Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Advances in Colloid and Interface Science 137 (2008) 57 – 81

www.elsevier.com/locate/cis

The role of particles in stabilising foams and emulsions


Timothy N. Hunter a , Robert J. Pugh b,⁎, George V. Franks c , Graeme J. Jameson a
a
Centre for Multiphase Processes, The University of Newcastle, Callaghan, N.S.W. 2308, Australia
b
Institute for Surface Chemistry (YKI), Stockholm SE11486, Sweden
c
Department of Chemical & Biomolecular Engineering, University of Melbourne, Melbourne, VIC 3010, Australia
Available online 7 August 2007

Abstract

The use of particles as foam and emulsion stabilising species, with or without surfactants, has received great interest in recent years. The
majority of work has studied the effects of particles as stabilisers in emulsion systems, but recent successes has widened consideration into foams,
where industries such as flotation and food processing have encountered the effects of particle stabilisation for many years. This review seeks to
clarify studies into emulsions, highlighting new research in this area, and relate similarities and differences to foam systems. Past research has
focused on defining the interaction mechanisms of stability, such as principles of attachment energies, particle–particle forces at the interface and
changes to the interfilm, with a view to ascertain conditions giving optimum stability. Studied conditions include effects of particle contact angle,
aggregation formations, concentration, size and interactions of other species (i.e. surfactant). Mechanisms can be complex, but overall the
principle of particles creating a steric barrier to coalescence, is a straitforward basis of interaction. Much research in emulsions can be applied to
foam systems, however evidence would suggest foam systems are under a number of additional constraints, and the stability ‘window’ for particles
is smaller, in terms of size and contact angle ranges. Also, because of increased density differences and interfilm perturbations in foam systems,
retardation of drainage is often as important to stability as inhibiting coalescence.
© 2007 Published by Elsevier B.V.

Keywords: Foams; Froths; Emulsions; Stabilized by particles; Mechanisms; Surfactant free; Thin films interactions; Bubbles; Oil droplets

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.1. Basic chemistry of foams and emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.2. Particles in emulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
1.3. Particles in foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2. Stabilisation mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.1. Particle–interface interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.1.1. Particle detachment energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
2.1.2. Maximum capillary pressure of coalescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.2. Particle–particle Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.2.1. Air–water interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2.2.2. Oil–water interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3. Investigations into particle stabilised systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.1. Emulsion systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.1.1. Particle network structure and droplet surface coverage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.1.2. Droplet–droplet particle bridging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.1.3. Surfactant addition and synergy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

⁎ Corresponding author.
E-mail address: bob.pugh@surfchem.kth.se (R.J. Pugh).

0001-8686/$ - see front matter © 2007 Published by Elsevier B.V.


doi:10.1016/j.cis.2007.07.007
58 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

3.2.Foam systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2.1. Density effects and disproportionation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2.2. Non-adsorbing Inter-film stabilisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.2.3. Particle effects on surface tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.2.4. Particles as foam stabilisers and destabilisers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.2.5. Particles in metal foams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

1. Introduction There are also some differences that play a key role in observed
behaviour. Firstly, most gas in liquid foams are formed with
1.1. Basic chemistry of foams and emulsions bubbles up to 103 times as large as emulsion droplets, mainly the
result of the increased gas solubility in water. Also, gas bubbles
Foams and emulsions are practically important to many che- are about 105 times as compressible as oil droplets, meaning they
mical and engineering fields, and as such, investigations into the are easily deformed, which changes foam structure from an
stability, interaction and structural relationships are of great sig- individual ‘bubble’ type to typically a polyhedral type structure
nificance. Liquid and solid foams are encountered extensively in [5,6]. These are not observed for emulsions, except in some cases
the food and beverage industries (e.g. cake and beer), dermato- where the disperse phase is highly concentrated. Other significant
logy and personal care industries (e.g. shampoo), textile indus- differences, such as the higher disproportionation of gas bubbles,
tries, fire retardants, general polymers/plastics, oil recovery and the higher Hamaker constants of gas–water systems (i.e. stronger
mineral flotation processing [1]. Emulsions are also heavily used van der Waals attractive force), the different surfactant systems
in the food industry, cosmetics and paints, the pharmaceutical required for stability and of course the differences in particle
industry, agricultural products and the petroleum industry [2,3]. stabilised systems is given in Table 1 below. These features will be
Emulsions are typically categorized as oil-in-water (denoted O/ discussed throughout this review.
W), such as milk dairy emulsions, or water-in-oil (W/O), such as Foams, like pure liquid-in-liquid macroemulsions (emulsions
oil based skin creams [4]. General pure liquid foams can be where droplet size ranges from 1 to 100 μm) can be considered
considered a highly concentrated “emulsion” of, say, air and water thermodynamically unstable, as their decay results in a decrease of
(denoted A/W). Hence there are many similarities between the the free energy [15]. However, kinetic mechanisms involved with
mechanisms and underlying physical principles of stable breakdown can be so slow that the foams or emulsions can be
emulsions and foams; so the inclusive study of both fields can thought of as at least metastable for their applications [4]. The
yield results significant for a spectrum of industrial and practical primary processes for instability in foams and emulsions are crea-
applications. An example of a general similarity is the high ming and sedimentation (for O/W and W/O emulsions respective-
surface area per volume of the interface. ly), analogous to drainage in foams, coalescence and flocculation

Table 1
Similarities and differences between emulsions and foams
Interaction Foam and emulsion differences
Size of dispersed phase Foams have large gas bubbles, up to 103 times as large as oil/water emulsions, due to increased gas solubility compared to oil droplets [5]
Deformability of Bubbles are much more compressible and are easily deformed into polyhedral structures [6]
dispersed phase Only very concentrated emulsions can form polyhedral structures [7]
Main forms of instability Foam instability occurs by thin film drainage, coalescence and rupture [6]. Emulsion instability also occurs by creaming and
sedimentation and kinetics are usually slower [5]
Effects of diffusion Disproportionation in foams occurs by gas diffusion leading to coarsening of bubble size [8]
Diffusion in emulsions occurs by Ostwald Ripening which is controlled by molecular solubility, and is generally significant only in
droplets b0.1um [9]
Non-ionic surfactant Foams are generally stabilized by low HLB surface active agents (b10) [10]
effect Emulsions of O/W form are stabilized by high HLB agents (N10), while emulsions of W/O form, like foams, are stabilised by low
HLB agents [11]
Hamaker constants Aqueous foam Hamaker constant (a controlling factor in the destabilising van der Waals force) normally taken as air–water–air
A(0) = 3.6 ⁎ 10− 20 Nm [12]
Emulsion Hamaker constants are smaller, generally around 1 ⁎ 10− 20 Nm. (eg. Toluene–water–toluene A(0) = 8.6 ⁎ 10− 21 Nm) [12]
Particle stabilised Foams stabilised by particles with θ b 90° (often θcrit ∼ 70°), and are susceptible to piercing from non uniform particles [13]. Particles
systems can strongly inhibit drainage by residing in bubble interfilm, as well as form a steric barrier at the interface [8]
Particles of a greater θ range (below and above 90°) can stabilise emulsions. Emulsions more stable to piercing, and have lower natural
drainage. Hence significant stability mainly from interfacial steric barrier [14]
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 59

[5,16]. Creaming and sedimentation are gravity induced move- 1.2. Particles in emulsions
ments, caused by the difference in density of the oil and water
phases. Creaming involves the rising of oil in an O/W emulsion, Particles are naturally present in many types of emulsions, yet
whereas sedimentation involves the settling of water in a W/O despite this, the pointed use of particles either solely, or in addition
emulsion [5]. Because gravity induced movement is increased as to other traditional surfactants as emulsion stabilisers, has only
differences in density become more pronounced, water drainage been recently (in the last 20 years) given extensive attention. In
(and gas “creaming”) is most significant in foams. In fact, interfilm addition to the many recent developments, published study of
drainage is the basic driving force forming the ‘dry’ polyhedral particles in emulsions has been documented for 100 years. The
type structures in foam systems, which in turn heightens conditions person credited with first investigating particle stabilisation in
for bubble coalescence [6,15]. Coalescence involves the irrevers- emulsions is Pickering, who in 1907 (using work documented
ible binding of two or more foam bubbles (or emulsion droplets), four years earlier by Ramsden [22]) noted that particles more
where the interfacial film drains and is eventually ruptured forming wetted by water than oil, stabilised O/W emulsions by residing at
a single larger bubble, and is generally the most severe form of the interface [23]. From this report, the term ‘Pickering emulsions’
instability [17]. Flocculation involves interaction between emul- has been given to emulsions stabilised by solid particles.
sion droplets (and can also be observed in ‘dilute’ foam systems, The role of particle wettability and emulsion stability was
such as the initial bubble zone of a flotation column) where the free discussed further by Finkle et al [24], who considered that
energy minimum is appreciably low for the droplets to interact. particles at an interface in an emulsion would most likely pre-
Unlike coalescence, the interfacial film wall does not drain to ferentially reside in one of the liquids, and this would become the
rupture, and the droplets remain held together as ‘clusters’[5]. disperse phase. This was considered in line with resulting
In foam systems, another form of instability that is important is emulsions from surfactant adsorption. It was not until the work
disproportionation. Disproportionation is the process whereby air of Schulman and Leja [25], that particle wettability and interface
molecules diffuse through the disperse phase, between bubbles. contact angle was comprehensively investigated. Using barium
Because of the increased Laplace pressure of smaller bubbles, the sulphate crystals and surfactant, they found that conditions
diffusion flux generally results in the shrinking of smaller, and the giving particle contact angles (θ measured through the aqueous
growth of larger gas bubbles [5]. Coarsening of the bubble size phase) slightly below 90° resulted in O/W emulsions, and for
distribution from disproportionation, can also lead to increased conditions giving θ slightly above 90°, W/O emulsions were
drainage and coalescence [8]. This diffusion process is also formed. Interestingly, conditions giving extreme contact angles
present in emulsion systems; termed ‘Ostwald ripening’. As it is (close to 0° or 180°) no stable emulsions were formed.
controlled by molecular solubility, it is negligible for immiscible The wetting behaviour of solid particles is generally
fluids, such as most oil and water emulsions, and is usually only considered a key descriptive rule for particle emulsion beha-
significant for droplets under 0.1 μm [9,17]. viour. Consider a particle at an oil–water interface. For a θ
To help stabilise foams or emulsions of pure components, it is measured through the aqueous phase of below 90°, the particle is
normally necessary to add an additional third component, a preferentially wetted by the water (i.e. it is preferentially hydro-
foaming/emulsifying agent. In, say, a traditional air–water philic in nature). For a particle θ of above 90°, it is preferentially
system, foaming agents are required to first produce conditions wetted by the oil (i.e. hydrophobic). For a contact angle of 90°, it
that will create a foam, as well as stabilise it for required time is equally wetted by both phases, and is considered at the point of
periods, as simply, “clean” air bubbles passed through water will inversion [14]. Similar to surfactant behaviour, the preferentially
burst immediately on drainage [15]. Foamers/emulsifiers are wetting phase becomes the disperse phase, as the particles
classically molecular surfactants (e.g. fatty acids or alcohols [18]), minimise energy by curving round the inhibited phase, forming
polymers or larger protein type agents (e.g. egg albumen) [19]. stable emulsion droplets [26,27]. For particles with contact
Molecular surfactants generally contain a polar (hydrophilic) angles near 90°, whether O/W or W/O emulsions result depends
head group and a non-polar (hydrophobic) chain tail. Surfactants on the particle and solution properties such as concentration
therefore preferentially adsorb to the air/oil–water interface. This [28,29]. Fig. 1 shows how the bending behaviour of droplets
reduces the free energy involved with producing a high surface coated in particles may dictate emulsion formation.
area interface, and as a result, reduces the interfacial surface There have been a number of different particle types used as
tension [4]. Polymers and proteins cause stability largely through stabilisers in both O/W and W/O emulsions, including silica, latex
electric and steric repulsion, controlled by the extent of unfolding particles, metal oxides and sulphates, clays and carbon. The
(or ‘denaturing’ as unfolded proteins are known) and conforma- effectiveness of a specific particle type in stabilising an emulsion,
tional layer structure on droplets [20,21]. ‘Semi-dilute’ polymers depends on the emulsion medium, the particle shape and size,
and larger ‘globular’ proteins, which do not denature to the same particle wettability and inter-particle interactions [14]. Particles
extent, can further cause stability through changes to the rhe- impart stability on emulsions primarily through a steric barrier,
ological properties of the dispersion medium (namely through created at the interface. With a certain concentration of particles, a
increase in viscosity) [5]. This may also be achieved through close packed network is generally formed, which further increases
addition of agents such as glycerine [18]. Of course, many natu- stability. It is also thought that a certain level of particle
rally occurring emulsions and foams will contain a variety of flocculation is advantageous for stability [30]. In addition, at
types of agents, leading to potentially very complex interactions certain concentrations, particles may affect the rheological pro-
giving overall characteristics. perties of the disperse medium (much like free-proteins), and
60 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

Fig. 1. Bending behaviour of emulsion droplets coated in particles.

particles such as clays, can also act to increase stability by for- stabilised by asphaltene particles [39,40]. In waste water, foams are
mation of inter-droplet networks [31–33]. Unlike surfactants, stabilised by colloidal particles such as bacteria, soil and viruses. In
particles do not affect emulsion stability by significantly reducing other cases, heterogeneous nucleation of insoluble precipitates
the oil–water surface tension [34]. (hydrolysed cations such as iron hydroxides) at the air/water
There is a large size range of particles that can be used to surface can also cause foaming in waste water or rivers. Many other
stabilise emulsions. For successful stabilisation, it is necessary everyday processed products frequently involve particle stabilised
that the particles be approximately orders of size smaller than the foam systems, which ensure long-term stability. In food and drink
droplets, for the particles to be properly located around the production (bread and beer have foam like structure) and in the
droplets [2]. The actual size of particles that can be used ranges dairy industry when whipping cream, partially crystalline ‘solid’
from small nanometre [35] to micrometer [36] to successfully oil droplets accumulate at the interface and stabilise the foam [5].
stabilise emulsions. The size of the particles does correlate to the It is also important to note that the adsorption of particles at
size of the droplets formed in stable emulsions, with even droplets bubble interfaces is the basic driving mechanism in many
of up to millimetre having been found stable to coalescence, dynamic foaming processes, such as dissolved air flotation to
something not easily possible with surfactants [14]. Generally the treat waste water, effluent treatment and froth flotation [41]. In
overall stability is inversely proportional to particle size, with the froth flotation process, which is widely used as a particle
smaller particles giving a higher packing efficiency, and so separation method for minerals, the attachment of particles to
producing a more homogenous layer [16,35,37]. bubbles occurs when the particles are dispersed in the liquid
Because of the extensive attention particle stabilised emul- pulp [42]. The bubble/particle aggregates then ascend to the
sions have gained in recent years, behaviour has previously interface where the behaviour of the particles has a critical
been well studied and documented, as evidenced from reviews influence on the stability of the thin films in the froth. It is of
by Binks, Aveyard and co-workers, from the University of Hull interest that much early work on particles as foam stabilisers has
[14,26] and a recent book covering this topic edited by Binks & occurred, from a perspective of mineral processing and flo-
Horozov [38] . Because of this, an effort will be made to focus tation. Throughout industry, there are many examples of particle
on presenting recent work on the subject of particles in emul- stabilised foaming processes in both aqueous and non aqueous
sions, while maintaining required continuity. The second and environments, with and without surfactant addition. For exam-
major focus of this report is to link well detailed emulsions ple, during the boiling of radioactive sludge suspensions, foam-
behaviour, to the study of particles in foams, a subject that has ing has been reported to frequently occur [43].
been less extensively investigated, but currently is an area of There has been published material, in which the role of particles
increasing interest. as stabilising species in foams has been noted, dating back to 1913,
where Hoffmann suggested finely divided particles had a role to
1.3. Particles in foams play in stabilisation of foams or froth, specifically in mineral
frothers and ore flotation ([44] as cited in [18]). In 1925, Bartsch
As with particles in emulsions, the phenomena of particle found partially hydrophobic particles stabilised froths, whereas
stabilisation of foams, has only recently been given pointed completely wetted particles had no effect on the stability [45].
attention, and interest is largely the result of successes in emulsions. Similarly Hausen found stable three phase froths formed with
Yet, it is well known that finely dispersed particles play an various metal oxide and clay particles [46], and Lekki and
important role in the stabilisation of many different types of foams. Laskowski found that particles could enhance the stabilisation
For example, particles in the absence of surfactant are known to power of frothers such as diacetone and ethyl acetal [47]. Ottewill
cause foaming in rivers, treatment of radioactive wastes, dispersed et al., in a non-aqueous system, reported increased foam stability in
sludges, distillation towers, oil-well drilling, pulping in the paper systems of alkali metal carbonates, that was attributed to increase in
industry, fabrication of cellular metal foams and the preparation of bulk viscosity of the dispersion with solid particles [48]. More
foods, etc. In the oil industry, obnoxious foams are produced in recently, Tang [49] showed that fine silica particles enhanced
boilers and various stages of distillation and are thought to be stability of an SDS foam, and that stability was proportional to
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 61

particle concentration, and inversely proportional to particle size frothers, it is difficult to first establish frothing conditions in clean
(much like emulsion systems). particle–water systems. So collectively, the increased bubble–
This initial evidence would signify that many similarities lay bubble contact area from drainage drying, deformability, and
in the stabilisation mechanism of particles in foams and emul- inhibitive effects of hydrophobic particles means particle
sions. The reliance on particles to create a steric barrier to coal- stabilised foams are under a number of additional constraints
escence is again a major contribution to foam stabilisation, and as when compared to emulsions.
with emulsions, smaller particles in high concentrations that form Further research into the finite stabilisation mechanism of
a more complete layer gives the most effective barrier. Because of particles in foams may possibly yield results of great significance,
the increased effects of drainage in foam systems, the affect of and is in some ways, a more complex proposal than in emulsions.
particles on the interfilm (for example, causing retardation of Overcoming the constraints of stable foam behaviour may lead to
drainage and increasing maximum capillary pressure) is also sig- many potentially profitable investigations, using particles in
nificant in some systems, and generally is a more important addition to, or as a way of completely removing surfactants from
mechanism than in emulsions. Bridging of particles between a foaming process. Particles can prove both economically and
bubbles is also a very important process, because of the high environmentally attractive to industry. This is because particles
forced contact area in foam phases, as they become dryer [6]. Fig. 2 may cause stability over and above generally used surfactants
shows how particles may stabilise polyhedral type foam. and also because of the benefits of reducing dependence on
Hydrophobic particles (of contact angle greater than 90°) potentially harmful organics. Knowledge of the stabilisation
should not work as effective stabilisers in foams, obviously as process is especially important to the flotation industry, where
foams will break (collapse), upon opposing bending energies, as accurate prediction of mineral yields is a difficult prospect, and
there is no ability for foams to ‘invert’ (although, recent evidence the great range of particles often encountered leads to many
suggests so-called free-flowing liquid marbles and powders possibilities in overall foam structure and so flotation potential.
containing numerous water drops, termed ‘dry water’, may be
formed [50]). In fact, hydrophobic particles can pointedly be used 2. Stabilisation mechanisms
as foam breakers, further destabilising the foam from bridging
drainage, as comprehensively investigated by Dippenaar [51– 2.1. Particle–interface interactions
53]. Also, because of the deformability and the ‘delicate’ nature
of foams, they can be subject to foam destruction from particle 2.1.1. Particle detachment energy
piercing, especially in the case of larger, non uniform, particles A useful factor that theoretically links how well particles of
(an environment often encountered flotation). Further, because of different contact angles stabilise emulsions and foams is the
the weak affect on surface tension without the use of other particle detachment energy. This detachment energy is related to
the free energies involved in removing an adsorbed particle from
an interface (either oil–water or air–water). As particles create a
steric barrier to coalescence, it is obvious that strong particle
detachment energy will result in more force being required to
disrupt the particle layers and allow coalescence [54]. The
detachment energy becomes higher, as generally an area of great
line tension (oil/air water interface) is lost due to particle
adsorption, where the energy is influenced most significantly by
the area (i.e. its radius ‘R’) and immersion of the particle at the
interface (i.e. θ) and the initial surface tension γow/γaw [26].
Consider a single spherical particle moved to an air/oil–water
interface (Fig. 3), with an equilibrium immersion depth and
contact angle θ [54].
The energy required to move the particle from this equili-
brium into the bulk water can be calculated from fundamental
principles (if buoyancy/gravity effects are neglected) [16,27].
Taking into account the area and depth of immersion, and
respective surface tensions, this detachment energy (ΔGremove)
is shown.

DGremove ¼ pR2 gOW=AW ð1FcoshÞ2 ð1Þ

This energy is considerably large (in the order of 103 kT for a


10 nm particle) at angles of around 90°, but falls quite rapidly as
the particle contact angle decreases or increases. At θ b 30° or
θ N 150°, the contact energy is essentially negligible, suggesting
Fig. 2. Particle stabilised foam. that in principle these particles would not create stable emulsions
62 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

or foams. Another important consequence of the large energy of via a number of pathways, from chemical grafting [59], seed
attachment, is that once at an interface, a particle can be thought polymerization [60], to methods building particles of hydropho-
of as irreversibly absorbed, a behaviour unlike surfactants, bic cores and hydrophilic shells [61], and the use of composite
which are generally thought to be in a state of dynamic equili- materials [62]. The manufacture and use of these particles in
brium, absorbing and desorbing on a fast timescale [26]. The various systems, for example, in cosmetic emulsions, has pro-
energy of attachment is important, especially in determining duced a number of patents, and there is little doubt, interest in the
overall stability relationships, but many factors lead to it being properties of amphiphilic particles will continue.
useful generally only in a qualitative sense. A significant
mitigating factor in quantitatively assessing the relationship in 2.1.2. Maximum capillary pressure of coalescence
real systems is the great difficulty in accurately determining Although detachment energy theory can explain much of the
particle contact angles, especially for particles in the nanometre overall stability phenomena observed in emulsions, it is often
size range. Other problems, such as particular contact angle found that in foam systems (as described in detail in Section 3.2),
hysteresis in certain phases, and effects of irregular meniscus and the contact angles for maximum stability occur below 90°, and
surface roughness on clean particle wetting, make very accurate particularly in the region of 60–70°. One mechanism that may
assessments almost impossible. explain this behavior is the affect of particles on the capillary
A theoretical comparison can be made between the pressure between two bubbles. Thus, instead of looking at how a
detachment energy of uniform particles, with that of true am- particle is attached to a single interface, one considers how
phiphilic particles. It is possible to partially treat particles, particles residing between two interfaces affect stability by
creating both a hydrophobic (apolar) section, and a hydrophilic reducing the thinning of the draining interfilm. In particular we
(polar) section on the particles. This more truly mirrors attributes consider the pressing force required to bring two bubbles to
of molecular surfactants, which are naturally amphiphilic in coalescence (i.e. to reduce bubble-bubble distance to approxi-
character [55]. These types of particles have been termed ‘Janus’ mately zero) with particles holding them apart. This pressing
particles, a description originally coined to denote particles with force is described as the capillary pressure, or the pressure
equal parts hydrophilic and hydrophobic surface [56], yet, with gradient between the bubbles (P1) and the interfilm fluid (P2)
modern synthesis techniques, the fractional distribution of the (analogous to the ‘disjoining’ pressure) [63].
polar and apolar sections can be varied [57]. Amphiphilic parti- For different arrangements of particles (if it's assumed they
cles can also exist naturally in some systems, such as radioactive remain embedded and stationary between the enclosing bubbles)
waste sludge, which consists of noble metal oxides and hy- it is possible to calculate a theoretical maximum capillary pres-
droxides of broad chemical composition that are often bi-philic. sure sustainable by a given system. A higher maximum capillary
These particles are problematic, as they create very stable, pressure (Pcmax) indicates a more stable system. Ivanov and co-
unwanted boiling foams [43]. workers were the first to consider such as system [64], and
As with uniform particles, theoretical attachment energies can models have been further developed by others including Nush-
be related to contact angles. Binks and Fletcher [55], using the taeva & Kruglyakov [65] and Kaptay [66]. A recent review by
work of Ondarcuhu et al. [58], considered the energy rela- Kaptay [67] has brought this mechanism back into focus, and it
tionships for amphiphilic Janus particles. The detachment energy will be briefly surmised here.
is increased significantly with increasing amphiphilic behavior, Kaptay [67], derived an analysis for a single hexagonal layer
and is as much as three times that of a homogeneous uniform of particles between two bubbles (or droplets); in a continuation
particle. Janus particles could therefore lead to marked increases of the work of Visschers and co-workers [68]. Here particles
in particle stabilised emulsion and foam performance. With the with a zero contact angle (and radius R) that are completely
many recent developments in chemical manufacture, there has resting in the interfilm, lie between the two droplets as shown in
been extensive recent interest in the production and use of Fig. 4. For this system, if the capillary pressure is zero (i.e. no
amphiphilic type particles. Modern methods allow production net ‘sucking’ pressure between the droplets) then the contact
line between the particle and bubbles is flat (irregular meniscus
profiles are not considered). As drainage occurs, the bubbles
form a meniscus around the particles as shown. The curvature of
the meniscus is given from an assumed sphere radius ‘r’. This
increases the capillary pressure between the bubbles. The con-
tact angle between the meniscus profile and particle tangent is
given by Θ, while the contact area is derived from the angle α.
As drainage increases, the meniscus profile continues to
curve around the particles, and the overall separation of the
bubbles ‘H’ decreases. As the capillary pressure continues to
increase accordingly, the bubbles will eventually touch, as H
decreases to zero, and the pressure associated with this coale-
scence event is given as Pcmax. The basic equation of maximum
capillary pressure can be resolved by equating this system to the
Fig. 3. Particle at an air/oil–water interface. Laplace equation, and combining the various parameters
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 63

Fig. 4. Particles residing in a droplet–droplet interfilm effecting Pcmax for coalescence [67].

associated with the changing structure of the interface. This is interfaces in drying polyhedral foams, occasions where laterally
given by Kapkay [67] as shown in Eq. (2). static interfaces are bridged by particles is more common than in
emulsions, and the proposed capillary mechanism more aptly
2gAW=OW applies.
Pcmax ¼ FP cosh ð2Þ There are some general shortcomings to the theory of
R
capillary pressure however. Perhaps most important is that
Here, as with particle detachment energy, a positive value particles are themselves assumed to be statically located and not
relates to bubble and O/W systems, while a negative value relates influenced by fluid drainage or flow. In effect, even particles
to W/O systems. The factor ‘P’ is a theoretical packing parameter, with almost no adsorption are not assumed to be ‘washed away’
which by various approaches is used to associate the influence of from the contact area by liquid drainage. Also, the effect of
particle concentration and structure on the capillary pressure. The particle–particle interactions is important. Kruglyakov et al. [69]
other important implications of this equation are firstly that and Slobozhanin et al. [70] found in experimental studies, that
particle size ‘R is directly and inversely proportional to Pcmax, Pcmax values followed the general stability trend, but were much
showing smaller particles with higher curvature and smaller lower than theoretically predicted. The flocculated nature or
interparticle packing are more effective ‘blockers’. Secondly the incomplete structure of particle networks was assumed to be a
maximum capillary pressure is directly affected by contact angle. key mitigating factor, highlighting the importance of categoriz-
This predicts highest Pcmax values for contact angles close to ing such mechanisms, and why generally, equating system
zero, and lowest values (or lowest stability) at contact angles stability to only particle–interface interactions often only gives
around 90°. This is an opposing trend to detachment energy, but half the picture.
here makes sense as obviously the more a particle resides in the
interfilm, the more the liquid film has to drain around it (and H 2.2. Particle–particle Interactions
reduce) until coalescence. In fact, Kaptay has looked at the
combined effect of detachment energy and capillary pressure on The theoretical understanding of particle–particle interactions
calculating contact angles of high stability. Assuming a relative at interfaces are complex, and still leave unexplained much
change in energy for each competing mechanism, he associated phenomena observed experimentally [14]. However, the princi-
the combined relationship as shown in Fig. 5 for a single layer of ple of particle network structures generating a steric barrier is
particles between two bubbles [67].
This relationship highlights the potentially significant role
that capillary pressure can have on a given systems stability
performance. Most interesting however, was how the contact
angle for maximum stability theoretically changes when a
system of two particle layers is considered (two completely
coated emulsion droplets coming into contact). Here, Kaptay
found the contact angle of maximum stability was located at 85°
[67], well within the region found often with emulsions [14],
where these types of high concentration systems are encounted.
In terms of foams, often lower concentrations of particles are
encountered (especially in flotation [41]) so this relationship
may be even more critical. Also, the increased drainage in foams Fig. 5. Combined effects of Pcmax and ΔGremove on the relative energy of particle
will add to its relevancy. Because of the high contact between stabilised systems versus particle contact angle [67].
64 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

simple, and can be surmised in terms of the nominal steps in There has been some conjecture on the role of dipole–dipole
droplet coalescence. For coalescence to occur, film drainage and repulsion forces in air–water systems. Abdel-Fattah and El-Genk
thinning must occur between bubbles (or emulsion droplets). In [72,73], and Williams and Berg [77] called into question the
the final stages of film thinning, dark spots (or holes) are formed, strength of dipole–dipole forces, with work on particle adsorption
which expand until the film becomes unstable and ruptures [27]. to air–water interfaces. Both found that at even moderately low
If a droplet or bubble surface is well covered in a stabilising salt concentrations, conditions led to extensive particle aggrega-
species, an energy barrier must be overcome to form a critical tion on the surface, at salt levels up to two orders of magnitude
sized hole in the film wall. Thus the rate of coalescence can be below bulk conditions [77], and below concentrations expected if
considered in line with the energy required for hole formation significant dipole–dipole repulsions were evident. Aveyard and
[17]. Because of the high free energies involved with strongly co-workers also found unexpected aggregation of particles at air–
adsorbed particles [54,71] (as discussed in Section 2.1.1), they water surfaces [78], but especially in this case, very low particle–
are far more likely to be laterally moved along the contact water contact angles (close to 0°) may have lead to the observed
interface, rather than expulsed into the open medium during film trends (as particle–interface contact energy was therefore
thinning [33]. Hence, particle–particle forces such as electric minimal).
double layer repulsion and dipole–dipole repulsion, as well as Other workers, using more hydrophobic particles have found
van der Waals attraction and capillary forces are of utmost surprising stability of particles at air–water interfaces, and have
importance to overall stability of foams and emulsions, and may established that dipole–dipole forces are indeed significant
dominate interactions over particle–interface attachment. [79,80]. Further, any observed aggregation can be accounted for
without disregarding such repulsions, by equating other attractive
2.2.1. Air–water interface forces [80–84]. Robinson and Earnshaw [79], showed that by
Equating the role of various forces in interface systems is a decreasing particle contact angles (rendering particles more
complex proposal, but may be achieved by breaking down the hydrophilic) dipole–dipole forces were reduced in line with
overall interaction potential into a number of defined mechan- increased counter ion screening, and aggregation was increased.
isms and their related activities. Firstly, once a particle is attached Ghezzi, et al. [80] further showed that by applying charge to the
to an air–water interface, particle–particle DLVO repulsion air interface, aggregation was induced by retarding dipoles
works through the aqueous phase, and augmented van der Waals through counter ion screening in the air phase. They also found
attraction working through the aqueous and air phases [72,73]. that under certain conditions, particles aggregated into complex
So, as in bulk colloidal conditions, effects of surface change and chain and meso-structures, an interesting result they attributed to a
salt counter ions can increase repulsion or cause interfacial secondary energy minimum (attraction force) at long-range
coagulation. Levine, from the theoretical considerations, deter- distances (beyond possible van de Waals attraction) [80]. This
mined that double layer repulsion was the most significant secondary long-range attractive force was also independently
repulsive force, and dipole–dipole repulsive energies were small established by Ruiz-Garcia and co-workers [81,82].
(in the range of ∼ 0.8 kT) [54,71]. However, Levine also felt that The origin of this secondary attractive force is debatable.
an additional force was required as a thermodynamic equilibrium Ghezzi [80] felt that the secondary minimum was caused by
balance with the single particle energy well. Further, current inhomogeneous charge distributions around the particles,
consensus points to interactions, such as dipole–dipole repul- leading to certain particle alignments, whereas Ruiz-Garcia et
sion, becoming significant in systems where particles contain al [81] and Meija-Rosales et al [84] felt that the secondary
ionisable surface groups and a medium of low dielectric constant minimum could be predicted even with isotropically charged
is used (i.e. air) [74,75]. particles, by taking into account interaction potential oscillations
Dipole–dipole forces were first proposed by Pieranski [76], with particle distance, from the dipole–dipole repulsions.
who determined that dipoles are produced, by asymmetrical Stamou and Duschel, considered the secondary minimum to
charge distribution on particles, as the portions of particles in the be caused by capillary attraction [85], and this is currently the
aqueous phase become ionized [54,76]. In effect, for particles most popular theory. It was felt that particle surface roughness
with ionisable surface groups (e.g. latex or silica), the part of the could lead to irregular shaped surface meniscus around particles,
particle immersed in the aqueous phase will become charged,
creating asymmetric charge distribution on the particle, and
causing a dipole perpendicular to the interface [75,76]. The
asymmetrically charged particles undergo partial neutralisation,
from counter-ions in the aqueous phase. Particles with smaller
aqueous immersion, will attract a lower concentration of dis-
solved counter ions, and the dipole will work normally through
the air phase, causing repulsion between particles [75]. The
dipole–dipole forces are considered to work in association with
DLVO double layer repulsion through the aqueous phase (see
Fig. 6). Also, as charge neutralisation is controlled by the extent
of aqueous immersion, more hydrophobic particles will attain
stronger dipole–dipole interactions. Fig. 6. Dipole–dipole formation on particles at an interface.
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 65

Fig. 7. Possible causes of meniscus deformations and resulting capillary attraction [92].

which could lead to an attractive capillary interaction, from considered, along the lines of Stamou and Duschel [85] that
particle re-alignments forced by complex meniscus associations. irregular particle wetting was possibly the major cause of
Whatever the attraction basis, it has been shown through film meniscus associations in most systems [92]. Fig. 7, shows
balance and computation analysis, the secondary minimum these potential origins of capillary attraction for both large and
cannot be attributed to DLVO interactions or hydrophobic small particles.
attraction [83,86]. The basis of an irregular meniscus on small particles for a
It is perhaps pertinent at this point, to more fully assess the capillary interaction was first proposed by Lucassen [94]. He
origins and strength of general capillary interactions in modelled the effects of a small particle, with a sinusoidal edge
interface systems. The capillary force at work is a lateral disturbance (undulated contact line), and showed there could
interaction force. It is present in any systems where the particles be significant interaction with neighbouring particles. The
cause surface deformations (menisci), and generally (depend- interaction is based on the minimization of the total surface
ing on shape) the overlap of menisci between two close area, and could result in a capillary force, acting over distances
particles causes an attraction [87]. The origin of the initial of the order of the sine wavelength, and magnitudes pro-
deformations is dependent on particle size. For larger particles, portional to the square of the amplitude. The model showed
the deformation of the menisci is produced from gravitational that, in cases where there is a difference in amplitude or
effects, but for particles below 5 μm, gravity effects are wavelength, or when the sines are out of phase, an attraction at
negligible, and behaviour cannot be accounted for in this way large distances resulted, which changed into repulsion as
[88]. Nikolaides [89] considered that small particle meniscus particles approached on another. This theory wa further de-
deformations were caused by electric field induced fluctuations veloped by Stamou and Duchel [85] and by Kralchevsky [92].
to the interface, from the potential differences between the Here it was assumed that irregular wetting causes changes to
aqueous and non aqueous phases, resulting in a net long-ranged the fractional area affected by charge neutralisation, which
attractive force. Questioning work from Foret and Wurger [90], causes further asymmetric charge distributions. In effect the
and Megans and Aizenberg [91] suggested electric field particles dipoles augment to form a quadrupolar type charge
differences could only result in a small negative (repulsive) distribution, and can then align themselves according to the
perturbation, and could not be the cause of any capillary placement of the quadrupoles (Fig. 8), in an attempt to mi-
attraction in small particles. It must be noted that in a reply, nimise energy. This could account for the particle rearrange-
Nikolaides, stated that potential fluctuations may still lead to ments, and meso chain like structures evidenced by Ghezzi
capillary attraction, if trapped ions evidenced in O/W interfaces [80].
are present [91]. Again, this could not be a potential cause in From the quadrupolar theoretical basis, it has been shown
air–water systems, where such trapped ions do not occur (see that surface irregularities can cause remarkable strong attrac-
Section 2.2.2). tions relative to kT [85,92,94]. Other authors have also
The general interaction phenomenon of capillary attraction assessed the aggregation patterns of particles subjected to
has been a focus of research for Kralchevsky and co-workers quadrupolar capillary interaction. Brown et al used a novel
[88,92,93]. They have detailed two possible causes of required etching technique to produce various particle types with noble
meniscus deformations in small particles. Firstly, immersion in metals and silica dioxide [95]. Using disk shaped particles,
three phase systems (i.e. thin film studies on flat plates) with both a hydrophilic and hydrophobic face, surface
meniscus deformations, and resulting capillary attraction, can deformations and cluster formations were studied, and
be formed from restricted surface wetting. They also generally could be approximated by considering quadrupolar
66 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

Fig. 8. a) Quadrupolar charge on particles with an undulated contact line, b) resultant particle realignment in reduced energy conformations [92].

analysis. The conformations were successfully modelled, and water systems [12] may account for some reduction in
calculating energy relationships, in terms of two and many- attractive aggregation. The same principles of dipole–dipole
body interactions by Fournier and Galatola [96]. repulsion are also evident, working through the oil phase.
Capillary attraction is still evidenced in certain systems, as
2.2.2. Oil–water interface observed by Kralchevsky and co-workers, finding both
Forces at work in oil–water systems show many of the theoretical and experimental evidence of a capillary force
same characteristics to air–water interfaces, but with a small sufficiently strong enough to cause partial aggregation at O/W
number of important differences. As with air systems, DLVO interfaces [87,93]. Further, Loudet et al used a model O/W
forces will again work through the aqueous phase, and interface to study the importance of particle shape on capillary
strength is dependent on the extent of particle immersion. attraction, by looking at ellipsoidal silica and latex particles
However, the weaker Hamaker constants present in most oil [97]. In such cases it seems the capillary attraction is enhanced
along the quadrupolar lines, producing very long-range
particle chains. Also, chaining conformation was proven to
be a direct result of particle hydrophobicity, with ellipsoidal
silica chaining almost exclusively side to side, while lenticular
latex particles formed end to end chains (see Fig. 9). The
strength of such capillary attraction is in fact remarkable for
latex spheres, considering the repulsive forces at work in oil
systems.
In many systems of very hydrophobic particles, capillary
attraction is greatly reduced, or at least is dominated by a
further force in oil–water systems. Apart from short range
double layer and long-range dipole–dipole forces, there is also
evidence for an additional ultra-long-range repulsive force.
Aveyard and workers [78] observed that latex particles in an
octane-water interface produced extremely stable networks,
even at high electrolyte concentration. This stability could
therefore not be fully attributed to dipole–dipole forces, as
these are thought to decrease significantly at high electrolyte
concentration (from increased neutralisation from counterions)
[78]. From the experimental evidence, they deduced that the
origin of this repulsive force lies in charge–charge repulsion
from the surface sulphate groups. It is thought that a small
percentage of the surface charge groups are trapped by surface
water molecules in rough surface pores, and in the oil phase act
in unscreened charge–charge repulsions [98]. Molecular
surface dynamics simulations were also fitted successfully to
models including this charge–charge force [99], and further
studies with hydrophobic silica produced similar ultra-high
order packing (from trapped hydroxide ions) [100]. In
addition, there is no evidence of these ‘ultra-long-range’
repulsive forces in air–water interfaces. This confirmed the
idea that water-trapped charged ions are the cause of the
Fig. 9. Aggregation patterns of A) ellipsoidal silica (scale bar 22 μm) and repulsion, as these groups would not remain trapped in an air–
B) ellipsoidal PS (scale bar 33 μm) at O/W interface [97]. water system [78].
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 67

3. Investigations into particle stabilised systems detachment energy. Major differences in packing structure
were observed between hydrophilic and hydrophobic silica,
3.1. Emulsion systems seen under freeze fracture SEM (Fig. 10). Hydrophilic silica
bonded to the interface as individual particles (no evidence of
Previous research has focused on particle emulsion sta- pre-flocculation); whereas the hydrophobic silica produced
bilisation with a view to ascertaining the inherent mechanism highly flocculated mono and multi-layers, even at salt
of stabilisation, including determination of the structure of concentrations orders of magnitude below bulk flocculation.
interparticle networks at the interface, conditions that give This would show either evidence of an additional hydrophobic
optimum emulsion stabilisation and also ways to augment attraction working through the bulk phase [17], or that because
emulsions for specific purposes. Work has been undertaken to of the increased particle attraction to the interface, closer
investigate aspects such as determination of particle to droplet interparticle distance on contact (in relation to the bulk) leads
size relationships, comparing particle types and performance, to increased flocculation.
mixtures of particles and resultant effects, oil phase polarity Horozov and Aveyards work on aggregation at static
and type, optimum percentage droplet coverage, and inversion particle interfaces found opposite trends [100]. Using silica of
principles. These systems were comprehensively studied in varying hydrophobicities, they found more hydrophilic silica
the review by Aveyard [14]. The focus here will be on produced disordered aggregated networks, and very hydro-
important areas that can be compared to foam systems, and phobic silica produced a well ordered hexagonal network, in
areas highlighting the complex questions remaining in particle line with the apparent charge–charge repulsive force.
emulsion behaviour. However, there were a number of important differences
between experimental situations. Firstly, Horozov and Ave-
3.1.1. Particle network structure and droplet surface coverage yard found the well-ordered networks were formed at
The types of singular and flocculated network structures that
can be observed in emulsions (and indeed foams, although less
studied) is vast, and very hard to categorise. The role of
particle–particle interactions can give a key indication of
expected structures, however the great energies involved in
emulsion formation may counterbalance this in some systems.
Insight has been gained from both the study of static network
interfaces, where detailed observations and analysis can be
made to degrees not possible with dynamic systems, and
emulsion interfaces (where often information can only be
gained from TEM/SEM data).
Stancik et al investigated the strength of close packed
particle networks, by subjecting particle coated surfaces to
extensional flow [75]. It was found that networks resisted
conformational changes, but under certain flow, elongated into
rhombic structures. Under the studied close-packed conditions,
the networks produced significant stability, yet with a finite
limit. Aveyard et al also found similar rhombic conformational
changes with particle networks in a Langmuir trough subjected
to compression, and showed that particle systems would ‘fold,’
when a certain collapse pressure is reached, holding arrayed
structures (particles remained at the interface, not expelled into
the open medium), indicating again significant particle
network stability [78,100,101].
Simovic and Prestidge have completed a number of studies
into the structure and behaviour of particle networks,
subjected to a series of conditions, and described a number
of differences between hydrophilic and weakly hydrophobic
silica in terms of stabilisation characteristics [30,102]. Model
polydimethylsiloxane (PDMS) oil droplets were used, which
could be synthesised to high monodispersity (and extent of
droplet cross-linking manipulated). As expected, for both
particle types, increased salt concentration lead to increased
packing (in line with decreased DLVO repulsion forces Fig. 10. A) SEM (high magnification) of hydrophilic silica on PDMS droplets;
between particles) and hydrophilic particles performed poorly, B) SEM (low magnification) of hydrophobic silica on PDMS droplets, both with
in line with the very low contact angle and attributed 10− 3 M NaCl [102,103].
68 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

hydrophobicities above those studied by Simovic and hibiting Brownian motion while trapped at the oil interface.
Prestidge (contact angles above 118°). Also, differences This would further suggest that below concentrations causing
between a static interface and real emulsion systems (studied packed networking, individual particles can easily move
by Simovic and Prestidge) cannot be discounted, as energies laterally around the droplet surface. There was also evidence
involved in emulsion formation would introduce far more of particle clustering, attributed to capillary attraction forces at
particles to the oil interface, possibly causing additional the interface. Fig. 11 shows the range of network behaviour
aggregation for even particles in strong networks. These types seen for different particle concentrations. Most surprisingly,
of effects may be a mitigating factor in the general science of they discovered that when two oil droplets were brought into
equating interface systems to real emulsion and foam systems. contact, particles would collect around the drop-drop contact
Effects of changes in oil phases and salt conditions could have region, thus still imposing a significant steric barrier (i.e. the
further led to the apparent structural differences. particle distribution was non-homogeneous, and was affected
Midmore discovered interesting structural packing differ- by neighbouring droplets).
ences between systems of silica and poly-ethylene oxide (PEO) The range of structures observed with low concentration
or hydroxypropyl cellulose (HPC) surfactants [104,105]. In systems pulls into focus the general problem of defining
PEO-silica systems, silica networked at the oil drop interface to particle networks. In many emulsions, and also in foams (see
coverage of approximately 72% monomer layer, an average Section 3.2) many authors have noted potential aggregated
packing efficiency, and expected for stabilised emulsions. With ‘clumps’, small flocs, long range lateral flocculated networks,
HPC-silica systems, packing of only 30% was observed for or coagulated systems. These are very difficult to categorise
stable systems, a figure far below a hard packed layer distinctly, because of the great difficulty in directly observing
considered necessary for strong steric repulsion. It was their nature at a nano level. The range of networks is greatest in
suggested that free HPC could be adsorbed at the droplet low to moderate concentrations of particles, where they may
interface, also inhibiting coalescence, so less silica was either adsorb as individuals, or in small multi-particle units, or
required for stabilisation. It was also thought that although possibly progressively acuminate and grow from a ‘seed’ type
actual coverage was only 30%, it consisted of long range aggregate.
partially flocculated networks, giving strong lateral support. Fig. 12 illustrates four basic categorised network struc-
Others who have considered the effect of low coverage tures, which may form under different bulk and interfacial
stabilisation are Vignati and Piazza [34]. They investigated conditions in emulsions and foams. Condition (A) shows
systems where the silica concentration was only 15% of a a classical hexagonally close-packed (HCP) monolayer,
packed monolayer, and excitingly found that it could still where particle–particle interactions dominate, and balanced
impart substantial stabilising power in this range. Observations repulsive and attractive forces hold the network in place.
of fluorescent core silica under an electron microscope showed Condition (B) shows partially-flocculated or coagulated
that there was strong evidence of individual particles ex- ‘islands’, where particles are initially flocculated into small
networks in the bulk prior to interface adsorption, where they
may undergo further rearrangement. Condition (C) is another
form of partial flocculation, where capillary attraction forces
small networks together into meso or chain like structures at
the interface (thus giving strong lateral stability in low con-
centration ranges, as envisaged by Midmore [104]). Finally a
high concentration flocculation is shown (D), where signif-
icant multilayer aggregation occurs, and forces are domi-
nated by particle–interface attraction or bulk hydrophobic
attraction rather than particle–particle repulsion forces. These
mechanisms may help explain some of the range of network
behaviour.

3.1.2. Droplet–droplet particle bridging


A stability mechanism that can be associated with a specific
low-particle-concentration interaction is particle bridging.
Here, if oil droplets have an incomplete coating, particles
adsorbed onto one droplet can directly interact with a
neighbouring droplet interface (much like in constricted foam
systems), causing stable droplet flocs. Gosa and Uricanu in
their work with PEO polymer surfactants and particle mixtures
[106] considered that polymer chains could bridge from the
Fig. 11. Concentration specific network structures of polymer particles silica between oil droplets, causing clustering, but increased
(∼ 770 nm) on octanol emulsion droplet: A) dense layer B) Colloidal ‘clumps’ overall stability from additional hindrance to film drainage.
C) individual ‘loose’ particles D) homogenous HCP layer [34]. This type of cluster stabilisation was also considered by Vignati
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 69

Fig. 12. Framework of possible particle network structures at oil or air interfaces: A) HCP layer B) Low concentration island ‘clumps’ C) low concentration long-range
lateral meso chains D) High concentration multilayer aggregation.

and Piazza to help stabilisation in their droplet systems of low replaced) not only found that these ‘zips’ resisted interfilm
particle coverage [34]. Here, even clean particles (without changes, but when interfilm liquid was replaced to a point that
bridging polymers) could cross the inter-film region and bond could eventually spit the zip, the two interfaces returned to
to both oil droplets. strong individual networked monolayers [109].
Stancik et al have also found evidence of particle bridging
across droplet interfaces [74,107]. They used droplets mounted 3.1.3. Surfactant addition and synergy
at the ends of capillary tubes, which were contacted with a flat Masliyah, Wang and co-workers, through investigations on
interface, using both oil and water. As particle coated droplets petroleum emulsions, looked at the effect of negatively-
were brought into contact with the surface, particles arranged charged kaolinite clay particles on emulsion stability, with
themselves into ring structures around the outer contact area, addition of anionic palmitic acid and cationic dodecyl amine
thus showing the lateral movement necessary to produce a (DDA) surfactants [36]. The addition of DDA to O/W
‘clean’ drop-drop interface, required for coalescence. However, emulsions stabilised by kaolinite caused a significant drop in
the droplets would only partially coalesce, and when removed, interfacial surface tension, dependent on the amount of
resisted, but were cleanly separated. kaolinite in the system. It was thought that particles at the O/
In follow up work, both Stancik and Fuller [108] and W interface surface attracted the oppositely charged DDA
Horozov et al. [109] found that interfaces in close contact could surfactant molecules, which then adsorbed onto the oil–water
undergo long-range particle bridging, called ‘particle zips’. interface, to a greater extent than if no particles were present.
Under certain circumstances, when two networked particle laden Attraction of kaolinite particles to DDA was great enough for
interfaces come into contact, the two interfaces will ‘zip’ to form DDA molecules to be attracted through the oil onto particles in
one tightly networked monolayer [108] (Fig. 13), and remark- the water.
ably this monolayer would continue to ‘zip’ nearby particles in In a similar approach, Tambe and Sharma focused on
both layers, growing outwards from the initial point of contact, idealised systems of naturally occurring oil-brine emulsions,
compressing the particles to a finite distance, analogous to using barium sulphate, calcium carbonate and silica [2,32,33].
crystal formation [109]. Horozov, using a vertical dual interface Stearic acid was used as a surfactant emulsifier. Although not
(where the interlayer liquid film could be withdrawn and focusing on surface tension, it was found that surfactant
70 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

addition to particle mixtures led to emulsion stability far greater


than found with either particles or surfactant alone (Fig. 14).
This synergy was attributed to the surfactant's ability to
increase the particle contact angle on adsorption. For highly
hydrophilic particles, surfactant addition caused the contact
angle to increase in parallel with stability to angles of 90°, and
for certain particle systems, they were able to induce emulsion
inversion from O/W to W/O by increasing the relative contact
angle from below to above 90°. At high surfactant particle
ratios, emulsion behaviour similar to surfactant only systems,
as surfactant molecules began to compete for active sites on the
oil–water surface, thus inhibiting particle adsorption.
In a similar finding Gosa and Uricanu found that systems of
silica and PEO–PPO–PEO block copolymers caused synergis-
tic stability [106]. Again, systems of low surfactant to particle
ratios created optimum conditions, and that increased surfactant
concentration lead to site competition between the particles and
surfactant molecules, leading to an overall stability decrease.
They attributed the synergy to principally increased flocculation
of silica with surfactant adsorption. It was thought that co-
polymer bridging mechanisms would lead to small oligomeric
silica flocs, which when adsorbed onto the oil–water interface
would cause stronger networks, increasing hindrance to lateral Fig. 14. Half life of emulsions stabilised by stearic acid with (open squares) and
movement (the main form of instability). without (closed squares) calcium carbonate particles [2].
Midmore furthered these findings, in systems of silica and both
PEO [110] and HPC [104,105] surfactants. Again clear synergy
was found between silica and each of surfactants, with stabili- continued to gain experimental backing. Very recently Binks and
sation initially increasing and then decreasing, once surfactant co-workers looked at the synergy between moderate concentra-
concentration was heightened to compete for the oil–water tions of nanosilica (20 nm) and non-ionic PEO type surfactants
interface. PEO polymers of greater chain length gave enhanced [111]. For systems with high stability clear evidence of small
results, because of their preference to adsorb onto silica, and aggregate formation was present, confirmed with bulk rheology
hindered preferential absorption onto the oil–water interface. In tests showing increases in viscosity linked to this flocculation.
terms of synergic action, there was evidence for increased Interestingly, they found that the interfacial surface tension was
flocculation of silica with surfactant addition, and increased con- negatively affected (increased) by particle addition. This was
tact angle. Again a likely scenario was that upon emulsion balanced by changes in the particle contact angle, which overall
agitation, silica flocs initially formed would be broken into small produced a net increase in the theoretical detachment energy (i.e.
oligomeric and monomeric flocs that underwent further network- increased stability) for moderate concentrations of both species.
ing on the oil–water surface (See Fig. 15). In terms of flocculation stability, questions remain as to how
The hypothesis that synergistic stability is caused by unique particle–surfactant species can intrinsically aid stability of
flocculation mechanisms between particles and surfactant has emulsions and foams. From bulk rheology tests, it is not hard to
judge how partial flocculation may impose more significant
bulk viscosity and so inhibit drainage. However, for agglom-
erates adsorbed at the interface, it is not immediately obvious
how these particle–surfactant clusters could hinder coalescence
more than, say, a strong HCP layer of individual particles. In
fact pre-flocculated clusters should not undergo the same
network rearrangement as individual particles, and so may not
produce such homogenous steric layers.
The key to understanding the synergy between particles and
surfactant may in fact lie in how their joint arrangements at an
interface could change interfacial rheology properties. It is
hypothesised here that surfactant–particle mixtures may
increase interfacial elasticity and cohesiveness over particles
or surfactant acting alone. In effect, interlocking surfactant
Fig. 13. Formation of particle zips: A) interfaces initially brought into contact chains between particles may act as elastic ‘springs’ in the layer.
B) particles conform to bond to interstitial sights on the opposing interface Also, combined effects could impart significant changes the
C) monolayer ‘zip’ is formed, expanding out from initial contact point [108]. interfacial viscosities (essentially the intrinsic hydrodynamic
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 71

3.2.1. Density effects and disproportionation


The effect of increased density difference between air–water
foams over oil–water emulsions is evidenced simply by an
increased rate of film drainage [7]. Arditty and co-workers
highlighted these differences in work with bi-liquid foams. The
bi-liquid foams generated were essentially condensed O/W type
emulsions, where the water-continuous phase was withdrawn to
only 10%, leading to condensed partially coalesced oil ‘bubbles’.
These foam like droplets held strong structure and stability far
beyond what would be considered for gas–water type systems,
because of the decreased effects of gravity induced drainage
[7,114,115].
It would seem that the most difficult de-stabilising influence
to overcome is that of molecular diffusion, or disproportionation,
as even a highly packed steric barrier (such as proteins or
Fig. 15. Partial flocculation characteristics of particle–PEO surfactant
particles) would leave holes at a molecular level [8]. In fact
aggregates on a O/W emulsion droplet (TEM) [110]. Dickinson, in and extensive investigation into foam stabilisation
by proteins, found for a range molecules that augmented inter-
facial viscosities and elasticity, stabilities were similar, and
disproportionation could be decreased no more than 2–3 orders
resistance to flow of the layer), much like increases to the bulk of magnitude [116]. It seems in response to disproportionation,
viscosity. Given in Fig. 16 is a possible ‘elastic’ arrangement of greater retardation can occur if interface particles have a high
particles and surfactant at an interface. degree of inter-networking, as evidenced by Martin (using
It can be generally surmised that there is a possibility for proteins) [117], and this is backed up by the view that proteins
surfactants to have three synergistic effects on particle–interface can only create a truly stable foam, if the species undergoes
stabilisation. Surfactant–particle mixtures may firstly decrease coagulation or aggregation at the interface, creating a strong
the interfacial tension (or at least greatly alter it, as further lateral network [118]. This would suggest that particles that
discussed in Section 3.2.3). Secondly surfactant adsorption acts undergo partial coagulation/flocculation may be perfect candi-
to change particle contact angle (and so alter stabilisation dates for foam stabilisers. It must also noted here that affects of
energies). Lastly, surfactants may help promote partial particle disproportionation are most evident in systems using gases of
flocculation and network rearrangements at the interface, high molecular solubility, such as CO2 [5]. As such, dispropor-
perhaps observable through changes interfacial rheology. This tionation is commonly a problem encountered in the food
last topic is an area not well researched, but is one of increasing industry (i.e. with carbonated beverages) and is not as significant
interest. The best indicators can be found in studies of proteins in in most air–water foams (although still an important mitigating
food systems, where changes to hydrophobicity, and addition of mechanism).
surfactant have led to differences in measured viscous elastic
moduli [112]. One of the first papers to directly study dilatational 3.2.2. Non-adsorbing Inter-film stabilisation
changes to particles and particle–surfactant mixtures is work by As well as studies relating to particles as attached stabilisers,
Ravera et al [113]. Here, aqueous drops containing nanosized there has been work into the mechanism by which non-adsorbing
silica and CTAB, allowed dynamic monitoring of changes in particles may stabilise foams, specifically the phenomena of
network structure under harmonic perturbations. The measured
dilatational modulus was markedly higher for droplets with
CTAB–particle mixtures than CTAB alone, indicating the
potential for high stability emulsion and foam systems.

3.2. Foam systems

When compared with their effect in emulsions, the action of


particles in the stabilization of foams are inherently more com-
plicated. Of particular importance are the effect of larger density
differences between the respective phases in air–water mediums,
the increased effects of disproportionation and the pointed ability
of hydrophobic particles to act as foam de-stabilisers [8]. The
study of particles in foam systems still remains a very challenging
and important prospect, because of the potential consequences for
related industrial fields such as particle flotation [26] and the
metal foam industry. Fig. 16. Possible joint network of particles and surfactant at an interface.
72 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

‘stratification’ (or stepwise film thinning). Stratification occurs the surface tension is a critical factor in defining stability. There
when particles (or particle type species) are contained within a have been a number of interesting effects reported for certain
thinning inter-bubble film, and are forced into ordered layers of types of particle only systems, in addition to the synergic action
particles, that are expelled out of the film surface into the Plateau often reported with surfactants. Nushteava and Kruglayakov
boarders in a stepwise manner, with each step relating to a specific [129] found similar surface tension behaviour to Masliyah and
concentration of ordered species [15]. This stratification process Wang [36]. They observed a dramatic drop in the surface tension
increases stability, through an additional contribution to the of an oil–water interface, when particles were introduced to a
capillary disjoining pressure [15]. In terms of foam stability, CTAB surfactant system. It was concluded that the surfactant
stratification has been shown in systems containing micelles and adsorbed onto the particles, increasing agglomeration and
globular proteins [119–122]. attachment at the interface, in line with results and theory
Sethumadhavan, Wasan and co workers, have completed presented by Tang at an air–water interface [49]. Recently some
extensive research into the stratifying behaviour of non-adsorbing more foam based behaviour has shown the same distinct
silica nanoparticles in thin film air–water systems [123,124]. It reductions in surface tension. Gonzenbach and Gauker have
was found that indeed silica produced strong stratifying observed surface tension decreases for nanosized aluminium
behaviour, with the number of thinning layers increasing with oxide and a range of short-chain fatty-acid surfactants [130]. The
concentration as expected. Stratifying behaviour was attributed to greatest decreases were seen in surfactant ranges over a critical
the existence of a gradient in the chemical potential of the par- concentration (where adsorption was maximised) (Fig. 18). It is
ticles, or an osmotic pressure gradient, where particles leave the theorised that the particles disrupt the natural equilibrium of the
film and vacancies appear in their place. Smaller particles pro- surfactant. Firstly adsorption onto particles will remove free
duced lower rates of black spot formation, and overall produced surfactant, but secondly the hydrophobised particles move to the
films more stable to rupture, indicating the importance of particle interface as surfactant carriers. Here, the action of the
size. There was also evidence for an order-disorder transition of surfactant–aluminium particles may collectively bring more
the particle layers, as the particles approached monolayer co- surfactant to the interface than a free equilibrium of surfactant
verage [124], from a distinctive increase in packing efficiency, only. However, the extremely high concentration of particles
leading to greater film stability. Basheva found similar well used in this system (35%) makes comparison of this phenom-
ordered colloidal structure, in studies with latex suspensions, and enon with other experimental systems difficult. It is also poten-
calculated the effective dynamic viscosity as being approximately tially a very complex interfacial system, and the free energies
100 times larger than pure solvent viscosity [125]. In terms of involved with particle, surfactant and mixture adsorption are
general packing structures, Basheva used interference methods to difficult to associate directly. There is still room for better
study the structure of stratified latex particles [126]. It was understanding of this process, and the combined effect of
observed that the particles form hexagonal close packed structures particles on surface tension show very wide ranging outcomes,
exclusively, with no evidence of tetragonal type packing. as evidenced by other authors.
Sethumadhavan et al considered the changes to stratifying An interesting study of surface tension changes with surfactant
behaviour, in systems containing non-uniform particles [127]. and particle systems was conducted by Marquez and Grady,
The stratifying behaviour was seen to reduce markedly upon the studying polystyrene (PS) latex and SDS structure with a
polydispersity of particles being increased. This behaviour was Langmiur-Blodgett trough [131]. Correlations were found
observed, because of the degradation of ordered particle layers between surface tension and particle structure (Fig. 19). For
with size variation in particles, and was even more apparent in bi- SDS/PS, the cmc is shifted to higher SDS concentrations, but a
disperse systems; giving systems of very low film stability (see sharp increase in surface tension occurs between 0.5–2 wt. SDS.
Fig. 17). The relative flat plateau region (dotted circles) corresponds to the
Velikov and Velev looked into the inter-film behaviour of range of surfactant concentrations at which ordered monolayers of
latex particles, with various surfactants, rendering the particle latex spheres form on the surface. Above this plateau, the surface
behaviour either adsorbing or non-absorbing [128]. With anionic tension returns to values near the SDS solution, and this is
surfactant, the particles became fully dispersed in the aqueous reflected in latex cluster formation and possible phase separation.
phase, and were expelled between the film surfaces in a step wise For nonionic surfactant addition, no ordered structure was found,
(stratifying) manner. With particle adsorption of protein, there consistent with no appearance of the unique feature in surface
was some observed aggregation of particles between the film tension results. It was concluded that the tensiometer can be used
surfaces, because of decreased mobility with protein coating, to determine the conditions under which well ordered latex films
and long-range capillary attraction. With cationic surfactant, the form on substrates.
latex hydrophobicity was increased, and the particles adsorbed Okubo used surface tension measurements to study the
onto the film walls in a bridging manner similar to that observed distribution and ordering of colloid spheres in deionized water
with emulsions. at interfaces (without surfactant addition) [132]. Two kinds of
colloidal systems were studied: monodispersed polystyrene
3.2.3. Particle effects on surface tension latex particles (strongly hydrophobic surface) and silica (polar
An interesting approach that has not been comprehensively surface). A total of 19 kinds of colloidal systems were
studied, is the ability of particles to affect interfacial surface investigated in the size range of 6–460 nm. In the case of the
tension. However, if one relates back to surfactant systems [19] colloidal silica suspensions (for size ranges 10–200 nm)
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 73

Fig. 17. Black Spot expansion vs. film diameter, mono-disperse and bi-disperse silica [127].

relatively minor reduction in surface tension from that of water These reports highlight the complexity involved in the
values (− 2 MN/m) suggested very weak surface activity. On the relationship between surface tension fluctuations and particles at
other hand, with polystyrene spheres the surface activity was or near the interface. One may make the interpretation that partial
unusually high for spheres of a particular critical size range crystallisation and aggregation of particles in certain size ranges
diameter of 100 to 200 nm, and resulted in surface tension can lead to a decrease in surface tension, especially for hydro-
reduction by values as great as 20 mN/m, at relatively high phobic particles. Marquez and Grady's work [131] would also
particle concentrations (10–20%). suggest certain particle structuring may lead to surface tension
From the surface energy data, the ordering of the particles at increases. The role of surfactant is also unclear. Indeed, it seems
the interface was described in terms of crystal-like, liquid-like conditions producing bulk decreases in surface tension may be
and gas-like. To explain the differences in surface activity, it very specific, as a great deal of other work has shown the
was firstly considered a result of the hydrophobicity of the latex, depletion of free surfactant by particles dominates surface beha-
and also that the surface tension of interfacial crystal-like viour and results in an increased surface tension [113,136,137].
suspensions was lower than that of liquid-like and gas-like As indicated by Ravera and workers, the behaviour could be a
suspensions. The importance of particle size in determining dynamic result of surfactant depletion, surfactant transport to the
surface activity was a significant finding (Fig. 20), and it was interface by attached particles, and as a consequence, particle–
concluded that the critical size range corresponded to sizes, surfactant rearrangements at the interface [113]. It is obvious the
which led to crystalline structures in bulk conditions, again hydrophobicity of particles and surfactant-particle species is
indicating the importance of structure on surface activity. critical in determining overall performance. The role of capillary
More recent studies by Dong and Johnson, using charged attraction in changing surface activity (as theorised by Dong
stabilised (and hydrophilic) TiO2 dispersions, demonstrated the
influence of bulk particle size and concentrations on the surface
tension [133,134]. The surface tension first decreases to a
minimum at around 5–7% mass solids, before rising back to
initial values with very high particle concentrations of around
10% (Fig. 21). It was suggested that at lower concentrations, the
movement of particles to the interface decreased the total free
energy of the system, which caused a decrease in surface tension.
At higher particle concentrations, in a situation similar to the
immersion of thin films, the capillary force due to the wetting of
the particles becomes significant. The strong capillary forces
between particles will resist interfacial deformation and increase
local effective surface tension. Results with silica micro-
particles showed a weaker, but still distinct effect [135]. In line
with Okubo [132], they also felt that particle size was a Fig. 18. Surface tension of butyric acid with (closed) and without (open) 35%
significant factor in behaviour differences [135]. alumina suspensions [130].
74 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

between particles and adjoining gas-liquid interfaces, causing


changes to the maximum capillary pressure of coalescence [67]
(as stated in Section 2.1.2). Most research in this area has been
concerned with particle flotation, with a focus on the effect of
particle hydrophobicity on mineral recovery, highlighting the
practical importance of producing stable froths.
Van Der Zon et al., using a moving bubble column, found as
expected, the use of hydrophobic coal particles in flotation caused
extensive froth coalescence, and the most hydrophobic coals
produced the weakest froths [141]. Pugh and Johansson
considered the effects of the hydrophobicity of quartz particles,
in regards to mineral processing [13]. They found particles with
θ b 40° gave no effect, whereas particles with contact angle
around 65° gave optimum froth stability. More hydrophobic
particles lead to collapsing froth, with contact angles still below
90°. For particles at the optimum contact angle, froth stability was
Fig. 19. Surface tension changes with polystyrene (PS) and SDS addition [131]. shown clearly to exceed stability for surfactant alone (Fig. 23). As
well as this, results indicated that smaller particles worked better
[134]) may also be an important regulating factor in overall at stabilising froths, and particle mixtures of hydrophobic and
effect. Capillary influence was further observed by Casson and hydrophilic particles lead to no increase in stability. It was thought
Johnson, with suspended latex particles in a dish [138]. Here that even small concentrations of hydrophobic particles lead to
capillary attraction between particles caused random initial ag- dominating de-wetting interactions, although instability could be
gregation, leading to fluctuations in the local surface tension, to also due to changes to capillary action, as hypothesized in mixed
such an extent to cause tension-gradient driven particle flow. emulsions [17].
Schwarz found similarly that optimum particle contact angles
3.2.4. Particles as foam stabilisers and destabilisers for stability occurred around 63° [142], and Jameson and Ata, in
The use of particles as foam stabilisers is more complicated experiments using froth columns to selectively separate
than their use in emulsions; primarily because hydrophobic hydrophilic and hydrophobic particles, found particles of around
particles can act as foam de-stabilisers. The antifoam action of 66° contact angle gave optimum stability conditions [143,144].
hydrophobic particles is caused by heightened capillary de- In both cases, increasing particle hydrophobicity led to dry
wetting action, from the extenuated water receding angle partially coalesced froths, because of increased film drainage.
[15,139]. Given a smooth particle where θ N 90° for particles Ata and Jameson also looked at mixtures of hydrophobic and
that bridge between bubble films, the receding contact angle will hydrophilic particles (in terms of selective flotation perfor-
cause de-wetting of the interface film. This drainage is propelled mance) and found that hydrophobic particles dominated inter-
by a positive Laplace pressure in the film adjacent to the particle actions, and through bridge film drainage, entrapped water wet
[15]. For particles where θ b 90°, the drainage stops as the Laplace
pressure difference reduces to zero and the film becomes
essentially planer, when the two air interfaces attain steady
contact angles with the particle (assuming both air interfaces are
similar and give θ b 90°) [15]. In fact, in hydrophilic particle
bridging systems, any further drainage will cause the Laplace
pressure to draw liquid back into the particle film, creating a stable
bridging mechanism. For hydrophobic particles, as both air
surfaces will try to attain a steady contact angle above 90°, de-
wetting will accelerate until the inter-film is completely drained
(Fig. 22). Particles can further be used in conjunction with various
non spreading oils as very effective foam breakers, where the
particles protrude to create an unstable asymmetric air–water-oil
interface, allowing the entry of oil globules, that rupture the foam
through dewetting bridging & stretching mechanisms [15,140].
Although these wetting mechanisms imply that particles with
any contact angle below 90° should act as foam stabilisers, much
research has shown that particles can begin to act as de-foamers
with contact angles much lower. This is partly from the effects of
meniscus formation and surface roughness, leading to ‘non-
equilibrium’ wetting characteristics. However, it may also be a Fig. 20. Surface tension measurements of silica (X) and latex (O), with
result of the relatively low concentration and high interaction increasing diameter (mass fraction 0.2) [132].
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 75

required for particles to potentially stabilise foams, indicating


that particles above 3 μm would probably not be effective
stabilisers. This is significant, because it shows foam systems
rely more critically on particle size than emulsions. Whereas in
emulsions, larger particles will tend to form larger emulsion
droplets, large particles in foam systems will tend to pierce and
break foam systems. Another factor important to stability is the
shape of particles, especially relevant to the flotation industry,
which regularly encounters larger, non-symmetrical particles.
Dippenaar, in investigations [51,52], established that ‘flaky’ flat
type particles generally result in more stable three-phase froths
Fig. 21. Surface tension changes due to TiO2 particle suspension [134]. from as a result of lateral alignment along the contact region.
Square (or ‘rhombic’ type particles, such as galena) can align
hydrophilic particles, inhibiting separation [144]. Along with the either laterally or diagonally, resulting in differentiating stability
work of Aveyard [139], it can be generally surmised that the profiles. In further examples, rounded or spherical particles
range of contact angles in which particles stabilise foams is accomplished thinning and rupture of a liquid film in
small, and there is a sudden drop, in line with the critical rece- approximately 0.1 s, whereas sharp-edged particles ruptured
ding angle, from where particles begin to work as de-stabilisers. the liquid film in only 0.02 s.
This signifies that particle stabilised foam systems are far Despite the extensive research into particle foam stability in
more susceptible to conditional changes than particle–emulsion terms of flotation, there has been relatively little work done on
systems. the establishment of stabilising mechanisms of pure foam and
Kaptay looked at general forces governing stability in foams particle systems (without the surfactants and frothers used in
[66]. Unsurprisingly he found that particles could stabilise foams flotation). Of note is the largely unpublished work of Wilson
under different conditions in a range of contact angles, in all [145] (see Binks [26]). Using froths stabilised by latex particles,
cases between 20° and 90°. Structurally, in line with network Wilson found that the most stable froths were made, if conditions
formations on emulsions, particles could form loose monolayers were developed close to bulk particle coagulation conditions
to packed multilayers on bubble surfaces. He also determined a (either through salt addition, pH changes, or surfactant addition).
dimensionless factor PR/γ (where P = pressure destabilising This is again in line with the theory that for particles to work
foam, R = particle radius, γ = liquid surface tension) for calcu- effectively as stabilisers in pure systems, partial coagulation is
lating potential foam stability. It was theorised that PR/γ b 40 is necessary [118]. Wilson also found that the optimum contact

Fig. 22. Bridging particle behaviour in a foam. A) hydrophilic, B) hydrophobic particle [139].
76 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

conditions by Murray and co-workers (see Fig. 25) [152]. Foam


stability is seen to drop more significantly at NaCl levels below
1 mol l− 1, relating to a ‘minimum’ salt concentration needed for
strong stability This was not only due to the coagulation and
gelling characteristics (shown by rheological measurements),
but also because salt screening increased the contact angle of
adsorbed particles. Interestingly almost no stability was seen in
systems without salt.
The apparent high stability to disproportionation is a most
significant result, as even considering the coagulated nature of
the particles; diffusion should still be possible on a molecular
level. This was considered by Subramanian et al., who modelled
the effects of gas diffusion on bubbles with an arbitrary number
of partially wet spheres [153]. Modelling showed that if the
local particle laden interface could be approximated to a flat
surface, it is stable to slight perturbations caused by fluctuations
in the gas volume, as any gas diffusion would cause changes to
Fig. 23. MIBC froth stability, with (filled) and without (open) 2 wt.% 26–44 μm
quarts [13].
the curvature restoring equilibrium.
These studies point to a few general rules guiding foam
stability in the presence of particles. It would seem that foam
angle for stability was closer to 85o, a value higher than gene- stabilisation is increased if a high concentration of small
rally found in flotation type systems. (nanosized) particles of moderate hydrophobicity (generally
Armes and co-workers [146,147] have furthered Wilson's 60–70°) are used, and de-stabilisation is increased with the use
research into latex stabilised foams, and have revealed some
interesting possible differences to emulsions. They firstly found
that particle stability increased with decrease in particle size only
to a point. In fact a minimum size of 260 nm was needed to
produce stable foam [146], which is about an order of magnitude
above the stable size range in emulsions [37]. However, they
also found the stability was linked strongly with the polymer-
ization and initiator species, and that (for the latexes tested) the
smaller particles were more polydisperse, with interface net-
works of smaller lateral ordering. This shows the polydispersity
of a system can be more important in forming homogenous
intrinsically strong networks than simply the given size of a
particle. Most importantly, SEM micrographs (Fig. 24) show in
all stable systems, strong ordered hexagonal networks with no
evidence of flocculated aggregates.
This is quite different to the surmised behaviour of silica
systems that Dickinson and co-workers have detailed in com-
parisons with protein stabilisation [148,149]. Silica nanopar-
ticles (of around 20 nm) could impart outstanding stability
in foams, especially in respect to disproportionation and bub-
ble shrinkage, giving results far above those achieved with
proteins. Interestingly (in line with Wilson [145]) although foam
stability was high, it was initially quite difficult to generate the
foams in clean systems [148]. This fact was also noted by Bindal
and Sethumadhavan, who found even a small degree of poly-
dispersity drastically reduced silica suspension foaminess [150].
Dickinson determined that in terms of overall stability, silica
nanoparticles should also be partially coagulated at the interface
[149], where salt concentration should be at a point to form small
oligomeric flocs (upon system sonication), which would
undergo further networking at the interface. This conclusion
was also found by Binks in similar systems of nanosilica foams
[151]. The importance of salt and electrolyte screening was Fig. 24. SEM of a latex foam fraction (A) and (B) a close up SEM of latex
furthered from research into particle behaviour in high salt bubble packing, both with 1.14 μm latex particles [146].
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 77

of even a small concentration of larger more hydrophobic


particles. Also, as with emulsions, partial coagulation of particle
networks on the surfaces of the bubbles is found to be advan-
tageous for stability, especially in silica systems of low to
moderate contact angles. The role of salt coagulants may also be
complicated by a possible effect on particle contact angle, and
specific ion effects [152]. One other generality is that the ability
of particles to stabilise foams is not necessarily associated with a
given particles ability to generate a foam, and because of this,
the role of co-surfactants or foamers maybe even more critical
than in emulsion systems.

3.2.5. Particles in metal foams


An interesting comparison can be made with particle stabi-
lising behaviour in metal foams. Cellular metals are a novel Fig. 26. Examples of “METCOMB” aluminium foam [155].
class of lightweight material with unique mechanical, thermal
and electrical properties. They have potential applications as produced by firstly adding calcium to an aluminium melt in the
heat exchangers, in fuel cell systems, steam generation, in the presence of air to aid oxidation. Then TiH2 particles are admixed
automotive industry (car bodies) and for light-weight structural to instigate foaming [157,158]. These inner oxide particles
elements in building [154]. In addition, since they have the (films) are in the submicron range. With the CYMAT and
ability to absorb large amounts of energy at almost constant METCOMB processes, liquid metals reinforced with micron-
pressure without generating high peak stresses, they can be used sized ceramic particle (eg. SiC and Al2O3) are foamed by direct
as crash absorbers and sound isolation elements (as in tunnels) gas injection [155,157]. These latter techniques represent the
[154]. Today, although several methods of fabrication of metal most promising, and offer the highest quality foams and superior
foams exist, it is still difficult to achieve homogenous structure technology to produced foamed parts. Generally however, a
with uniform properties. The manufacture of metal foams is larger fraction of particles is added in direct injection systems,
very challenging since it involves the simultaneous occurrence and problems occur when trying to disperse submicron sized
of solid, liquid and gaseous phases at varying temperatures and particles in these processes [155], a problem, which if overcome,
the morphology of the solidified foam is quite complex [155]. could lead to dramatic improvements in stability and economy.
Several metals and metal alloys can be foamed (Al, Zn, Mg, Pb, In all manufacturing routes known up to now, foam stability is
Fe), and among these, aluminium is the most important for achieved by solid particles with particle sizes ranging from
applications. Several routes and competing manufacturing tech- 30 nm to 20 μm. As a rule, smaller stabilizing particles require a
nologies for foamed metals are now known. Melts can be lower volume content of such particles. Metal foams actually
foamed directly by gas injection, or by indirect foaming from give one of the clearest and significant proofs of the potential for
metal power compacts [156]. Fig. 26 shows an example of particle-stabilising species. Yet, even here (partly extenuated
‘direct injected’ aluminium metal foam. because of differences in the various manufacturing systems) the
There are a number of characteristic foam types and trade nature of action of these particles is still under dispute. There is
names, all with different potential properties and uses. Examples conflicting evidence of whether a particle's major contribution
include the European ALULIGHT, produced by compacting lies in the inter-film layer with consequential effects on rheology
metal and alloy powders with gas releasing TiH2 particles [157]. and drainage, or whether (as is most evident in, say, emulsions) a
The foams develop during baking of this precursor, which particle–interface steric barrier is the major component. Gergely
contains nanometre-sized resident oxide particles from the and Clyne stated that high particle concentration could increase
former surfaces of powder mix. The Japanese ALPORAS is viscosity, and inhibit liquid drainage [159]. Ip suggested
particles that were attached to the interface caused a flattening
of the bubble curvatures around the Plateau border, reducing the
capillary pressure, and retarding drainage [160]. Babcsan and
co-workers found that additionally, Al2O3 particles in liquid
aluminium alloy suspensions reduced the surface tension sig-
nificantly at high concentrations, highlighting another potential
basis of stability [156].
The many possible stabilisation mechanisms observed in metal
films were recently categorised by Hailbel and workers [161],
seen in Fig. 27. Here (a) represents a clean foam interfilm, with a
pressure difference between the bubbles and the interfilm and also
Fig. 25. Comparison of foam fraction (F) with increasing time for 20 nm silanated the plateau boarder causing drainage. Possibility (b) highlights a
silica and salt. Filled diamonds, 3 mol l− 1 NaCl; open squares, 2 mol l− 1 NaCl; single layer of particles of low absorptivity increasing the
closed triangles, 1 mol l− 1 NaCl; open triangles, 0.5 mol l− 1 NaCl [152]. maximum capillary pressure of coalescence. Mechanism (c)
78 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

suggests that small particles of low-to-moderate adsorbance at did yield significant improvements in initial gas expansion of the
interfaces can alter the capillary curvature of the surface, and so foamed melt. Bubbles/pores had smaller average diameters, with
reduce the capillary flow between interfilm and plateau border. (d) smaller interfilm thicknesses. Optical micrographs (Fig. 28)
Indicates particles flocculated in the bulk may bridge between clearly show the TiB2 particles of the composite film residing
bubbles as aggregates, while (e) highlights how non adsorbing mainly in the interfilm (i.e. non-adsorbing). Stabilisation in this
particles may affect bulk viscosity. No apparent bridging between manner could simply be a result of particles impeding film
either aggregates or individual particle layers, were found in their drainage, as changes to capillary curvature are not immediately
experiments, as interfilm thicknesses were quite large. They evident.
therefore concluded that stability mainly came from adsorption Another factor in equating the dominating stability interac-
based steric repulsion, with a few non-adsorbing particles trapped tion from particle addition is the effects of particle size. First
in the interfilm region (seen as mechanism (f)). considered in line with volume fraction by Jin et al [166], and as
However, the majority of evidence does seem to indicate that revised by Kaptay [66], there are some well grounded
particles can still significantly reduce capillary flow and drainage conclusions. Generally, cell walls will rupture at larger minimum
[158], as many systems with low particle absorbance still show thicknesses in metal foams than traditional aqueous foams, and
increased stability. A further complicating factor is the high the diameters of the stabilising particles, have a direct affect on
concentrations of particles often used, where static repulsion and the attainable minimum cell thickness. Larger particles will give
also bulk viscosity changes should be even more apparent. Ip weaker films that rupture at greater interparticle thicknesses
[160], found that in direct injection systems, a minimum particle [167]. The direct affect of particle size on cell wall thickness
coverage of 50% was required, while Babcsan et al. [162] supports the concept of interfilm stabilisation (eg. capillary
concluded that optimum packing of SiC and Al2O3 particles was pressure interactions); yet particle size relationships are also
45% in nitrogen gas foaming systems. Both Lietlmeier et al. [163] indicative of a steric barrier, as smaller particles can naturally
and Ip [160] suggested that for particle sizes of around 10 μm, a form a more complete, well ordered layer. Esmaeelzadeh et al
mass fraction of at least 10% was required for long-term stability [168] and Ip [160] have experimentally shown the affect of
(generally mass fractions can range from 10–20% [157]). Ip also different particle sizes on metal foam stability. Ip produced clear
showed that an increase in system temperature led to reduced and direct support for a strait inverse relationship between
system stability, attributed to a decrease in bulk viscosity, in line particle size and stability (Fig. 29) [160], which is also present in
with the proposal of Gergely and Clyne [159]. emulsions [37] and aqueous foams [49].
The effects of particle wetting are a complicated matter, The evidence suggests that the high level of stability in metal
owing to general kinetically-controlled changes of substrate foams can be gained from both drainage and coalescence
wetting on molten metals (reactive wetting), often causing shifts retardation. Difficulties in defining the major contributor in a
with temperature, time or with parameters such as drop volume particular situation could hinder future progress. Despite the
[164]. Some trends are consistent, with those in non-metal improvement in quality over the past 10 years, metal foams still
foams. Ip [160] found a critical wetting angle of around 75° gave suffer from non-uniformities and other deficiencies. In order to
optimum performance, values close to flotation air–water improve the cellular structure of the materials, and also to make
foams. Kennedy and Asayayisitchai [165] looked at the effects the production technologies more reliable and reproducible, the
of TiB2 particles on indirect power-compact Al foams, which stability of liquid metal foams needs further investigation,
contain large amounts of oxides. The oxides inhibited TiB2 particularly aimed at avoidance of rupture and the control of
particle wetting, giving low TiB2 particle contact angles, a liquid drainage. In addition, the high content of stabilising
reason stated for the low long term foam stability. It was also ceramic particles makes direct injection foams such as MET-
found that despite the low long term foam stability, the particles COMB difficult to cut. A strong research effort in this field is

Fig. 27. Mechanisms of particle stabilisation in metal foams: (A) clean bubble interfilm, (B) stabilisation from bridging particles altering Pcmax, (C) weakly adsorbing
particles reducing capillary flow through meniscus alterations in curvature, (D) bridging flocs transmitting mechanical ‘disjoining’ force, (E) non-adsorbing particles
altering viscosity, (F) strongly adsorbing particles creating steric barrier, with added trapped particles inhibiting drainage and bulk rheology [161].
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 79

still questions relating to whether the long-range meso structures


shown at static interfaces, are present in many droplet/bubble
systems. It may be thought that the large energies involved in
interface formation, and effects of other factors such as salt and
surfactants, may produce more small ranged coagulated networks
(not complex capillary induced structures). Conclusive evidence
of network formation is unfortunately inherently difficult in real
emulsion or foam systems at a nanoscale level. This kind of
information is critical however in determining the reasons for
stability, especially with low concentrations of particles
In both emulsion and foam systems, work with surfactants
warrants attention. Because of the great range of surfactants,
polymers and other absorbing species available, the range of
possible interaction mechanisms is quite extensive. It is known
that surfactants may help to promote particle emulsion stability
by increasing particle contact angle, by partial flocculation of
particle–interface networks and also by lowering surface
tension. It is not clear which of these processes will dominate
in a particular situation, particularly the effect on surface tension.
The overall stability for any given system is a delicate balance of
particle–interface, surfactant-interface and particle–surfactant
interactions. Categorising these interactions and their affect on
stability is an area of important future research. There could be
potential benefits in synergistic action, which could improve the
formulation of a wide range of naturally occurring systems (such
as in food colloids, cosmetics and flotation) where a number of
different species work (and compete) together. A largely unan-
swered question is how solid particles act with other species,
such as proteins. Such large steric molecules are more perma-
nently adsorbed at interfaces, so it could be presumed that
particles would not preferentially adsorb to the same extent, as
with dynamic surfactant systems. This may mean that particles
Fig. 28. Al Foam inter-film A without and B with TiB2 nanoparticles [165].
will adsorb onto the proteins at the interface, and not the
interface itself, possibly changing stabilisation dynamics.
needed to achieve a break-through in economical production The successes with particle-stabilised emulsions have contin-
technology. In order to implement these goals, synergies between ued to lead to benefits in a number of fields. The principles of O/W
different scientific disciplines such as chemistry, physics, physical emulsion stabilisation is being used in many other systems, such
metallurgy, materials science and engineering have to be as the use in stabilisation of liquid CO2 and water emulsions
established, but it can be generally stated that advances in metal [169]. In terms of foam stability performance, probably the most
foams is generating possibilities for a whole range of foaming
systems.

4. Conclusions

The recent interest in particle-stabilised foams and emulsions


has identified the major factors affecting the stability of respective
systems. It should now be possible equate rules governing the
effect of particle size, concentration, aggregated nature and
contact angle to help improve the performance of a given system.
However, great complexities arise from the fact that these
structures attain equilibrium as a result of colloidal forces acting
on the particles and bubbles in the bulk (such as the effects of
salts) as well as interfacial forces. Difficulties also arise in
associating characteristics with the method of preparation and the
energies involved in emulsion and foam formation. A related
problem is the possible differences in aggregation kinetics,
between static interfaces and real bubbles & droplets. There are Fig. 29. Relationship between particle size and Al metal foam life [160].
80 T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81

pressing issue is to increase the range of effectiveness for a given [32] Tambe DE, Sharma MM. J Colloid Interface Sci 1994;162:1–10.
particle (to produce a larger stability ‘window’). In industrial [33] Tambe DE, Sharma MM. Adv Colloid Interface Sci 1994;52:1–65.
[34] Vignati E, Piazza R, Lockhart TP. Langmuir 2003;19:6650–6.
processes, it is impossible to entirely predict system parameters, [35] Binks BP, Whitby CP. Colloids Surf A Physicochem Eng Asp
and variation to operating conditions is inevitable. The delicate 2005;253:105–15.
nature of foams means they are very susceptible to system change, [36] Wang W, Zhou Z, Nandakumar K, Xu Z, Masliyah JH. J Colloid Interface
and particles can quickly go from acting as stabilisers to de- Sci 2004;274:625–30.
[37] Binks BP, Lumsdon SO. Langmuir 2001;17:4540–7.
stabilisers, with relatively minor alterations (such as changes to
[38] Binks BP, Horozov TS. Colloidal Particles at Liquid Interfaces.
surfactant concentration). Augmentations that may enhance Cambridge University Press; 2006.
particle stabilisation over a wider range of environments could [39] Speight JG. The Chemistry and technology of Petroleum. Marcel
be important in many areas, from the flotation industry, to food Dekker; 1991.
and textiles and production of metal foams. [40] Spiecker PM, Kilpatrick PK. Langmuir 2004;20:4022–32.
[41] Nguyen AV. Colloidal Science of Flotation, vol. 1. Marcel Dekker; 2003.
[42] Matis KA. Flotation Science and Engineering. Marcel Dekker; 1995.
Acknowledgments [43] Bindal SK, Nikolov AD, Wasan DT, Lambert DP, Koopman DC. Environ
Sci Technol 2001;35:3941–7.
Mr. Hunter would like to thank the Commonwealth Scien- [44] Hoffman Z. Phys Chem 1913;83:385.
tific and Industrial Research Organisation (CSIRO) Australia [45] Bartsch O. Kolloidchemische Beihefte 1924;20:50–77.
for additional project funding. [46] Hausen DM. Can Metall Q 1974;13:659–68.
[47] Lekki J, Laskowski J. Proc.- Int. Miner. Process. Congr., 11th; 1975.
References p. 427–48.
[48] Ottewill RH, Segal DL, Watkins RC. Chemistry & Industry; 1981. p. 57–60.
[1] Prud'homme RK, Khan SA. Foams-Theory, Measurements, and London, United Kingdom.
Applications. Marcel Dekker Inc.; 1996. [49] Tang F, Xiao Z, Tang J, Jiang L. J Colloid Interface Sci 1989;131:
[2] Tambe DE, Sharma MM. J Colloid Interface Sci 1993;157:244–53. 498–502.
[3] Lissant KJ. Emulsions and Emulsion Chemistry Part 1, vol. 6. Marcel [50] Binks BP, Murakami R. Nat Mater 2006;5:865–9.
Dekker Inc.; 1974. [51] Dippenaar A, Harris PJ, Nicol MJ. Rep Natl Inst Metall S Afr 1978;1988
[4] Nielloud F, Marti-Mestres G. Pharmaceutical Emulsions & Suspensions, 34 pp.
vol. 105. Marcel Dekker Inc; 2000. [52] Dippenaar A. Int J Miner Process 1982;9:1–14.
[5] Dickinson E. An Introduction to Food Colloids. Oxford University Press; [53] Dippenaar A. Int J Miner Process 1982;9:15–22.
1992. [54] Levine S, Bowen BD, Partridge SJ. Colloids Surf 1989;38:325–43.
[6] Pugh RJ. Adv Colloid Interface Sci 2005;114-115:239–51. [55] Binks BP, Fletcher PDI. Langmuir 2001;17:4708–10.
[7] Arditty S, Whitby CP, Binks BP, Schmitt V, Leal-Calderon F. Eur Phys J [56] Casagrande C, Fabre P, Raphael E, Veyssie M. Europhys Lett 1989;9:
E Soft Matter 2003;11:273–81. 251–5.
[8] Murray BS, Ettelaie R. Curr Opin Colloid Interface Sci 2004;9:314–20. [57] Paunov VN, Cayre OJ. Adv Mater Weinh Ger 2004;16:788–91.
[9] Hunter R. J, Foundations of Colloids Science. Oxford University [58] Ondarcuhu T, Fabre P, Raphael E, Veyssie M. J Phys Fr 1990;51:1527.
Press; 1987. [59] Xia R-K, He W-D, Pan C-Y. Colloid Polym Sci 2002;280:865–72.
[10] Laskowski JS. Fizykochem Probl Mineral 2004;38:13–22. [60] Yasuda M, Kobayashi M, Kotani T, Kawahara K, Nikaido H, Ueda A,
[11] Courtney Sr DL. Surfactant Sci Ser 1997;68:127–38. et al. Macromol Chem Phys 2002;203:284–93.
[12] Coons JE, Halley PJ, McGlashan SA, Tran-Cong T. Colloids Surf A [61] Yun X, Li H-q. Gaodeng Xuexiao Huaxue Xuebao 2001;22:1929–31.
Physicochem Eng Asp 2005;263:258–66. [62] Fujii S, Read ES, Binks BP, Armes SP. Adv Mater Weinh Ger 2005;17:
[13] Johansson G, Pugh RJ. Int J Miner Process 1992;34:1–21. 1014–8.
[14] Aveyard R, Binks BP, Clint JH. Adv Colloid Interface Sci 2003;100-102: [63] Shaw DJ. Introduction to Colloid & Surface Chemistry. Fourth edition.
503–46. Butterworth-Heinemann; 1992.
[15] Pugh RJ. Adv Colloid Interface Sci 1996;64:67–142. [64] Denkov ND, Ivanov IB, Kralchevsky PA, Wasan DT. J Colloid Interface
[16] Binks BP, Lumsdon SO. Langmuir 2000;16:8622–31. Sci 1992;150:589–93.
[17] Simovic S, Prestidge CA. Langmuir 2004;20:8357–65. [65] Nushtayeva AV, Kruglyakov PM. Mendeleev Commun 2001:235–7.
[18] Weaire D, Hutzler S. The Physics of Foams. Oxford University Press; [66] Kaptay G. Colloids Surf A Physicochem Eng Asp 2003;230:67–80.
1999. [67] Kaptay G. Colloids Surf A Physicochem Eng Asp 2006;282+283:
[19] Shaw DJ. Introduction to Colloid and Surface Chemistry. Butterworths; 1966. 387–401.
[20] Dimitrova Tatiana D, Leal-Calderon F, Gurkov Theodor D, Campbell B. [68] Visschers M, Laven J, van der Linde R. Prog Org Coat 1997;31:311–23.
Adv Colloid Interface Sci 2004;108-109:73–86. [69] Kruglyakov PM, Nushtayeva AV, Vilkova NG. J Colloid Interface Sci
[21] Russev SC, Arguirov TV, Gurkov TD. Colloids Surf B Biointerfaces 2004;276:465–74.
2000;19:89–100. [70] Slobozhanin LA, Alexander JID, Collicott SH, Gonzalez SR. Phys Fluids
[22] Ramsden W. Proc Royal Soc 1903;72:156. 2006;18:082104/1–082104/15.
[23] Pickering SU. J Chem Soc 1907;91:2001. [71] Levine S, Bowen BD, Partridge SJ. Colloids Surf 1989;38:345–64.
[24] Finkle P, Draper HD, Hildebrand JH. J Am Chem Soc 1923;45:2780–8. [72] Abdel-Fattah AI, El-Genk MS. Adv Colloid Interface Sci 1998;78:
[25] Schulman JH, Leja J. Trans Faraday Soc 1954;50:598–605. 237–66.
[26] Binks BP. Curr Opin Colloid Interface Sci 2002;7:21–41. [73] Abdel-Fattah AI, El-Genk MS. J Colloid Interface Sci 1998;202:417–29.
[27] Morrison ID, Ross S. Colloidal Dispersions: Suspensions, Emulsions and [74] Stancik Edward J, Kouhkan M, Fuller Gerald G. Langmuir ACS J Surf
Foams. Wiley Interscience; 2002. Colloids 2004;20:90–4.
[28] Stiller S, Gers-Barlag H, Lergenmueller M, Pfluecker F, Schulz J, Wittern [75] Stancik EJ, Widenbrant MJO, Laschitsch AT, Vermant J, Fuller GG.
KP, et al. Colloids Surf A Physicochem Eng Asp 2004;232:261–7. Langmuir 2002;18:4372–5.
[29] Binks BP, Lumsdon SO. Langmuir 2000;16:2539–47. [76] Pieranski P. Phys Rev Lett 1980;45:569–72.
[30] Simovic S, Prestidge CA. Langmuir 2003;19:3785–92. [77] Williams DF, Berg JC. J Colloid Interface Sci 1992;152:218–29.
[31] Abend S, Bonnke N, Gutschner U, Lagaly G. Colloid Polym Sci 1998;276: [78] Aveyard R, Clint JH, Nees D, Paunov VN. Langmuir 2000;16:1969–79.
730–7. [79] Robinson DJ, Earnshaw JC. Langmuir 1993;9:1436–8.
T.N. Hunter et al. / Advances in Colloid and Interface Science 137 (2008) 57–81 81

[80] Ghezzi F, Earnshaw JC, Finnis M, McCluney M. J Colloid Interface Sci [124] Sethumadhavan GN, Nikolov AD, Wasan DT. J Colloid Interface Sci
2001;238:433–46. 2001;240:105–12.
[81] Ruiz-Garcia J, Gamez-Corrales R, Ivlev BI. Physica A Amst 1997;236: [125] Basheva E, Nikolov A, Kralchevski P, Ivanov I, Wasan DT. Surfactants
97–104. Solut 1991;11:467–79.
[82] Ruiz-Garcia J, Ivlev BI. Mol Phys 1998;95:371–5. [126] Basheva ES, Danov KD, Kralchevsky PA. Langmuir 1997;13:4342–8.
[83] Tolnai G, Agod A, Kabai-Faix M, Kovacs AL, Ramsden JJ, Horvoelgyi [127] Sethumadhavan G, Bindal S, Nikolov A, Wasan D. Colloids Surf A
Z. J Phys Chem B 2003;107:11109–16. Physicochem Eng Asp 2002;204:51–62.
[84] Mejia-Rosales SJ, Gil-Villegas A, Ivlev BI, Ruiz-Garcia J. J Phys [128] Velikov KP, Velev OD. Langmuir 1998;14:1148–55.
Condens Matter 2002;14:4795–804. [129] Nushtaeva AV, Kruglyakov PM. Colloid J (Translation of Kolloidnyi
[85] Stamou D, Duschl C, Johannsmann D. Phys Rev E Stat Phys Plasmas Zhurnal) 2004;66:456–65.
Fluids Relat Interdiscip Topics 2000;62:5263–72. [130] Gonzenbach UT, Studart AR, Tervoort E, Gauckler LJ. Angew Chem Int
[86] Goulding D, Hansen J-P. Mol Phys 1998;95:649–55. Ed Engl 2006;45:3526–30.
[87] Danov KD, Pouligny B, Kralchevsky PA. Langmuir 2001;17:6599–609. [131] Marquez M, Grady BP. Langmuir 2004;20:10998–1004.
[88] Kralchevsky PA, Nagayama K. Adv Colloid Interface Sci 2000;85: [132] Okubo T. J Colloid Interface Sci 1995;171:55–62.
145–92. [133] Dong L, Johnson D. Langmuir 2003;19:10205–9.
[89] Nikolaides MG, Bausch AR, Hsu MF, Dinsmore AD, Brenner MP, Gay [134] Dong L, Johnson D. Adv Space Res 2003;32:149–53.
C, et al. Nature 2002;420:299–301 (London, United Kingdom). [135] Dong L, Johnson DT. J Dispers Sci Technol 2004;25:575–83.
[90] Foret L, Wurger A. Phys Rev Lett 2004;92:058302/1–4. [136] Somasundaran P, Snell ED, Xu Q. J Colloid Interface Sci 1991;144:
[91] Megans M, Aizenberg J. Nature 2003;424:1014 [London, United Kingdom]. 165–73.
[92] Kralchevsky PA, Denkov ND, Danov KD. Langmuir 2001;17:7694–705. [137] Adamczyk Z, Para G, Karwinski A. Tenside Surfactants Deterg 1998;35:
[93] Kralchevsky PA, Denkov ND. Curr Opin Colloid Interface Sci 261–4.
2001;6:383–401. [138] Casson K, Johnson D. J Colloid Interface Sci 2001;242:279–83.
[94] Lucassen J. Colloids Surf 1992;65:131–7. [139] Aveyard R, Binks BP, Fletcher PDI, Peck TG, Rutherford CE. Adv
[95] Brown ABD, Smith CG, Rennie AR. Phys Rev E Stat Phys Plasmas Colloid Interface Sci 1994;48:93–120.
Fluids Relat Interdiscip Topics 2000;62:951–60. [140] Marinova KG, Denkov ND, Branlard P, Giraud Y, Deruelle M. Langmuir
[96] Fournier JB, Galatola P. Phys Rev E Stat Nonlinear Soft Matter Phys 2002;18:3399–403.
2002;65:031601/1–4. [141] van der Zon M, Hamersma PJ, Poels EK, Bliek A. Chem Eng Sci
[97] Loudet JC, Alsayed AM, Zhang J, Yodh AG. Phys Rev Lett 2005;94: 2002;57:4845–53.
018301. [142] Schwarz S, Grano S. Colloids Surf A Physicochem Eng Asp 2005;256:
[98] Aveyard R, Binks BP, Clint JH, Fletcher PDI, Horozov TS, Neumann B, 157–64.
et al. Phys Rev Lett 2002;88:246102/1–4. [143] Ata S, Ahmed N, Jameson GJ. Int J Miner Process 2002;64:101–22.
[99] Sun J, Stirner T. Langmuir 2001;17:3103–8. [144] Ata S, Ahmed N, Jameson GJ. Miner Eng 2004;17:897–901.
[100] Horozov TS, Aveyard R, Clint JH, Binks BP. Langmuir 2003;19:2822–9. [145] Wilson J. A Study of Particulate Foams. University of Bristol; 1980.
[101] Aveyard R, Clint JH, Nees D, Quirke N. Langmuir 2000;16:8820–8. [146] Fujii S, Iddon PD, Ryan AJ, Armes SP. Langmuir 2006;22:7512–20.
[102] Simovic S, Prestidge CA. Langmuir 2003;19:8364–70. [147] Fujii S, Ryan AJ, Armes SP. J Am Chem Soc 2006;128:7882–6.
[103] Prestidge CA, Barnes T, Simovic S. Adv Colloid Interface Sci 2004;108- [148] Du Z, Bilbao-Montoya MP, Binks BP, Dickinson E, Ettelaie R, Murray
109:105–18. BS. Langmuir 2003;19:3106–8.
[104] Midmore BR. Colloids Surf A Physicochem Eng Asp 1998;132:257–65. [149] Dickinson E, Ettelaie R, Kostakis T, Murray BS. Langmuir 2004;20:
[105] Midmore BR. J Colloid Interface Sci 1999;213:352–9. 8517–25.
[106] Gosa K-L, Uricanu V. Colloids Surf A Physicochem Eng Asp 2002;197: [150] Bindal SK, Sethumadhavan G, Nikolov AD, Wasan DT. AIChE J 2002;48:
257–69. 2307–14.
[107] Stancik EJ, Gavranovic GT, Widenbrant MJO, Laschitsch AT, Vermant J, [151] Binks BP, Horozov TS. Angew Chem Int Ed Engl 2005;44:3722–5.
Fuller GG. Faraday Discuss 2003;123:145–56. [152] Kostakis T, Ettelaie R, Murray BS. Langmuir 2006;22:1273–80.
[108] Stancik EJ, Fuller GG. Langmuir 2004;20:4805–8. [153] Subramanian RS, Larsen RJ, Stone HA. Langmuir 2005;21:4526–31.
[109] Horozov TS, Aveyard R, Clint JH, Neumann B. Langmuir 2005;21: [154] Ashby MF, Evans A, Fleck NA, Gibson LJ, Hutchinson JW, Wadley
2330–41. HNG. Metal Foams- A Design Guide. Butterworth-Heinemann; 2000.
[110] Midmore BR. Colloids Surf A Physicochem Eng Asp 1998;145:133–43. [155] Babcsan N, Leitlmeier D, Banhart J. Colloids Surf A Physicochem Eng
[111] Binks BP, Desforges A, Duff DG. Langmuir 2007;23:1098–106. Asp 2005;261:123–30.
[112] Murray BS. Curr Opin Colloid Interface Sci 2002;7:426–31. [156] Babcsan N, Leitlmeier D, Degischer HP. Mater wiss Werkst tech 2003;34:
[113] Ravera F, Santini E, Loglio G, Ferrari M, Liggieri L. J Phys Chem B 22–9.
2006;110:19543–51. [157] Banhart J. Adv Eng Mater 2006;8:781–94.
[114] Arditty S, Whitby CP, Binks BP, Schmitt V, Leal-Calderon F. Eur Phys J [158] Wuebben T, Odenbach S. Colloids Surf A Physicochem Eng Asp 2005;266:
E Soft Matter 2003;12:355. 207–13.
[115] Arditty S, Schmitt V, Giermanska-Kahn J, Leal-Calderon F. J Colloid [159] Gergely V, Clyne TW. Acta Materialia 2004;52:3047–58.
Interface Sci 2004;275:659–64. [160] Ip SW, Wang Y, Toguri JM. Can Metall Q 1999;38:81–92.
[116] Dickinson E, Ettelaie R, Murray BS, Du Z. J Colloid Interface Sci [161] Haibel A, Rack A, Banhart J. Appl Phys Lett 2006;89:154102/1–3.
2002;252:202–13. [162] Babcsan N, Leitlmeier D, Degischer HP, Banhart J. Adv Eng Mater 2004;6:
[117] Martin AH, Grolle K, Bos MA, Stuart MAC, van Vliet T. J Colloid 421–8.
Interface Sci 2002;254:175–83. [163] Leitlmeier D, Degischer HP, Flankl HJ. Adv Eng Mater 2002;4:735–40.
[118] Walstra P. Physical chemistry of foods. Marcel Dekker; 2003. [164] Shen P, Fujii H, Matsumoto T, Nogi K. Metall Mater Trans A Phys Metall
[119] Kralchevski P, Nikolov A, Wasan DT, Ivanov I. Langmuir 1990;6: Mater Sci 2003;35A:583–8.
1180–9. [165] Kennedy AR, Asavavisitchai S. Scr Mater 2003;50:115–9.
[120] Bergeron V, Jimenez-Laguna AI, Radke CJ. Langmuir 1992;8:3027–32. [166] I. Jin, L.D. Kenny and H. Sang, 1992.
[121] Husband FA, Wilde PJ. J Colloid Interface Sci 1998;205:316–22. [167] Gergely V, Jones RL, Clyne TW. Trans JWRI 2001;30:371–6.
[122] Wasan DT, Nikolov AD, Kralchevski P, Ivanov I. Colloids Surf [168] Esmaeelzadeh S, Simchi A, Lehmhus D. Mater Sci Eng A Struct Mater
1992;67:139–45. Prop Microstruct Process 2006;A424:290–9.
[123] Sethumadhavan G, Nikolov A, Wasan D. J Colloid Interface Sci 2004;272: [169] Dickson JL, Binks BP, Johnston KP. Langmuir 2004;20:7976–83.
167–71.

You might also like