect, since the surfactant
may not impart any significant surface viscosities.
Although the reduction in surface tension is the main driving force for the genera-
tion of foam, it is the rheological properties that impact the stability. During foam
generation, it is the expansion and contraction of the air/solution interface under
strongly dynamic conditions that prevail, while during drainage, it is the compres-
sion, coalescence and disproportionation which are the most important factors.
During these processes, the initial equilibrium state becomes distorted, causing the
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
228 The stability/instability of bubbles and foams
amphiphiles to dissolve in the bulk solution or adsorb at the interface to restore the
equilibrium coverage. Several models were developed for understanding adsorption
of surfactant at the surface of bubbles, and two extreme cases were taken into
consideration: (a) rapid deformation of the interface so that a non-equilibrium
adsorption predominates and (b) slow deformation of the interface such that the
adsorption is in equilibrium. In considering the rheology of the interface, different
types of deformation occur. The main ones have been classified by (a) the dilational
shearing modulus, where the deformation involves an increase or decrease in area of
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.5 Interfacial rheology 229
(a)
(b) a da
Shear deformation
Fig. 7.6 Examples of a 2D dilational and shear deformation (measured upon expansion/compression)
of an interface. From ref (7).
the liquid surface whereas the shape of the area remains the same, and by (b) the
shearing deformation, where the shape of the liquid interface is changed while the
interface area is kept constant. These processes are illustrated in Fig. 7.6.
External perturbations can cause internal relaxation of the adsorbed surfactant film,
and this behavior can give useful information on the dynamics of adsorption layers
and the stability of the foam. The drop shape and the capillary pressure tensiometer are
suitable methods to measure the dilational rheology of the interfacial layers which
define the key properties for understanding the behavior of foams. Dilational elasticity
and viscosity are nominated as the important parameters for understanding the
dynamic behavior of the interfacial layers. Dilational elasticity of the adsorbed
surfactant can be better understood by introducing an intrinsic compressibility of
the surface layer. Surface shear rheology provides important information on the
structure of the interfacial layer.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
230 The stability/instability of bubbles and foams
interfacial tension over adsorption (E* = dγ/dln Γ), which also corresponds to the
derivation of interfacial tension over bulk concentration (E* = dγ/dlnc). E* is then
dependent on both thermodynamic and kinetic characteristics of the foam systems
and may be considered as a dilational or compression modulus (surface stiffness or
resistance to deformation) that relates the dilational change in surface area and the
resulting change in surface tension. Essentially, surface viscosity reflects the speed of
the relaxation processes that restore the equilibrium in the system after imposing
stress on it. It is also a measure of the energy dissipation in the surface layer.
In contrast, the surface elasticity is a measure of the energy stored in the surface
layer as a result of an external stress. However, it is generally difficult to distinguish
between the effects on surface viscosity and the surface elasticity, although the steric
factors which effect the packing of the mixed film presumably control the rheological
behavior of the viscous mixed layers. Isolating the interfacial viscosity is also
difficult since the interface is connected to the substrate so that interfacial viscosity
is coupled to the bulk viscosity.
In addition to diffusion from the bulk phase, other kinetic processes play a role inside
the adsorbed layer, such as orientation, aggregation and compression, and cause rear-
rangements of the molecular structure. These factors affect E* and the magnitude of the
forces stabilizing the foam films, and, overall, it is difficult to achieve a complete
understanding of the surface rheology of soluble and insoluble layers at the interface
without knowledge of the shear viscosity, shear elasticity, dilational elasticity, dilational
viscosity and transport effects. In 2005, Ivanov (8) carried out a detailed analysis of the
range of aspects of interfacial viscoelasticity and concluded that it was impossible to
present the exact definition of the surface dilational viscosity. In many cases, the
experimentally determined values depended on the interpretation of results and char-
acteristics of the model. At sufficiently high deformation frequencies, E* may reach
a limiting value, which corresponds to a state of equilibrium adsorption where the
molecular exchange is suppressed and the adsorption layer will behave as an insoluble
monolayer (no diffusion exchange). At much lower frequencies, there is no resistance
against area change and the surface remains in equilibrium.
In general, the most interesting region is within intermediate frequency range, which
for soluble surfactants is about 1–500 Hz, where relaxation processes take place in and
near the interface (e.g. diffusion exchange). This causes viscoelastic surface behavior
and, therefore, the viscoelastic modulus E* has both elastic and viscous components.
In the case of sinusoidal deformation in which adsorption and surface tension oscillate
around their equilibrium values, the surface dilational modulus is a complex function of
frequency. At high surfactant concentrations, the interfacial tension follows the area
change with a certain delay due to the various relaxation processes within the interfacial
layer and between the interface and bulk solution. In this more general case, the surface
behavior is complex and E* has both elastic and viscous components. In the case of
soluble surfactants, the magnitude of the elasticity modulus E is a strong function of
frequency of external disturbances. For a more complete understanding of the surface
behavior, the complex surface dilational modulus E* (or visco-elastic modulus) has been
expressed in the form of
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.5 Interfacial rheology 231
0
E ¼ E þ iE″ ð7:1Þ
0
where E is called the storage modulus (dilational elasticity) which is obtained as the
pure dilational elasticity and E″ is loss modulus and is proportionate to the viscous
contribution where
E″ ¼ ωηd ð7:2Þ
and ηd is the surface dilational viscosity and ω the circular or angular frequency. E*
is a complex function, and the molecular interpretation of the dissipative process
that gives rise to the imaginary part of the modulus is subject to some controversy.
Also, E* = dγ(t) /dlnA(t) where dγ(t) and dlnA(t) are incremental changes in surface
tension and area, respectively, and ɳd is determined by the changes from the phase
angle between the generated surface area changes and the response (changes in
surface tension). For pure elastic interfaces (insoluble monolayers), the dynamic
interfacial tension response γ(t) follows the area change A (t) without any phase
difference, ηd → 0 and E* = E̍ .
Bos (9) also stressed the importance of the process kinetics and interfacial rheology
in the generation and breakdown of foams. During the generation of foams, new
interfaces are produced in a millisecond or less, and time scales with rapid expansion
and compression and shear viscosity would probably be more relevant under these
conditions. However, high interfacial shear viscosity would probably make
a contribution to long-term foam stability. In conditions where bubble break-up
occurs, large deformations and high strain rates need to be considered, and since
these processes involve large deformations in area (mainly dilational), dilational
elasticity and dilational viscosity play important roles. However, the coalescence
process is not completely dependent on rheology since activation energies, bending
energies and the bulk rheological properties (particularly in high volume fraction and
high collision frequencies) play an important role. Under more quiescent conditions,
such as when the foam is standing, lower strain rates would be involved. During
drainage, several different types of deformation processes may occur. For example,
liquid flowing downward due to gravity from the top of the bubble droplet would be
subjected initially to dilatation forces, whereas on reaching the bottom will be com-
pressed. In the middle of the droplet, where the distance between bubbles is the
narrowest, the interface is subjected to shear deformation. This situation is illustrated
in Fig. 7.7. In these circumstances, the overflow cylinder method and the maximum
bubble pressure method would possibly be relevant methods of measuring the inter-
facial elasticity.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
232 The stability/instability of bubbles and foams
(a) (b)
x-y plane x-z plane
Dilation
Gas Gas
cell cell Shear Shear
Compression
was related to the frequency of the perturbation. This approach led to the establishment
of the Lucassen van Temple Hansen (LvdT) model. This was successful in accounting
for many surface elastic systems but failed to completely account for the viscoelastic
properties of the monolayer. The theory was modified by Wantke and coworkers (11)
where a sub-layer was built into the model, and it was assumed that the surface
viscoelasticity was associated with a delayed exchange process between the sub-
surface and the monolayer. It was also assumed that this exchange process was relatively
rapid compared to the deformation rate in surface elastic systems, and this model was
found to be useful in describing the data obtained from measurements of surface
elasticity and surface viscoelasticity of foam systems. Over recent years, several other
models were developed to include underlying dissipative processes such as surfactant
re-orientation, surface compression, phase transitions or surface reaction in the surfac-
tant monolayer (12).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.5 Interfacial rheology 233
elasticity and the physical interaction between the adsorbed molecules on the surface
viscosity.
Surface laser scattering techniques have proved particularly useful for determining
the viscoelastic properties of monolayers, and this technique is very sensitive to the
behavior at the air/solution interface. When transversal ripples occur, periodic dilation
and compression of the adsorbed layer occur which can be studied. This enables
the viscoelastic behavior of equilibrium and non-equilibrium monolayers at an air/
water interface to be characterized without disturbing the original state of the mono-
layer. Surface elasticity values were obtained from surface dilational modulus mea-
surements determined at different angular frequencies for films of specific thickness.
It is important to consider the thickness and adsorption behavior since these para-
meters are also inter-related with elasticity. More recently, Derkash and coworkers
(13) gave an extended review of the more recent methods of measuring the rheological
properties of interfacial layers. Magnetic needle viscometers and rods have also
been developed which enable sensitive measurements of the interfacial stress to be
achieved (14, 15).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
234 The stability/instability of bubbles and foams
(a) (b)
Oscillating bubble
Light
Liquid
Piezo Compression
Equilibrium
Pressure
Computer
transducer
Manual Dilatation
dosing
Interface
Fig. 7.8 (a) The measuring cell in which the change in surface tension is monitored by capillary pressure
measurements and (b) profile of the pulsing bubble showing the adsorption kinetics of the
surfactant. Fast oscillation up 100 Hz can be easily achieved.
~
A = A0 + Asin (2 P vt)
D
~
dA/A γ = γ0 + γ (2 P vt + D)
A ~
A
Elastic
surface
dγ
γ
~
γ
Amplitude is a
Visocoelastic surface measure of elasticity
Time
Fig. 7.9 The response to the interfacial tension (γ) and interfacial area (A) to the sinusoidal perturbations of
a pendant drop as a function of the component of frequency. For the viscoelastic surfaces the
interface storage shear modulus and the interface loss shear modulus can be obtained from the
phase difference between the two signals.
obtained at a fixed frequency are shown. In Fig. 7.10, a typical set of results are shown
for a surfactant which was measured at three different concentrations.
For a low amplitude harmonic perturbation, the complex dilational viscoelastic
modulus or the dilational viscoelasticity E can be expressed by the equation:
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.5 Interfacial rheology 235
(a) (b)
dA/A Amplitude is a measure of elasticity
C1 65
60
C1
C2
50
55 d γ/γ
C3 C2
A 40 γ
Phase log is a measure
30 of the dilational viscosity 45
C3
20 Area
10 35
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
Time Time
Fig. 7.10 Harmonic oscillations of the interfacial surface tension and the drop area at three different
concentrations of surfactant as represented by C1, C2 and C3.
Δγ
E¼ πr2 ð7:3Þ
ΔA=A0
The complex surface dilational modulus can be obtained from analysis of the
amplitude and relative phase shift between the change of area of the bubble image and
the change of interfacial tension as a function of time. More in-depth discussions of these
techniques are given by Ravera and coworkers (18) in 2010, by Miller and Liggieri (19)
in 2005, and Liggieri and coworkers (20). This oscillating bubble instrument has
been successfully used at fairly high frequencies, enabling interesting data to be
collected on the relaxation mechanisms occurring between the surface layer and
adjacent bulk phase within the surface layer. This technique has proved useful in
understanding the composition of the interfacial layer and strength of the inter-
action between adsorbed species. The pulsating bubble module has also been used
to investigate the influence of the surfactant concentration above and below the
CMC on E*. Studies have been reported on a wide range of different types of
surfactant systems, including blends of surfactants, polymers and surfactant/poly-
mer mixtures (20).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
236 The stability/instability of bubbles and foams
wave motion of aqueous surfaces covered with fatty acid and lecithin monolayers
were reported by Hard and Neuman (21) in 1981. Quantitative determinations of the
film dilational elasticity and film dilational viscosity were reported. Viscosity mea-
surements by light scattering were comparable in magnitude to surface viscosities
obtained by mechanical (shear) viscometers. Colegate and Bain (22) carried out
ellipsometry studies on a liquid jet with a micellar solution of a nonionic surfactant
(C14EO8). The results were discussed in terms of the diffusion of micelles to the
interface, the adsorption step and the immediate breakdown of the micelles into
monomer species. These observations contradicted standard micellar models since
it was suggested that the micelles may break down into monomers not only in bulk
solution but also in the sublayer before adsorption.
In 2003, Wantke and coworkers (23) measured the surface dilatational properties
of aqueous solutions of sodium dodecyl sulfate (SDS) and n-dodecanol in the
frequency range 1–500 Hz using the oscillating bubble method. It was found that
pure dodecanol solution exhibited an elastic surface without viscous effect, whereas
the surface of an SDS solution without added dodecanol exhibits a strong viscoe-
lastic behavior. Mixtures showed a graduated transition from elastic to viscoelastic
behavior and it was suggested that the dodecanol molecules drive the SDS mole-
cules slowly out of the surface. This study established that a one-component model
describing the surface dilatational modulus can be used also for these mixtures.
Surface dilational modulus measurements carried out in the mid-frequency range
were reported by Anderson and coworkers (24) in 2006. The dynamic state of the
surface layer was monitored using a pressure sensor and surface second-harmonic
generation (SHG) recorded. The pressure sensor measures the real and imaginary
part of the modulus while SHG monitored independently the surface composition
(the concentration of soluble surfactant) under dynamic conditions. Two aqueous
solutions were studied, a surface elastic and surface viscoelastic solution.
The results could be explained by the LvH model, which included a molecular
exchange process and provided evidence for a non-equilibrium state within the
surface phase.
Recent trends have led to the combination of the oscillating bubble method with
other advanced surface analysis techniques. Possible future studies could involve
coupling the surface rheology with more precise measurements of the adsorption
kinetics, adsorbed amount and surface layer composition. The combination of shear
rheometry, ellipsometry and Brewster angle microscope might give more advanced
insight into the different mechanisms involved in foaming processes.
The stability of foams can be controlled by using single surfactant systems and
controlling the aqueous environment by adjustment of pH or by adding inorganic
electrolytes to produce specific ion effects. There is also the possibility of adding
a second surface-active agent such as a low molecule long straight-chain unbranched
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 237
(a) (b)
Cohesive interactions
Air
Air
Solution Solution
Air Air
(c) (d)
Cohesive interactions
Air Air
Solution Solution
Air Air
Particle where θ > 90º
Polymer / Surfactant (mixed film)
(low degree of wetting)
Fig. 7.11 High packing densities result from certain types of mixed surfactant species which can cause
strong cohesive interactions similar to polymer or particle films. For example: (a) mixed
surfactants system, (b) polymers (c) polymer/surfactant (d) particles with high contact angle
attached to the air-solution interface.
7.6.1 Single surfactant systems, pH, electrolyte and specific ion effects
As discussed in Chapter 2, the foaming of long straight-chain hydrocarbon ionic
surfactant reaches a maximum value around the CMC, but the addition of inorganic
electrolytes to the surfactant solution decreases the CMC, enhancing the foaming.
Generally, at low electrolyte concentrations, specific adsorption effects change the
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
238 The stability/instability of bubbles and foams
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 239
Table 7.2 Correlation between foam stability and CMC lowering by additives for sodium
2-n-dodecylbenzene sulfonate solutions. From ref (25).
Foam stability
Decrease in CMC Foam volume (ml)
Additive CMC (g/L) (%) after 20 min.
None 0.59 18
Lauryl glycerol ether 0.29 51 32
Laurylethanolamide 0.31 48 50
n-Decyl glycerol ether 0.33 44 34
Laurylsulfolanylamide 0.35 41 40
n-Decyl-3-sulfolanyl ether 0.35 41 28
n-Octyl glycerol ether 0.36 39 32
n-Decyl alcohol 0.41 31 26
N-(3-Sulfolanyl)oleylamide 0.48 19 23
Caprylamide 0.50 15 17
Lauryl chlorohydrin glycerol ether 0.57 3 22
Tetradecanol-2 0.60 0 12
Further, a correlation between CMC lowering and foam stability indicates that the
more effective stabilizers are those which cause the greatest lowering of the CMC.
An early systematic study by Schick and Fowkes (27) with a series of different
detergents, showed the importance of different additives on reducing the CMC. From
this work it can be concluded that the additives which had the greater influence in
reducing the CMC were the most effective foam stability boosters. The additives with
unbranched paraffin chains (which were equal in length to the detergent molecule) and
that had highly hydrophilic nonionic polar groups at one end were the most effective.
This can be illustrated by the effect of different types of surfactants on the stability of
sodium 2-n-dodecylbenzene sulfonate, are shown in Table 7.2.
Other effective foam enhancers or boosters include alkanolamide-type surfac-
tants. Sanders and Knaggs (28) reported that these types of additives were extre-
mely effective in practical systems. Early studies suggested that extremely closely
packed layers are formed in the mixed surfactant films containing alkanolamides
due to the formation of multiple hydrogen bonds, which improves the rheological
properties. Rodriguez and coworkers (29) carried out an extensive phase study and
small-angle X-ray scattering with an alkanolamide-type foam booster, dodecanoyl
N-methyl ethanolamide (NMEA-12). The structure of the surfactant is shown in
Fig. 7.12.
This surfactant was reported to have a lower melting point (compared to other
alkanolamides) and formed a lyotropic liquid crystalline phase at room temperature
which was attributed to an attached methyl group. This increases the fluidity of the
molecule. In SDS/NMEA-12 system, wide hexagonal and lamellar liquid crystalline
regions were detected at relatively low surfactant concentrations. For both the SDS and
the C12EO8 surfactants, the addition of NMEA-12 caused the effective area per molecule
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
240 The stability/instability of bubbles and foams
to shrink, producing a tightly compact layer which correlated with the foam booster
characteristics.
Foam boosters are needed in detergent, body-care shampoos and hair-conditioner
formulations which are mainly stabilized by anionic surfactants. In these systems, oil
droplets from grease from the body or hair become dispersed in solution and act as
antifoamers. It is often necessary to add an amphoteric and nonionic cosurfactant to
suppress the antifoam effects of oil. Basheva and coworkers (30) evaluated a series of
surfactants as potential as a foam boosters in the presence of a model detergent system
containing silica oil. The surfactants included lauryl amide propyl betaine (LAPB),
lauryl acid diethanol amide (LADA), lauryl alcohol (LA) and a glycerine-based
nonionic. The weakly foaming solutions contained a commercial anionic surfactant
(sodium dodecyl polyoxyethylene-3-sulfate (SPD 3S)), and the silicon oil droplets
were dispersed in foaming solution as micrometer-sized droplets. From thin lamellae
film experiments, it was shown that the enhanced foam stability in the presence of
LAPB and LA correlated with the energetics required to overcome the entry of the oil
drop into the air/water interface. It was also shown that the size of the silicone droplet
had a significant effect on foam stability and that droplets above about 5 microns
caused foam destruction. The mechanism involved the compression of the droplets by
walls of the Plateau channels. This resulted in a reduction in film thickness during
drainage.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 241
Fig. 7.13 Mixed polyelectrolyte/surfactant film with (a) adsorbed polymer on the surface (b) via formation
of surfactant/polymer complexes or (b) via hydrophobic attraction and (c) mesoscopic structured
poylectrolytes chains in the film bulk. The distance Δh between the neighboring branches is
independent of the kind of surfactant but differs with the kind of polyelectrolyte. From ref (32).
been shown to be involved (31). These have been shown to include (a) the redistribution
of the surfactant between the bulk solution and the coil regions, with the surfactant
molecules bound individually along the chain, (b) surfactant molecules clustered around
hydrophobic sites on the polymer and (c) polymer molecules wrapped around surfactant
micelles in such a way that the polymer segments partially penetrate and wrap around the
polar headgroup regions of the micelles. It has also been suggested that electrostatic
interactions are important in these systems, and the adsorption of these types of complex
species at the interface has an important influence on the foaming.
Klitzing and Muller (32) reviewed the different parameters involved in stabilizing
binary polymer/surfactant films under dynamic and static conditions. The importance of
intermolecular interactions, including DLVO and steric forces, was considered for
relatively low and high molecular weight amphiphilic mixtures. The influence of
the complex architecture of the polymers in the case of block copolymers and
tri-copolymers was discussed in terms of electrostatic screening effects and coiling of
the grafted chains. These factors were particularly important under high ionic strength
conditions. Fig. 7.13 indicates the different types of interaction which may occur within
a thin foam film.
It has also been well established that surface complexes formed below and above the
CAC can cause a synergistic lowering of surface tension in these systems. In some cases
above the CAC, the stability becomes even higher due to the formation of gel-like
structures in the thin film. An overview documenting the different interactions involving
combinations of polyelectrolytes and surfactant with different charges was published
by Langevin and coworkers (33) in 2001. It is well known that with mixtures of cationic
and anionic polyelectrolytes, the thin film stability can be enhanced, particularly if
one species is only slightly charged. However, in addition to electrostatic interactions,
hydrophobic interactions and hydrogen bonding can also play a role in complex
formation. With polymer/surfactant interactions, an increase in polymer concentration
above CAC may cause desorption of complexes from the interface which will
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
242 The stability/instability of bubbles and foams
destabilize the film and reduce the foaming stability. For many industrial applications,
cellulose polymers are used in mixed formulation with low molecular weight surfac-
tants, and it is often necessary to optimize the formulation in order to prevent foaming.
Guerrini and coworkers (34) studied the interactions of an amino-alkyl carbonyl
cellulose derivative with SDS and reported that maximum foaming occurred in the region
where the charge of the polymer/surfactant complex was neutralized. Precipitation was
also reported to occur in this concentration range. For mixtures of a SDS surfactant and an
ethyl hydroxyl cellulose (EHEC) polymer, Djuve and coworkers (35) showed that the
formation and adsorption of complex polymer/surfactant aggregates at the interface had
a pronounced effect on foam stability in aqueous solution. The driving force of the
interaction has been discussed in terms of the hydrophobic interactions and the electro-
static interactions between polar SDS and the EHEC polymer. In these systems,
a synergism in foaming was established. This behavior was explained by the adsorption
of aggregates at the air/water lamellae interface leading to the formation of a coherent
gelatinous layer which reduced drainage.
In 2011, Langevin and coworkers (36) reviewed the stability and drainage of free
suspended film stabilized by surfactant/polymer mixtures, and in this paper, parti-
cular attention was paid to protein systems. Proteins alone are known to form brittle
monolayers at the air/water interface that can easily break and leave unprotected
areas. However, for mixed surfactant/protein films, the hydrophilic sections and
neutral hydrophobic segments can produce compact aggregates which may also
partially unfold at interface. In the case of a mixed monolayers produced from
surfactants and rigid polyelectrolytes of opposite signs, the surfactants may trans-
form the rigid proteins layer into more flexible and mobile entities that might be
capable of responding to applied stress without rupturing. This may be the reason
why protein/surfactant film mixtures are more stable than surfactants alone. There is
also a possibility that at high surfactant concentration, the proteins may be displaced
toward the film interior. From optical observations of the thin films containing
proteins, it could be suggested that the inhomogeneous films consisted of gel-like
aggregates, and interferometry showed different colors indicating a wide variation of
the film thickness. The adsorption of the aggregated proteins also corrugated and
deformed the surface, and their confinement in the film possibly provides a gel-like
jamming behavior. This prevented measurements of the disjoining pressure between
the interfaces of the thin films, but it has been shown by Bergeron and coworkers (37)
that it was possible to relate visually the surface characteristics of its the film to the
stability. Gel-like structures were observed within critical concentration levels in thin
film produced from polyelectrolytes with opposite charge. For example, the thin
film images produced from the PAMPS and DTAB systems are shown in Fig. 7.14.
These were reported to be extremely stable and near to the precipitation limit of the
complex.
Gel-like structures have also been observed in films produced from sodium caseinate
(Fig. 7.15). It was reported that the inhomogeneities in such films are mobile at
low concentration but become more strongly attached and static at higher
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 243
Fig. 7.14 Image of a film made from PAMPS and DTAB at a concentration near the precipitation
boundary (top and bottom: Schematic illustration of the film structure). From ref (37).
Fig. 7.15 Top view of casein thin films for three different casein bulk concentrations (a) c = 0.05 g/l, (b)
c = 0.3 g/l and (c) c = 0.83 g/l and showing heterogeneous thickness and the presence of
aggregates. The films become stable as soon as the aggregates percolate. From ref (37, 38).
concentrations (> 0.1 g/l, which corresponds to the stability threshold of foams gener-
ated with these solutions).
In industries, frequently mixtures of oppositely charged surface-active agents
(anionic and cationic surfactants and surfactant mixtures) are used in foaming for-
mulations; for example, mixtures produced from xanthane-gums, proteins, mixtures
of partially hydrophobic silica nanoparticles and short-chain amine surfactants.
These mixtures are sometimes referred to as catanionics. Usually these types of
mixtures produce bulk aggregates or soft entities which compact in the Plateau
borders of the foam preventing drainage.
Many basic research experiments using self-supporting foam films have been used to
study these systems. Schelero and von Klitzing (39) prepared foam films with a mixture
of a cationic surfactant (alkyl trimethyl ammonium bromide, CnTAB) and variable
amounts of anionic surfactants. Three different types of anionic surfactants with head
groups of different size and charge (sodium decanoate, sodium sulfate and sodium
styrene sulfate) were investigated. From these studies it was observed that with increas-
ing concentration of the anionic surfactant the film thickness and stability decreased to
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
244 The stability/instability of bubbles and foams
a minimum value that was dependent on the type of head group interaction, as well as the
hydrophilic/hydrophobic balance of molecules. Stronger interactions, due to hydropho-
bic interactions and the head groups, appeared to dominate the synergism. From the
study it could be suggested that the amount and type of anionic surfactant could be
adjusted to tailor the stability of the foam. Also, surface tension was confirmed as
a useful technique to study these systems.
7.6.5 Nanopatterning
In 2008, extremely stable microbubbles (stable for over 1 year) were prepared by
Dressaire and coworkers (41) from a solution of sodium stearate which was dispersed
and sheared in a highly viscous glucose syrup phase. The enhanced stability has been
explained by the formation of an elastic condensed surfactant phase with hexagonal
domains buckled outward from the bubble covered surface. It was suggested that it was
the elastic response of the interface which prevented the shrinkage of the bubbles.
The systems were found to have a wide bubble size distribution (ranging
from hundreds of nanometers to tens of microns). From cro-SEM and TEM imaging it
was shown that a self-assembled surfactant film coated the air/water interface.
Thermodynamic and molecular models based on the energetic competition between
the bending elasticity of the interface and the formation of domain boundaries were used
to describe the pattern geometry. It was suggested that these features of the nano-scaling
could be modified by changing the chemical composition of the film. Controlling the
nature of the surfactant self-assembly could prove to be a useful tool in designing
complex coatings to stabilize bubbles and foams. The structure of a neatly formed
hexagonal surface pattern has been clearly demonstrated (41).
7.6.6 Hydrophobins
Biomolecules such as hydrophobins can outperform synthetic molecules in many
applications, and there has been a growing interest in utilizing these natural polymers
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 245
in stabilizing bubbles and foams. These molecules are part of a class of small cysteine-
rich surface-active proteins composed of amino acids, and they are commonly found in
button mushroom. The structures contain distinctive hydrophobic and hydrophilic
amino acid residues in the protein sequence and can be considered to have similar
features to Janus particles. They form self-assemblies that are strong and irreversibly
adsorb at the air/water interfaces, and they have been described as the most surface-
active protein, known and produce surface tension reductions as low as 27 mN/m.
In 2009, Cox and coworkers (42) showed that bubbles could be stabilized for long
periods (for at least 4 months) and even up to several years in some cases.
The unique properties of hydrophobins have been reviewed by Linder (43) in 2009.
Hydrophobins have also been evaluated as bubble stabilization agents in aerated foods
such as ice cream, and early studies enabled them to be classified into two distinct
groups: class I and II. In class I, aggregates form in water which are highly insoluble,
whereas class II aggregates which are more soluble. On dispersion in water, class II
produce oligomers which dissociate into aggregates and membranes that disintegrate
when subject to shear deformation and transform into a viscoelastic layer similar to
a highly elastic membrane. It is interesting to note that macroscopic bubbles formed in
HFB II solution have been shown by Basheva and coworkers (44) to preserve the
irregular non-spherical shape they had at the moment of solidification, as shown in
Fig. 7.16.
There has been a considerable effort to understand the relationship between the
functionality and the molecular rearrangements of the hydrophobin fragments, which
Fig. 7.16 Photograph of millimeter-sized bubbles resting below the surface of a 0.005 w% HFB class II
solution containing added CaCl2 (25 mM). From ref (44).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
246 The stability/instability of bubbles and foams
are produced under shear. Emphasis has been directed toward determining the size
and geometry of the hydrophobic and hydrophilic head groups and the packing
density at the interface. Basheva and coworkers (44) also carried out monolayer
compression experiments and foam film studies which indicated that more mechan-
ical, stiffer films were formed with hydrophobins than with more common proteins
such as β-casein and β-lactoglobulin. These molecules appear to have sufficient
hydrophobic and also hydrophilic moieties to form compact films with high packing
densities, strongly attach to the bubble interface and reduce gas diffusion. They show
periodic wrinkles (ripples) with well-defined bending elasticity. It has also been
suggested that the protein monolayers of HFB II exhibited high mechanical strength
compared to typical amphiphlic proteins. From free foam film studies, it was found
that the interface was electrostatically stabilized, but near the pHiep, aggregates were
produced from dimers and tetramers that tended to adhere to each and to the air/water
interface. Generally, these large aggregates in water could be considered to exhibit
behavior similar to sticky particles, and this to some extent explains the compact
adsorption layers. Also, it was suggested that expanding monolayers vulnerable to
rupture could be repaired by the molecules filling the voids.
Adsorbed layers of hydrophobin HFB II with added β-casein were studied by
Radulova and coworkers (45) using a rotational surface shear rheometer. Experiments
were carried out at different shear rates and at different concentrations of added β-casein.
High values of the surface shear elasticity were explained by the formation of an
interfacial bilayer, with the layer from the more hydrophobic HFB II facing the air
phase. With increase in shear rate, the layers were found to fluidize, and the addition of
casein gave more rigid adsorption layers with faster fluidization. It was also possible to
study the different processes that occur in the viscoelastic adsorption layer in different
flow regimes which involve the breakage and restoration of intermolecular bonds and
solidification of the adsorption layers. In Fig. 7.17, the results from surface rheological
studies are shown together with a model illustrating the interfacial bilayer formed from
the casein/HFB II aggregate.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.6 Stability control agents 247
(a) (b)
Decreasing angular velocity
HFBII
40 adsorption
layer
Torque, τ (mN.m)
30
Adsorbed
HFBII tetramers
20
β–casein
10
HFBII +
0
β–casein
0 10 100 1000
Time, t (s)
Fig. 7.17 Surface shear rheology studies on hydrophobin II: (a) plot of the torque versus time at six different
angular velocities. The data indicate an increasing fluidization with rise in shear rate. (b) Top: HFB
II dimers and tetramers can adhere to the film due to attractive interactions between hydrophilic
parts of HFB II molecules. Bottom: With β-casein, an interfacial bilayer consisting of more
hydrophobic HFB II facing the water phase and a layer of more hydrophobic casein that faces the
water phase is formed. From ref (45).
reproducible experimental data on these systems. Particles can also jam on the bubble
surface and become irreversibly adsorbed, and these layers can form effective barriers
which decrease and even arrest the rate of bubble coarsening.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
248 The stability/instability of bubbles and foams
(a) (b)
100 100
80 80
60 60
40 40
CO2 / PFH
N2 / Water
20 0.42%, 0.83%, 3.75%
20 N2 / PFH
0.29% CO2 / Water
0.20%
0% 0.091%
0 0
0 100 200 300 400 500 600 0 250 500 750 1000
Time (sec) Time (sec)
Fig. 7.18 (a) Drainage rates of foams produced from mixtures of CO2 gas containing different
concentrations of Perfluoro vapor. In this experiment FC104 vapor was selected. Solution contains
1 g/l of Pluronic F68. Initial volume of solution = 50 ml, temperature 25°C and rate of foam
formation = 0.22 cm/s). (b) Comparison between two gases, CO2 and N2, with and without PFH
solution. 1 g/l of Pluronic F68. V0 = 50 ml, temperature 30°C and rate of foam formation =
0.22 cm/s. From ref (48).
structure is formed. This concept can be generally used to inhibit coarsening. For
example, by bubbling a soluble gas through a small amount of liquid such as perfluor-
ohexane (PFH), a small amount of insoluble vapor is introduced into the liquid causing
the lifetime of a foam to be extended from only a few hours to several days.
The theoretical background to this behavior has been presented by Weaire and cow-
orkers (47) and is based on the differences in mixing entropies of the gas molecules
which cause a counteractive driving force for gas diffusion.
The effectiveness of this method of stabilizing foams was evaluated in some detail
in experiments by Gandolfo and Rosano (48) in which foams were generated in
a column using a milk protein surfactant. Two systems were studied in which the
stability of the foams (generated by N2 and O2) was measured from the changes in the
drainage rates. Further experiments were made in which small concentrations of
insoluble secondary gases (vapors) were introduced into these gases. Experiments
were carried out with several different types of insoluble gases: Pluronic (F67: a block
copolymer of ethylene oxide and propylene oxide condensate), a PFH and other types
of commercial fluoro-inert (linear perfluoro molecules) designated FC 77, FC 104, FC
40 and FC 70. Fig. 7.18(a) shows a typical set of data of the drainage rates for foams
generated by CO2 containing increasing levels of the perfluoro vapor, compared to
foams generated by bubbling CO2 alone through the aqueous solution. The foams
were generated at a fixed flow rate, and the gas supply was switched-off after the foam
reached a pre-determined height.
These results show that an increase in the level of FC 104 in the gas mixture to above
0.42% caused the drainage curves to be reduced, and they finally level off. Also, in
Fig. 7.18(b), the drainage results for CO2 and N2 are shown, and it can be seen that
without the PFH vapor, the CO2 causes a large increase in drainage rate compared with
N2, as expected. Nevertheless, it can be seen that by using gas mixtures containing PFH,
the drainage can be substantially reduced for both gases.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.7 Interfacial rheology and gas permeability 249
(a) (b)
C12TAB C12TAB
Permeability coeff (cm/s)
SDS
0.4 0.4
0.0 0.0
Air
N2
Ar
O2
Air
N2
Ar
O2
Air
N2
Ar
O2
Air
N2
Ar
O2
Air
N2
Ar
O2
Air
N2
Ar
O2
Air
Xe
Fig. 7.19 Comparison of (a) gas permeability of CBF stabilized by β–C12G2, C12TAB, SDS to different
gases and (b) gas permeability of NBF stabilized by β – C12G2, C12TAB, SDS and AOS to
different gases. From ref (49).
Farajzdah and coworkers (49) measured the permeability rates (diffusion rates) of
foam films prepared with anionic, cationic and nonionic surfactant solutions with several
different types of gases which included argon, oxygen, nitrogen, air and in one case
xenon. All the surfactant solutions were above the CMC, and the electrolyte concentra-
tions were adjusted so that the film type could be changed. In the lower electrolyte
concentrations, CBFs were formed with the ionic surfactants with typical equilibrium
thicknesses of 10–15 nm, but at high electrolyte concentrations, NBFs were formed with
thicknesses about 5 nm at 0.5 M electrolyte. The results of the gas diffusion experiments
are summarized in Fig. 7.19 (a) and (b) for the two different film types, and this enables
a comparison to be made between different surfactants, gases and types of film.
The results from this study indicate that the gas permeability of the CBFs were
generally higher than that of the thinner NBFs, but this effect was mostly more
pronounced for films stabilized by SDS surfactants. It is difficult to give a complete
explanation of this behavior, but it is possible that the gas permeability is governed not
only by the film thicknesses but also by the nature of the interacting surfactant mono-
layers, which would be dependent on the surfactant adsorption densities.
There has been considerable debate on the influence of the surface elasticity on
disproportionation. It has been thought that during this process, as small bubbles
shrink, the film compacts and after a time the compression forces reach a sufficient
magnitude to retard the shrinkage. However, it is possible that the adsorbed surfactant
layers may collapse upon increasing compaction and form multilayers. A further
possibility is that the molecules desorb from the surface. In 2008, Tcholakova and
coworkers (50, 51) carried out a series of fundamental experiments using a mixed
surfactant system in order to study the influence of gas permeability on interfacial
rheology and drainage. Initially, a surfactant system was prepared by adding low
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
250 The stability/instability of bubbles and foams
2.8
SLES+CAPB
2.6
2.2
(Rt /R0)
SLES+CAPB+LAc
1.8
1.4 SLES+CAPB+MAc
1.0
0 1000 2000 3000
Time (seconds)
Fig. 7.20 Change of mean bubble radius R with time as a result of Oswald ripening. The bubbles are
observed at the foam contact with a glass plate (the size of the first layer of bubbles was measured).
From ref (50).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.7 Interfacial rheology and gas permeability 251
GAS GAS
P1 P2 P1 P2
JGAS JGAS
Fig. 7.21 Presentation of flux of dissolved gas across a foam film stabilized by surfactant with (a) high
surface modulus and (b) low surface modulus. The formation of surface-condensed phase of
surfactant molecules by myristic acid leads to a very low solubility and diffusivity of the dissolved
gas molecules in the condensed adsorption layers on the foam film surface. From ref (51).
200
WPI = (whey protein isolate)
150
Rb (μm)
Soy β-lacto
100
glycinin globulin
Oval-
50 WPI bumin Gelatin
Sodium
caseinate
0
0 50 100 150 200 250
r (min)
Fig. 7.22 Plot of bubble radius versus time for isolated bubbles near a planar air/water interface. Stabilized
by different proteins at pH 7. From ref (52).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
252 The stability/instability of bubbles and foams
Although the basic framework for studying gas diffusion to the atmosphere from single
bubbles is fairly well understood, there has been only limited work carried out to
understand the gas diffusion in real aerated food systems. These usually consist of well-
dispersed, high gas fraction systems with a large number of bubbles. In 2004, Dutta and
coworkers (53) generated two aerated food products by whipping mesophase mixed
dispersions (containing 33% sugar 65% water) with a 2% surfactant blend consisting of
sucrose stearate together with either propylene glycol stearate (PPGS) or triglycerol
stearate (TGS). The foams were allowed to stand, and the disproportionation process
was evaluated by measuring the change in bubble size using video frames and micro-
scopy. In addition, overrun, which was the decrease in the volume fraction of air in the
foam, was measured. This could be considered as the difference in the degree of air lost
to the atmosphere between the two systems due to the difference in diffusion rate in the
two systems. The results are shown in Fig. 7.23(a) for the TGS-sucrose stearate system
(a) (b)
20 30
Volume fraction
Volume fraction
of system (%)
15 25°C 25°C
whipping 10°C 20 whipping 10°C
10
10
5
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 80
Bubble diameter (microns) Bubble diameter (microns)
Fig. 7.23 The long-term stability of (a) triglycerol stearate surfactant system and (b) propylene glycol
stearate surfactant system. From ref (53).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.10 Stratification 253
and in Fig. 7.23(b) for the PPGS-sucrose stearate system. These samples were prepared
immediately after whipping and after a period of one week.
For the TGS system, the volume fraction did not change drastically after a week,
whereas the PPGS was found to be more unstable. Also, the temperature of storage did
not affect the size distribution to a great extent. The decrease in the overrun was
explained by the diffusion of air from the bubbles to the atmosphere and was found to
be higher at higher temperatures.
Two different types of models were developed to account for this behavior. The first
model described the gas diffusion in terms of an inter-bubble diffusion occurring
between bubbles of different sizes and driven by the difference in capillary pressure.
The second model described the diffusion process in terms of the gas flowing from the
larger bubbles to the atmosphere. Interestingly, the experimental results obtained from
this study did not agree with the prediction of either models and it was necessary to
construct a new gas diffusion model based on the combination of both models, which
assumed that both processes occurred simultaneously. This combined diffusion model
was found to give the most satisfactory fit to the experimental data.
7.10 Stratification
Johnnott (54) in 2006 and Perrin (55) in 1918 initially demonstrated that at surfactant
concentrations beyond the CMC, organized molecular structures can be formed within
thin films. This caused thinning of the film through a stepwise drainage mechanism
known as stratification. It has been well established over the past 50 years that stratifica-
tion can occur in dispersions of surfactant micelles, silica particles, microemulsions,
globular proteins polymeric macromolecules and colloidal nanoparticles. Extended
lifetimes have been observed for stratificating of liquid films, resulting in higher shelf-
stability of foams. On thinning of the film, the particulates are forced into the restricted
volume of the film which delays the drainage process and stabilises foam lamellae.
The different steps in the drainage of the thin film containing a surfactant (above the
CMC) are shown in Fig. 7.24(a) and for films containing low levels of surfactant in
Fig. 7.24(b).
Organized structures can be formed from micelles produced from both ionic
and nonionic surfactants in the aqueous solution and these provide an additional con-
tribution to the disjoining pressure. These micelle layers flow out of the film to the
Plateau borders, thus causing stepwise thinning, with each step corresponding to
specific black spot concentrations (Cb). The phenomenon was earlier summarized by
Ivan and Dimitrov (57), and several early experiments by Wasan and coworkers (58, 59,
60) showed that the ordering was caused by the interaction (via repulsive forces)
of surfactant micelles or colloidal particles with narrow size distributions which were
forced into the restricted volume of the film. These workers clearly showed that the
ordered microstructure could play an important role in stabilizing foam lamellae.
They suggested that the driving force for the stepwise thinning of the film was
caused by the gradient of the chemical potential of the micelles at the film’s periphery.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
254 The stability/instability of bubbles and foams
n=5 701 Å
n=4 592 Å
Thinning Thinning
n=3 483 Å
n=2 374 Å
n=1 261 Å
n=0 162 Å
Interferogram Interferogram
Stepwise thinning First or
common black
Film thickness
Film thickness
10 nm
Second or
micellar Newton black
diameter
Time Time
Fig. 7.24 (a) Stratification of thinning mechanism at high surfactant concentration (> CMC) and the
main stages in the evolution of a thin film and (b) thinning at low surfactant concentrations
(< CMC). From ref (56).
Both the experimental and theoretical aspects of these studies were summarized in 2007
by Wasan and Nikolov (61).
In 1992, Bergeron and Radke (62, 63) determined disjoining pressure isotherms for
single isolated foam films stabilized by SDS at concentrations greater than the CMC.
They reported that the oscillating component extended out to film thicknesses as large as
50 nm and that the discrete change in film thickness for each black film transition ranged
from 6 to 10 nm and depended on the ionic strength and surfactant concentration. Studies
of the oscillatory form of the disjoining pressure permitted quantitative interpretation of
the stepwise thinning behavior.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.10 Stratification 255
Fig. 7.25 The layer-by-layer thinning of the micellar films is triggered by the formation of black spots
inside the film. One can distinguish up to five stratified layers shown by photocurrent
time plot. From ref (65).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
256 The stability/instability of bubbles and foams
Fig. 7.26 Interdroplet interactions and resulting oscillating structural forces within the thin liquid film
separating the bubbles. Both attraction depletion forces and structural barriers occur between
bubbles. From ref (69).
particles, it does not fit the nonionics where the transition is a reversible process. In fact,
for the nonionics, the energy of both stable states must be almost identical.
Several theoretical models have been developed over the years by Wasan and co-
workers (66, 67, 68, 69, 70, 71) during the period 2001–2004 to rationalize the experi-
mental results. Generally, the oscillation component comprised an attractive depletion
component opposing the repulsive structural energy barrier inside the film. In the case of
very thin intervening film when the thickness is less than the micellar size, the micelles
became excluded from the film, and osmotic pressure inside the film is reduced. This
causes an attractive depletion force (i.e. Asakura–Osawa depletion attraction). However,
when the film thickness is of the order of several micellar diameters, the micelles are
ordered in the confines of the film surfaces, thereby resulting in a structural energy
barrier separating (stabilizing) the bubbles (Fig. 7.26).
Further studies were carried out based on statistical mechanics, and numerical
Monte Carlo simulations have been performed using an effective one-component
fluid potential model that incorporated particle hard-core repulsion. Both the
screened Coulombic electrostatic interaction and the entropic contribution due to
the discrete nature of the solvent and the finite size of the electrolyte ions were
included in these models. Concise formalism was developed based on the general-
ization of Ornstein–Zernike equations for fluid mixtures, and it was found that in
hard-sphere colloidal suspension, comprised of larger entities, the solvent molecules
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.10 Stratification 257
enhance the structural forces both among the larger particles themselves and between
them and the film interfaces (68, 69, 70).
(a) (b)
Film area: S = 2.6 10–4 cm2 Film area: S = 5.9 10–4 cm2
Capillary pressure: 40.7 Pa Capillary pressure: 38.5 Pa
Photocurrent (film thickness)
Photocurrent (film thickness)
42.7 nm
42.7 nm
33.2 nm
33.2 nm
25.0 nm 25.0 nm
17.0 nm
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
258 The stability/instability of bubbles and foams
larger film. In addition, it was observed that when the thickness is reduced from 250
to 175 microns, the film slowly increases in thickness from 17 back to 25 nm (from
one micellar layer to two micellar layers), indicating that micellar lamellar layer
thinning occurred as a reversible process. This unusual thinning behavior of micellar
nonionic films could not be explained by the previously proposed diffusive-osmotic
mechanism which had been successfully applied to anionic or cationic micellar
films. For charged ionic surfactants, the formation of black spots in micellar films
was based on a micellar vacancy diffusive-osmotic mechanism, which assumed that
the film at each stepwise transition is in a metastable state. In fact, for the nonionics,
the energy of both stable states must be almost same.
Generally, it was concluded from these studies that the stability of the micellar films is
governed by the film size (the bubble size), and the capillary pressure but the concept of
disjoining pressure isotherm alone cannot predict film thickness stability. Experiments
show that film stability depends more on the lamella or film size and micelle concentra-
tion and less on the capillary pressure.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.10 Stratification 259
charge distribution, and a cell model. The JA model assumes that micelle concentration
was uniformly distributed, despite the field of electric double layers. It was found that the
JA gave the best agreement between theory and experiment. In fact, two theoretical
approaches which were based on energy and osmotic pressure were found to give
reasonable agreement. The experimental data providing the suspension of charged
micelles was represented as Jellium. This model proved to be useful for determining
the aggregate number of ionic surfactant micelle from the experimental heights of the
steps. It was also concluded from theory (based on the dispersion of charged particles
behaving as jellium) and experiment that the thinnest stable films (h0) experimentally
observed were thick enough to contain micelles (at very low concentration) to ensure the
osmotic balance between the film and bulk solution.
2500
Area (mm2)
1500
500
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
Time (s)
Fig 7.28 Black spot expansion bidispersed mixtures at different film sizes. From ref (63).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
260 The stability/instability of bubbles and foams
2500
Area (μm2)
1500 8 nm (monodispersed)
Slope = 356 μm2/s
500
with time for both systems, but rate of expansion of the bidispersed systems was the
greatest for the large film sizes.
In a second series of experiments, the spot expansion rate for monodispersed system
was compared with a bidispersed system for films of about the same diameter (135
microns), and the results are shown in Fig. 7.29. In this case, it can be seen that the
monodispersed particles and the bidispersed particles left the film more rapidly.
Generally, it was concluded from these studies that the rate of particle movement from
the film to the bulk was the lowest for the monodispersed system, intermediate for the
polydispersed (prepared by mixing three different sized particles) and the highest for the
bidispersed system, and in general, monodispersed particles exhibit slower expulsion
and more stable films. However, in case of the experiments in curved and plain films, an
increased particle efflux from the film to the meniscus was reported for the polydispersed
systems. It was also found that the black spot area expands linearly with time and the
larger films exhibit faster rates of spot expansion which was explained by the DO model.
The influence of polydispersity on the structure barrier oscillation and oscillation forces
was discussed by Wasan and coworkers (65), and a representative picture of the inter-
action is shown in Fig. 7.30.
It was stressed that confined in-layer structures in films were essential for retarding
thinning and that polydispersity leads to the weakening of particle laying (decreasing in
the structural barrier) and film stability. But, in addition, it is the degree of polydispersity
that plays an important role in the height of the energy barrier. The interactions between
two approaching bubbles in a polydispersed system are shown in Fig. 7.30(a),
where initially a weak barrier (Us) is encountered but on further thinning smaller
particles are ejected from the film (Fig. 7.30(b)). This results in a reduction in poly-
dispersity and an improvement in layer structuring which causes the height of the
structural barrier to increase. However, on reaching a critical thickness the particles
are finally ejected and a depletion well is encountered. In Fig. 7.30(c), the relationship
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.11 Stabilization by liquid crystals 261
(a) (b)
Attraction Repulsive structural
(c)
80
Disjoining pressure
(d)
40
0.2
0
Height of structural barrier
U / kT
40 V% monodispersed 0.1
–40 40 V% with 30 V% polydispersity
[Polydispersity = (std. deviation)(mean size)]
–80
0 1 2 3 4 5 6 0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Film thickness,
Polydispersity
scaled by particle diameter
Fig. 7.30 Effect of particle polydispersity on the height of the structural barrier between film
surfaces. From ref (63).
between film thickness and the disjoining pressure is illustrated for a monodispersed
and polydispersed system. In both cases, the oscillating disjoining pressure is
increased as the film thickness is reduced for both systems. The height of the
barrier is a measure of the film stabilizing effect, and in Fig. 7.30(d), it is related
to the degree of polydispersity.
It is well known that liquid crystals can adsorb at the bubble interface and stabilize
both aqueous and non-aqueous foam systems. At higher surfactant concentrations,
the liquid crystalline surface phase is often in equilibrium with a bulk isotropic
phase. The stability in non-aqueous systems is covered in Chapter 9. Frequently,
liquid crystals have been formed in aqueous mixed surfactant systems, for exam-
ple, by the addition of small amounts of a nonionic surfactant to an anionic
surfactant, which can lead to enhanced foam stability. In addition, there are several
possible explanations to account for the increase in foam stability caused by the
precipitation or formation of crystalline or structural aggregates in the solution.
For example, they may cause a reduction in hydrodynamic drainage or retardation
in interfacial drainage caused by viscous layers or an increase in the mechanical
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
262 The stability/instability of bubbles and foams
Time (min)
0 50 100 150 200 250 300 350
C12EO9 C12EO7
Drainage volume (mL)
2 C12EO5 C12EO3
strength of a liquid film, which may increase the resistance between bubbles. Also,
the accumulation of the liquid crystal phase in the lamellae film may reduce the
diffusion rate of entrapped gas. In addition, there is a possibility that accumulation
in Plateau borders can distort the radius of curvature which may influence the
Laplace pressure. Although the liquid crystals may act as a reservoir of surfactant,
they are unlikely to enhance the Gibbs–Marangoni elasticity since early studies
have indicated that the dispersed liquid crystalline phase, rather than micelles,
exhibits the diminished rates of transport to the bubble interface. This would
suggest that relatively low levels of dynamic adsorption would lead to
a lowering of foamability (74).
Novales and coworkers (75) in 2010 carried out foaming studies on fatty
acid–lysine salts in aqueous solution. The alkyl chain length was varied from
dodecyl to stearic and an attempt was made to relate the foamability to the phase
behavior. Extremely stable foams were obtained with palmitic acid/lysine salts, but
for other chain alkyl length poor foams or no foaming was reported. It was found
that the structural aggregates played an important role in foaming, and quantitative
information on the aggregate structure could be obtained using a transmission
electron microscope, differential scanning calorimetry, solid state NMR and small-
angle neutron scattering. In aqueous solution, the lysine salt of the dodecyl chain
gave an isotropic micellar solution, whereas for longer alkyl chains, the fatty acid
vesicles were embedded in a lamellar arrangement that passes from gel to fluid state
on aging or upon heating. At low concentration no significant structure has been
detected.
Shrestha and coworkers (76) in 2006 studied the foaming characteristics of
penta-glycerol monostearate (C18G5) and penta-glycerol monoleate (C18:1G5) in
dilute aqueous solution. Both surfactants showed an increase in foaming with
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
7.12 Stabilization by emulsion and pseudo-emulsion films 263
concentration and foams prepared from C18:1G5 shown to be less stable than foam
prepared from C18G5. Also, dispersions of ɑ particles and Lɑ phase could be
detected in the water-rich regions for both systems but, it was not possible to
explain this difference in foaming, and no correlation between dynamic interfacial
properties and foaming could be established. Chen and coworkers (77) showed
that the stability of foams produced from the series of low molecular weight
polyoxyethylene-type nonionic surfactants (C12EOn) could also be explained by
the formation of liquid crystals. In these experiments, the foam stability was
evaluated from the drainage profiles (the volume of liquid drained plotted as
a function of time). A wide range of profiles was detected, and the stability
initially increased with increments of concentration for each nonionic. It was
found that maximum foamability occurred at 10 w% except for C12EOn which
occurred at 1 wt%. In Fig. 7.31, the drainage volumes of foams stabilized by the
different nonionic surfactants in the series C12En are shown corresponding to 30 w%
surfactant concentration. “S”-shaped profiles are shown and three stages in the drainage
curves were identified: an initial stable stage, followed by a rapid gravity-driven liquid
drainage region which accounted for removal of 90% of the liquid and the final
thinning of the foam films.
The viscosities and microstructures were also explored; for C12EO3 and C12EO5 the
maximum stability occurred at 30 wt% structures, which was explained by the presence
of high concentrations of lamellar liquid crystals adsorbed at the interface. These were
clearly visible between bubbles, and the foams and were stable for about 20 hours. In the
case of C12EO7 and C12EO9 the stability was considerably lower; it was only for
10 minutes and no liquid crystals were detected.
Wasan and coworkers (78) investigated the influence of oil (in the micellar envir-
onment) on the stability of foam. Two different types of emulsified oil systems were
studied which consisted of (a) a microemulsion (solubilized within the micelle) and
(b) a macroemulsion. It was found that in each case the foam stability was affected
by a completely different mechanism. In the first case, where the foam film contain-
ing oil, it solubilized within the micelle to form a microemulsion and the normal
micellar interactions are changed. It had been earlier demonstrated that micellar
structuring causes stepwise thinning due to layer-by-layer expulsion of micelles and
that this effect was found to inhibit drainage and increase the foam stability.
Generally, this stratification phenomenon was found to be inhibited by the oil
solubilized within the micelle, which decreased the micelle volume (representing
a decrease in the repulsion between the micelles). In the case of the macroemulsified
oil system, the important role of the so-called pseudo-emulsion film (formed
between the air/water interface and an approaching oil droplet) in the stability of
the aqueous foaming system was emphasized (Fig. 7.32).
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
264 The stability/instability of bubbles and foams
Pseudo-emulsion film
Oil drops
Plateau
border
Emulsion film
Air bubble
Fig. 7.32 Illustration of a macroemulsified oil system. The drainage of the film may be reduced due to
accumulation of emulsified oil droplets within the Plateau borders. The formation of the emulsion
film and pseudo-emulsion film are indicated. Factors affecting the foam stability were found to be
oil volume fraction, drop size and oil phase density.
References
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
References 265
(10) M. van den Temple and R. P. van de Riet, Damping of Waves by Surface Active
Materials, J. Chem. Phys., 42, 2769–2769, 1965.
(11) K. D. Wantke, H. Fruhner, J. Fang and K. Lunkenheimer, Measurements of the
Surface Elasticity in Medium Frequency Range Using the Oscillating Bubble
Method, J. Colloid Interface Sci., 208, 34–38, 1998.
(12) R. Miller and L. Liggieri, Eds., Interfacial Rheology, Progress in Colloid and
Interface Science, Series, Brill Publishers, Leiden, Boston, 2009; L. Liggieri and
R. Miller, Relaxation of Surfactants Adsorption Layers at Liquid Interfaces, Curr.
Opin. Colloid Interface Sci., 15, 256–265, 2010.
(13) S. R. Derkash, J. Kragel and R. Miller, Methods of Measuring Rheological
Properties of Interfacial Layers (Experimental Methods of 2D Rheology),
Colloid J., 71 (1), 1–17, 2009.
(14) J. Ding, H. E. Warriner, J. A. Zasadzinski and D. K. Schwartz, Magnetic Needle
Viscometer for Langmuir Monolayers, Langmuir, 18, 2800–2806, 2002.
(15) S. Reynaert, C. F. Brooks and P. Moldenaers, Analysis of the Magnetic Rod
Interfacial Stress Rheometer, J. Rheol., 52, 261–286, 2008.
(16) G. Kretzschmar and K. Lunkenheimer, Studies for Determination of Elasticity of
Adsorption Films of Soluble Surface Active Substances, Ber. Bunsenges. Phys.
Chem., 74, 1064, 1970.
(17) K. D. Wantke and H. Fruhner, Determination of the Surface Dilational Viscosity
Using the Oscillating Bubble Method, J. Colloid Interface Sci., 237, 185–199, 2001.
(18) F. Ravera, G. Loglio and V. I. Kovalchuk, Interfacial Dilational Rheology by
Oscillating Bubble/Drop Methods, Curr. Opin. Colloid Interface Sci., 15,
217–228, 2010.
(19) R. Miller and L. Liggieri, Surface Rheology as a Tool for the Investigation of
Processes Internal to Surfactant Adsorption Layers, Disc. Faraday Soc., 129, 125,
2005.
(20) L. Liggieri, and coworkers, Eds., Drops and Bubbles in Interfacial Research, Vol.
6, 239–278, Elsevier, 1998.
(21) S. Hard and R. D. Neuman, Laser Light Scattering Measurements of Viscoelastic
Monolayer Films, J. Colloid Interface Sci., 83, 315–338, 1981.
(22) D. M. Colegate and C. D. Bain, Adsorption Kinetics in Micellar Solutions of
Nonionic Surfactants, Phys. Rev. Lett., 95, 198302, 2005.
(23) K.-D. Wantke, H Fruhner and J. Ortegren, Colloids Surf., A, 221, 185–195, 2003.
(24) A. Anderson and coworkers, Oscillating Bubble SHG on Surface Elastic and
Surface Viscoelastic Systems; New Insight in the Dynamics of the Adsorption
Layers, J. Phys. Chem. B, 110, 18466–18472, 2006.
(25) N. Schelero, G. Hedicke, P. Linse and R. V. Klitzing, Effect of Counterions and
Co-ions on Foam Films Stabilized by Anionic Dodecyl Sulphate, J. Phys. Chem.
B, 114, 15523–15529, 2010.
(26) M. J. Schick and I. R. Schmolka, In Nonionic Surfactants Physical Chemistry, Ed.
M. J. Schink, Marcel Dekker, New York, 1987.
(27) M. J. Schick and F. M Fowkes, Foam Stabilizing Additives for Synthetic
Detergents: Interaction of Additives and Detergents in Mixed Micelles, J. Phys.
Chem., 61, 1062–1068, 1957.
(28) H. L. Sanders and E. A. Knaggs, Foams Stabilized by Alkloamides in Shampoos,
Soap Sanitary Chem, 45, 1953.
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
266 The stability/instability of bubbles and foams
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
References 267
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008
268 The stability/instability of bubbles and foams
Downloaded from https://www.cambridge.org/core. CERN Library, on 11 Jan 2021 at 19:52:30, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781316106938.008