Sunteți pe pagina 1din 47

M

Micellar Emulsion
Microemulsions

Micellar Systems
Tharwat Tadros
Wokingham, Berkshire, UK

Synonyms
Microemulsions

Definition
Micellar systems or microemulsions are a special
class of dispersions (transparent or translucent)
which are better described as swollen micelles.
A convenient way to describe microemulsions is
to compare them with micelles. The latter which
are thermodynamically stable may consist of
spherical units with a radius that is usually less
than 5 nm. Two types of micelles may be considered: normal micelles with the hydrocarbon tails
forming the core and the polar head groups in
contact with the aqueous medium and reverse
micelles (formed in nonpolar media) with
a water core containing the polar head groups

and the hydrocarbon tails now in contact with


the oil. The normal micelles can solubilize oil in
the hydrocarbon core forming oil-in-water (O/W)
microemulsions, whereas the reverse micelles
can solubilize water forming a water-in-oil
(W/O) microemulsion. A rough guide to the
dimensions of micelles, micellar solutions, and
macroemulsions is as follows: micelles, R <5 nm
(they scatter little light and are transparent);
macroemulsions, R >50 nm (opaque and milky);
and micellar solutions or microemulsions,
550 nm (transparent, 510 nm, translucent,
1050 nm). A thermodynamic definition of
microemulsions can be obtained from a consideration of the energy and entropy terms for formation of microemulsions. The increase in
surface area when forming a microemulsion is
DA, and the surface energy increase is equal to
DA g12. The increase in entropy is T DSconf (since
a large number of droplets can arrange themselves in several ways compared with one oil
drop, which has much lower entropy). According
to the second law of thermodynamics, the
free energy of formation of microemulsions
DGm is given by the following expression:
D Gm DA g12 TD Sconf . With microemulsions
DA g12  T DSconf (this is due to the ultralow
interfacial
tension
accompanied
with
microemulsion formation) and DGm 0. The
system is produced spontaneously, and it is thermodynamically stable. Thus, microemulsions
are better described as swollen micelles.

T. Tadros (ed.), Encyclopedia of Colloid and Interface Science,


DOI 10.1007/978-3-642-20665-8, # Springer-Verlag Berlin Heidelberg 2013

686

The best definition of microemulsions is as follows: System of Water + Oil + Amphiphile


that is a single Optically Isotropic and Thermodynamically Stable Liquid Solution. Amphiphiles refer to any molecule that consists
of hydrophobic and hydrophilic portions, e.g.,
surfactants, alcohols. The driving force for
microemulsion formation is the low interfacial
energy which is overcompensated by the entropy
of dispersion term. The low (ultralow) interfacial tension is produced in most cases by combination of two molecules, referred to as the
surfactant and cosurfactant (e.g., medium-chain
alcohol).

Cross-References
Interfacial Tension
Microemulsions
Surfactants

Micelle Formation
Micellization

Micellization
Tharwat Tadros
Wokingham, Berkshire, UK

Synonyms
Micelle formation

Definition
Micellization is a process of aggregation of surfactant ions or molecules above a critical concentration (that is referred to as the critical
micelle concentration, c.m.c.) to form selfassembly structures referred to as micelles.

Micelle Formation

The process of micellization is evident from


consideration of the solution properties of surfactants which show abrupt changes at a critical
concentration that is consistent with the fact that
above this concentration (the c.m.c.), surface
active ions or molecules in solution associate to
form larger units that are referred to as selfassembled structures or micelles. The latter can
be spherical, rod-like, or lamellar structures. The
c.m.c. depends on the surfactant structure and the
medium in which they are present. For
a surfactant series with the same polar head
group (ionic or nonionic), the c.m.c. decreases
with increase of the alkyl chain length of the
surfactant molecule. For a given alkyl chain,
the c.m.c increases with increase of the size
of the polar head group (e.g., number of ethylene
oxide units). For ionic surfactants, the c.m.c.
increases slowly with increase of temperature,
and it decreases with addition of electrolyte. The
driving force of the process of micellization is the
entropy increase on association. Two main
sources of entropy increase can be described.
The first is due to the so-called hydrophobic
effect. The surfactant monomers contain structured water around their hydrocarbon chain. On
transfer of such monomers to a micelle, these
water molecules are released and they have
a higher entropy. The second source of entropy
increase on micellization may arise from
the increase in flexibility of the hydrocarbon
chains on their transfer from an aqueous to
a hydrocarbon medium. The orientations and
bendings of an organic chain are likely to be
more restricted in an aqueous phase compared to
an organic phase. It should be mentioned that with
ionic and zwitterionic surfactants, an additional
entropy contribution, associated with the ionic
head groups, must be considered. Upon partial
neutralization of the ionic charge by the counterions when aggregation occurs, water molecules
are released. This will be associated with an
entropy increase which should be added to the
entropy increase resulting from the hydrophobic
effect mentioned above. However, the relative
contribution of the two effects is difficult to assess
in a quantitative manner.

Microcapsules and Controlled Release

Cross-References
Micelle Formation
Self-assembly Structures
Surfactants

Microcapsule
Tharwat Tadros
Wokingham, Berkshire, UK

Synonyms
Microcapsules and controlled release

Definition
Microcapsules with particles in the size range
1100 mm consist of a distinct capsule wall
(mostly a polymer) surrounding a biologically
or chemically active core. They are applied
for the controlled release of the active ingredient
(a.i.), its protection against the environment
in which they are dispersed or eliminating the
material encapsulated from interacting with
other substances in the formulation. Microencapsulation is mainly carried out by interfacial
condensation, in situ polymerization, and coacervation. Interfacial condensation is perhaps the
most widely used method for encapsulation in
industry. The a.i. which may be oil soluble, oil
dispersible, or oil itself is first emulsified in water
using a convenient surfactant or polymer.
A hydrophobic monomer A is placed in the oil
phase (oil droplets of the emulsion), and
a hydrophilic monomer B is placed in the aqueous phase. The two monomers interact at the
interface between the oil and the aqueous phase
forming a capsule wall around the oil droplet.
The role of surfactant in this process is
crucial since an oil-water emulsifier (with high
hydrophilic-lipophilic balance, HLB) is required.
Alternatively, a polymeric surfactant such as

687

partially hydrolyzed polyvinyl acetate (referred


to as polyvinyl alcohol, PVA) or an poly(ethylene
oxide-propylene oxide-ethylene oxide) (PEOPPO-PEO, Pluronic) block copolymer can be
used. The emulsifier controls the droplet size
distribution and hence the size of capsules
formed. On the other hand, if the material to be
encapsulated is water soluble, a water-in-oil
(W/O) emulsion is prepared using a surfactant
with low HLB number or an A-B-A block copolymer of polyhydroxystearic acid-polyethylene
oxide-polyhydroxystearic acid (PHS-PEOPHS). In this case, the hydrophilic monomer is
dissolved in the aqueous internal phase droplets.
In interfacial polymerization, the monomers
A and B are polyfunctional monomers capable
of causing polycondensation or polyaddition
reaction at the interface. Examples of oil soluble
monomers are polybasic acid chloride, bishaloformate, and polyisocyanates, whereas
water soluble monomers can be polyamine or
polyols. Thus, a capsule wall of polyamide, polyurethane, or polyurea may be formed. Some
trifunctional monomers are present to allow
cross linking reactions. If water is the second
reactant with polyisocyanates in the organic
phase, polyurea walls are formed. The latter
modification has been termed in situ interfacial
polymerization. One of the most useful microencapsulation processes involves reactions that
produce formation of urea-formaldehyde (UF)
resins.

Cross-References
Agrochemical Formulations
Emulsions
Polymeric Surfactant
Surfactants

Microcapsules and Controlled


Release
Microcapsule

688

Microemulsions
Julian Eastoe1, Marios Hopkins Hatzopoulos1
and Rico Tabor2
1
University of Bristol, Bristol, UK
2
School of Chemistry, Monash University,
Clayton, Australia

Synonyms
Micellar emulsion; Nanoemulsion; Swollen
micelle

Keywords
Enhanced oil recovery; Light scattering;
Microemulsions; Nanoparticles; Percolation;
Phase behavior; Self-diffusion NMR; Small-angle
neutron scattering; Soil remediation; Stability

Definition
Microemulsions are dispersions of two or more
immiscible or partially miscible fluids stabilized
by added surfactants. The dispersed domains are
generally in the nanometer size range. Visually,
they are transparent or translucent, and their
appearance does not change with time.
Microemulsions are considered to be thermodynamically stable.

Overview
Microemulsions may be considered as a subset of
emulsions, exhibiting certain unique properties.
Added surfactants perform a key role to reduce
interfacial tension sufficiently to promote thermodynamic stability of microemulsions. At the
same time, microemulsion droplets are short
lived; they are amenable to investigation by an
array of techniques, such as electrical conductivity, self-diffusion NMR, light scattering, and
small-angle neutron scattering. Microemulsions

Microemulsions

may be employed in processes of nanoparticle


synthesis, soil remediation, and enhanced oil
recovery. Microemulsion science is a steadily
developing and highly promising field with
many recent technological advancements.

Theory
Introduction
Microemulsions comprise two or more immiscible or partially miscible fluids, stabilized by
added surfactants. The surfactants must promote
sufficiently low interfacial tensions to permit the
formation of nanometer scale domains. This
characteristic nanometer length scale explains
the transparency (translucency), which is
a distinguishing feature of microemulsions. The
visual appearance of microemulsions does not
change with time. Microemulsions are distinguished from emulsions, which are generally
opaque due to of light scattering from the dispersed domains as a result of their size and refractive index difference from the continuous fluid.
The visual appearance of emulsions changes with
time as the dispersed and continuous fluids
resolve. Therefore, it can be reasonably assumed
that microemulsions are thermodynamically stable systems, whereas emulsions are thermodynamically unstable.
The visual appearance of microemulsions,
transparency (or translucency) and stability over
time, suggests that microemulsions could be considered as single-phase (monophasic) systems.
This immediately provokes consideration of
what is meant by single phase.
A phase is a region of a material that has the
same composition throughout, persisting up to
a boundary with another phase. A phase may
contain one or more components but is uniform
in composition. In the case of a pure fluid
(or a mixture), of a single phase, the time average
molecular distribution is constant. The Gibbs
phase rule (Eq. 1) indicates the number of phases
in equilibrium, based on the number of components and on the degrees of freedom. The degrees
of freedom are intensive properties of the
system which are independent from each other.

Microemulsions

689

Intensive properties are also independent of the


size of the system. For this purpose, chemical
compositions are defined in terms of mole fractions which are clearly intensive properties.
FCP2

(1)

Where F is the degrees of freedom, C the


number of components, and P is the number of
phases (Rock 1969).
It is commonly stated that microemulsions
form single phases; however, that does not hold
at all length scales, since microemulsions consist
of nanometer-sized dispersed domains in a bulk
fluid. Indeed, based on visual observations over
macroscopic length scales, microemulsions
are apparently monophasic. However, when
interrogated at microscopic length scales (i.e.,
1100 nm), because of the presence of identifiable dispersed nanometer-sized domains, it
becomes clear that microemulsions are not
monophasic but are in fact biphasic. In
microemulsions, domains (or droplets) of one
phase are dispersed in another phase, separated
by a monolayer of surfactant stabilizer molecules.
Microemulsions were first identified in the
early 1940s by Schulman et al. (Hoar and
Schulman 1943). Initially, they were referred to
as hydrophilic oleomicelles or oleophillic
hydromicelles, swollen micelles, and micellar emulsions, though these terms have largely
been discontinued nowadays. The term
microemulsion was coined in the late 1950s,
but until the mid-1970s, they were viewed as
something of a scientific curiosity with little
research being conducted on them. Research
interest picked up during the oil crisis in the
early 1970s because microemulsions can be used
in tertiary oil recovery (i.e., the partial removal of
the residual oil remaining in the well rock) but
faded again as the oil crisis receded and tertiary
oil recovery became commercially unrealistic
due to its high cost.
The unique properties of microemulsions
make them important systems for drug delivery
(Wei et al. 2005), synthesis of polymers and
high-value nanoparticles, in catalysis, enhanced
oil recovery, liquidliquid extractions, and

nanoparticle recovery (Abecassis et al. 2009;


Hollamby et al. 2010; Faizan et al. 2011), as
well as applications in pharmaceuticals (Lv
et al. 2006), detergency (Kling 1949), and lubrication (Hone et al. 2000). Therefore, the field
remains sufficiently important to continue to
attract considerable research interest. For that
reason, a great deal of progress has been made
in the last 20 years in understanding
microemulsion properties. In particular, interfacial film stability and microemulsion structures
can now be characterized in detail owing to the
development of new and powerful techniques
such as small-angle neutron scattering (SANS,
as described below). The following sections
deal with the fundamental microemulsion properties of formation and stability, surfactant films,
classification, and phase behavior.
To understand which kind of systems may
(and may not) be considered as microemulsions,
it is useful to compare them with normal
emulsions.
Emulsions and Microemulsions
Emulsions are systems of with at least one fluid
dispersed in another fluid: emulsions may be
called macro-, mini-, nano-, and microemulsions.
The fluids must be immiscible or partially
miscible. Most emulsions can be categorized as
oil-in-water (o/w) or water-in-oil (w/o), although
other more complex kinds are known. In this
field, the characteristic size of the internal
domain constitutes another important way to
distinguish between categories: microemulsions
have domains between 1 and 100 nm;
miniemulsions, between 0.1 and 1 mm; and for
macroemulsions, the internal domains (droplets)
are typically >1 mm.
Hence, the term microemulsion can be misleading, as the droplets tend to be submicron
sized being the 1100-nm-size range. As
a result, microemulsions are transparent (or translucent) which, confusingly, has also led to an
increased use of the term nanoemulsions.
Other names have been used for microemulsions
over the years, including swollen micelles and
micellar emulsions, though these terms have
largely been discontinued nowadays.

690

The principal difference between macroemulsions/miniemulsions and microemulsions


is explained in terms of stability: all other
emulsions are kinetically stable but thermodynamically unstable, whereas microemulsions are
thermodynamically stable but kinetically unstable. This is evident when it is considered how the
different systems are prepared. Emulsions
(macro-, mini-, nano-) are prepared by three
main routes:
(a) Batch comminution, which requires energy
being supplied to the two bulk fluids until
one becomes dispersed within the other.
(b) Continuous comminution by which one fluid
is forced through a capillary (or frits) emerging as droplets in the second fluid.
(c) Nucleation and growth is a method in which
two fluids are completely miscible under certain conditions. By change of one of the
intensive properties (such as temperature) in
a controlled manner, the fluids become
immiscible, and in the presence of surfaceactive stabilizers, a dispersion will form
(Vincent et al. 1998).
Emulsions are kinetically stable which means
the impact of droplet-droplet collisions is minimized. However, given time, the drops will eventually coalesce, the larger drops growing at the
expense of the smaller ones and the two fluids
finally separate into two distinct phases. The rate
of this phase separation process (emulsion resolution) depends on numerous physicochemical
factors, including the degree of miscibility (or
immiscibility) and on the viscosity of the dispersion. The process of emulsion resolution normally takes longer if the initial droplets are
smaller. To decrease the time of phase separation
and hence maintain the dispersion for longer
periods of time, stabilizers are employed. Stabilizers act by adsorbing on the interface between
the two fluids. Appropriate stabilizers can be
surfactants, polymers, and colloidal- and
nanoparticles, which serve to enhance electrostatic interactions (surfactants, polymers, particles) or steric interactions (polymers, nonionic
surfactants). The result is to promote repulsive
and hinder attractive interactions between the
dispersed fluid drops, hence retarding droplet

Microemulsions

coalescence so that the drops remain dispersed


for longer. Macroemulsions and miniemulsions
may form in the absence of stabilizers; however,
for microemulsions, the presence of surfactants is
essential. The reason for this becomes apparent
on consideration of the thermodynamics of formation (see section Formation). Under appropriate conditions and in the presence of added
stabilizers, microemulsions form spontaneously, whereas emulsions do not.
Microemulsions comprise of three (or more)
components: two immiscible or partially miscible
fluids and surfactant or polymer stabilizers. The
origin of the thermodynamic stability is the significant reduction in fluidfluid interfacial tension go/w brought about by adsorption of highly
efficient surfactants (amphiphiles). As a result of
promoting ultralow interfacial tensions (< 0.10
mN m1), larger internal surface areas are generated (surface area/volume  105 m1). It is often
found that standard surfactants (e.g., sodium
dodecylsulfate) are not capable of reducing go/w
to the required levels, and an additional
cosurfactant or cosolvent is needed, for example,
medium-chain-length aliphatic alcohols such as
pentanol and hexanol. Studies of microemulsions
indicate that chemical structure(s) of the stabilizer is all important, being one major factor
determining stability and formation. The other
factors are the chemical nature/composition of
the two fluid components and thermodynamic
variables temperature and pressure (Nave
et al. 2000a).
Thermodynamics
Formation
As stated above, the surfactant(s) lowers the
interfacial tension go/w between the two immiscible fluids sufficiently such that nanometer-sized
droplets are formed. The droplet formation can be
understood in terms of the free energy changes on
forming a dispersion DGform. Given that the dispersed phase forms small droplets, the configurational entropy change will be given by (Overbeek
1978):




1f
DSconf nkB ln f
ln1  f (2)
f

Microemulsions

691

where n and f are the number of droplets and the


volume fraction of the dispersed fluid, respectively. The free energy of formation (microemulsification) is then the sum of this entropic term
and the energy required to create new interfacial
area A, given in terms of the change in interfacial
area per droplet, DA (which would be equal to
n4pr2 for n spherical droplets of radius r) and the
interfacial tension go/w
DGform DAgo=w  TDSconf

(3)

respectively, are the only adsorbed components


(i.e., Gwater Goil 0), Eq. 4 becomes:
dgo=w Gs RTd ln Cs  Gco RTd ln Cco

(5)

Integration of Eq. 5 gives:


o
go=w go=w
Z


Cs

Cco

Gs RTd ln Cs 

Gco RTd ln Cco


0

(6)
For microemulsion formation to be spontaneous, DGform < 0, and hence DAgo/w > TDSconf. On
dispersion, the droplet number increases and so
DSconf is positive. Typically, the increase in interfacial area is of the order 104105, and the natural
fluidfluid interfacial tension of surfactant
free systems, go/w, is of the order 50 mN m1
(for the common example of wateroil). This
suggests DAgo/w is somewhere in the region of
1,000 kBT, hence, there is a need for the surfactant to lower the interfacial tension significantly,
to the region of 0.01 mN m1. Some surfactants
(double chain ionics (Kunieda and Shinoda 1980;
Chen et al. 1984) and some nonionics (Kahlweit
et al. 1990)) can produce extremely low interfacial tensions in the region of 102 to 104 mN
m1 but in most cases, such low values cannot
be achieved by a single surfactant. An effective
way to further decrease go/w, so as to promote
formation of a microemulsion, is to include a
second surface-active species (either a surfactant
or medium-chain alcohol), that is, a cosurfactant.
This can be understood in terms of the Gibbs
equation extended to multicomponent systems
(Hunter 1994). It relates the interfacial tension
to the surfactant film composition and the chemical potential, m, of each component in the system, that is,
dgwo 

X
i

Gi dmi  

Gi RTdlnCi (4)

where Ci is the molar concentration of component


i in the mixture and Gi the surface excess
(mol m2). Assuming that surfactants and cosurfactants, with concentration Cs and Cco,

Gs and Gco are the surface excesses of surfactant and cosurfactants, respectively. Equation 6
shows that go/w is decreased by both the surfactant
and cosurfactant so their effects are additive.
Stabilization

An important quantity defining the composition


of a microemulsion is the w value, which in the
case of a w/o system is defined in terms of the
molar concentrations of water and surfactant, and
is effectively the stabilization efficiency:
w

H 2 O
AOT 

(7)

For o/w systems, the water concentration in


Eq. 7 would be replaced by the molar concentration of the dispersed oily component.
Based on simple spherical droplet geometry,
an approximate droplet (domain) size can then be
calculated from w, as pointed out by Hilfiker et al.
(1987):
r

3wvw
lc
ao

(8)

where r is the radius of the microemulsion droplet, vw is the volume of a dispersed phase mole 3 in the case of water in w/o systems,
cule ( 30 A
but depending on molecular structure for O/W
counterparts), and lc and ao are the length of the
fully extended surfactant hydrophobic chain and
the area per head group of the surfactant, respectively. The range of w values accessible is dependent on the surfactant and solvent combination;

692

Microemulsions,
Fig. 1 Phase behavior of
AOT-stabilized w/o
microemulsions. A:
(a) n-octane, (b) isooctane,
(c) oct-1-ene; B: cyclic
alkanes and decalin,
(D) at fixed
[AOT] 0.10 mol dm3
(Reprinted from Fletcher
et al. 1987 copyright RSC)

Microemulsions

(b)
100 (b)
(a)
(a)

100
(c)

(c)

C7
C8 D C6 C8 C7 C6
D

R
50

50

15

30

some systems may only support very low w values


(below 10), but for others, values of 7080 are not
uncommon. One added advantage in introducing
w is that it can simplify the presentation of the
phase behavior of microemulsion systems, by
reducing the number of degrees of freedom, F,
in the Gibbs phase rule (Eq. 1). For instance,
a ternary compositional plot (water-oil-surfactant) can be reduced to a pseudobinary plot by
fixing the surfactant concentration and varying
the amount of water i.e., a convenient and
useful plot of w versus temperature T could then
be presented. This is shown in Fig. 1 for the case
of w/o microemulsions in some common organic
solvents, stabilized by Aerosol-OT (AOT,
sodium bis(2-ethylhexyl)sulfosuccinate).
These phase diagrams show a typical funnel
shape, which is characteristic of AOT
microemulsions. The lower phase boundary is
termed the water solubilization boundary, where
the macroscopically single phase system will
undergo separation to form a so-called Winsor II
system that is, separation into a water-in-oil
microemulsion phase in equilibrium with an
excess water part. At the upper temperature
phase boundary, the transition is into a Winsor I
system which is an oil-in-water droplet phase in
equilibrium with excess oil. The two alternative
possible microemulsion phases are a Winsor III
system, with a (bicontinuous) microemulsion
phase in equilibrium with both excess water and

45

0
T/C

20

40

60

excess oil, and a Winsor IV system, which is


a macroscopically single microemulsion.
Kinetic Aspects
This structural and compositional complexity of
microemulsions means that there are potentially
numerous relaxation processes. Despite this,
microemulsion kinetics have been relatively
well researched, and there have been several
important reviews of the area (Lopez-Quintela
et al. 2004a; Moulik and Paul 1998). The
dynamic processes which can occur can be summarized as: (1) motion of interfacial molecules,
(2) exchange of droplet contents, and (3) coalescence of droplets. Motion of interfacial molecules refers to the mobility of surfactant
molecules adsorbed at the interface. Ahlnas
determined that this occurs on the picosecond
timescale (Ahlnas et al. 1983). Exchange processes can be of surfactant, cosurfactant, or dispersed solvent. For dispersed water droplets in
w/o systems, the exchange rate of immobilized or
bound water (in a layer associated with the
polar surfactant head groups) and free water was
determined to be on the millisecond timescale
using NMR (Hansen 1960).
Exchange rates of cosurfactant molecules
between interface and bulk solution seem to
vary with molecular size (mobility) and environment, but a typical rate of around 108 s1
was found by Lang et al. for butan-1-ol in

Microemulsions

water/toluene/SDS microemulsions (Lang et al.


1975). Exchange of surfactant molecules may be
on a similar timescale but seems to be greatly
affected by many other parameters (Tondre
2005). It also seems that the exchange times
found in different studies are dependent on the
method of measurement, so the picture is not yet
complete. Exchange of material in droplets seems
to occur via a process of coalescence and then
rapid fission, with data supporting this theory
coming from Abe and coworkers (1987). They
found when mixing two three-component
microemulsions with differently sized droplets
a unimodal distribution was spontaneously
reached without stirring, demonstrating that
there must be an exchange process occurring
more rapidly than Ostwald ripening, which is
most plausibly interpreted as fusionfission. The
seminal paper in this field came from Fletcher,
Howe, and Robinson, who emphasized that
droplet exchange (via coalescence/fission) is
rapid and continuous (Fletcher et al. 1987).
They were even able to determine the activation
barrier to water exchange for AOT-stabilized
water-in-oil microemulsions (of the order of
70100 kJ mol1), with an associated
rate constant for water exchange of
106108 mol1 dm3 s1.
Therefore, for (typical) millimolar concentrations of droplets, the internal contents of the
microemulsion domains exchange, typically on
the millisecond timescale (Fletcher et al. 1987,
1990). If collisions are sufficiently violent, then
the surfactant film may rupture thereby facilitating droplet exchange, which is why the droplets
are kinetically unstable. Hence, there is an associated activation energy for droplet fusion. The
mechanism of droplet coalescence has been
reported for AOT w/o microemulsions (Fletcher
et al. 1987); the droplet exchange process was
characterized by a second-order rate constant kex,
which appears to be activation controlled (Milner
and Safran 1987a) (hence the activation energy,
Ea, barrier to fusion) and not purely diffusion
controlled. Other studies (Bancroft 1913a)
have shown that the dynamic aspects of
microemulsions are affected by the flexibility
of the interfacial film, that is, film rigidity

693

(see section Film Bending Rigidity), through


a significant contribution to the energy barrier.
Under the same experimental conditions, different microemulsion systems can have different kex
values (Fletcher et al. 1987): for AOT w/o system
at room temperature, kex is in the range
106109 dm3 mol1 s1, and for nonionics CiEj,
108109 dm3 mol1 s1 (Fletcher et al. 1987,
1990; Bancroft 1913a).
Predicting Microemulsion Type

A further microemulsion categorization (other


than w/o and o/w) is the aforementioned system
proposed by Winsor (1948) who introduced four
types of phase equilibria where two (or more)
phase portions are present in equilibrium:
Type I: The system consists of a small portion
of organic component (oil) and surfactant
coexisting with the larger volume of
a microemulsion dominated by the polar fluid
(water): for example, a oil-in-water (o/w)
microemulsion in equilibrium with excess dispersed oil (Winsor I).
Type II: The system is in effect the reverse of
type I, for example, with a w/o microemulsion
coexisting in equilibrium with excess water
(Winsor II).
Type III: For common water + oil
microemulsions, this consists of a three
regions, with the top region the oil phase and
the bottom phase being the polar fluid phase.
The middle region is a microemulsion stabilized by the majority of surfactant present
(Winsor III or middle-phase microemulsion).
Type IV: It is a homogenous system, comprising only one macroscopic phase (biphasic
on nanometer-length scales, see above)
(Winsor IV).
Phase transitions from Winsor I to Winsor II
via Winsor III phases can brought about by
increasing either electrolyte concentration (in
the case of ionic surfactants) or temperature (for
nonionic amphiphiles). To account for the type of
emulsion formed by a particular surfactant, Bancroft (1913a) and Clowes (1916) considered the
adsorbed film in emulsion systems to have an
inner and an outer interfacial tension acting independently (Adamson 1960). The interface would

694

Microemulsions

curve in such a way as for the inner surface to


always have the higher tension. Bancrofts rule
states the type of emulsion formed is dictated by
the preferred solubility of surfactant (i.e., the
tendency of a given surfactant to partition more
or less strongly into one or other of the fluids). If
the surfactant is preferentially water soluble, then
an oil-in-water (micro)emulsion will form, and
conversely if the surfactant is preferentially oil
soluble, then a water-in-oil (micro)emulsion is
expected. Other approaches supporting the Winsor and Bancroft classifications are discussed
below.

AHH, the interaction of the hydrophilic parts of


the surfactant molecules
The above parameters oppose the miscibility
of the surfactant in oil and water. For the (micro)
emulsion to prove stable, that is, to prevent phase
separation from occurring, the difference
between Aso and ASW should not be large. In
other words, the surfactant should have sufficient
miscibility in both fluids. Winsor expressed this
balance as:

The R Ratio

An extension of the theory (Bourrel and


Schechter 1988) accounts for all the cohesive
energies with:

Winsor (1948) proposed the R ratio as a means to


determine the type of (micro)emulsion formed.
The R ratio (Eq. 12) considers the strength of
interaction of the amphiphile with the two fluids
hence the dictating type of emulsion. The total
interaction of surfactant with the two fluids (Axy)
can be expressed as:
Axy ALxy AHxy

(9)

where ALxy and AHxy are the interactions of the


nonpolar parts of the two molecules (dispersion
forces) and the polar interactions (such as hydrogen bonding and Coulombic interactions),
respectively. In the case of an oilwater system,
the interactions are:
Aso ALso AHco

(10)

ASW ALsw AHsw

(11)

where AHco and ALsw are small and often ignored.


Aso is the cohesive energy responsible for the
solvation of the surfactant in oil. ASW is that
responsible for solvation in water. Other cohesive
interactions are the following:
Aww, the interaction of water molecules between
themselves
Aoo, the interaction of oil molecules between
themselves
ALL, the interactions of the hydrophobic parts of
the surfactant molecules

Aso
ASW

Aso  Aoo  ALL


ASW  Aww  Aww

(12)

(13)

Based on the R ratio, the type of emulsion


formed can be predicted. When R > 1, the interface increases its interaction with the oil phase at
the expense of water interaction and hence a w/o
emulsion forms. When R < 1, o/w emulsions
form, and when R 1, a type III or IV system
may form.
HydrophilicLipophilic Balance (HLB)

Another concept relating molecular structure to


interfacial packing and film curvature is hydrophiliclipophilic balance (HLB). The HLB scale
supports Bancrofts rule (Bancroft 1913b;
McBain 1950) which states that the type of emulsion formed is dependent on the preferential
partitioning of the emulsifier in either one of the
fluids. If the emulsifier is soluble in water more so
than in oil, then o/w emulsions form and vice
versa. The HLB is generally expressed as an
empirical equation based on the relative proportions of hydrophobic and hydrophilic groups
within the molecule. As such, HLB is an attempt
at quantifying hydrophilicity or lipophilicity
(hydrophobicity) of any given amphiphile. In
the early days, the majority of work on the HLB
was carried out on nonionic surfactants which

Microemulsions

695

were of primary industrial importance. Griffin


(1949) characterized a number of surfactants
and derived an empirical equation for nonionic
alkyl polyglycol ethers (CiEj) based on the surfactant chemical composition (Griffin 1954):
HLB

Ej OH wt %
5

(14)

where Ej wt% and OH wt% are the weight percent


of ethylene oxide and hydroxide groups,
respectively.
For bicontinuous structures, that is, zero curvature, it was shown that HLB  10 (Israelachvili
1994). Then, w/o (micro)emulsions form when
HLB < 10, and o/w (micro)emulsions when
HLB > 10.
Davies (1957) attempted to predict the type of
emulsion formed when water and oil were agitated in the presence of an emulsifier. Daviess
approach accounted for both ionic and nonionic
surfactants. The experiments conducted of relative coalescence rates for a particular oil and
surfactant, so that the rates of coalescence of an
oil drop and that of a water drop to the interface
were measured. From these rates, hydrophilic and
hydrophobic numbers were derived which were
proportional to the number of hydrophobic and
hydrophilic units of the surfactant. Introducing
these to Eq. 15, the HLB of the surfactant can
be calculated with good agreement in most cases
to those derived by Griffin.
HLB 7 SnH  SnL

(15)

This approach is popular in industry to guide


the preparation of (micro)emulsions and for the
choice of surfactants for a particular (micro)
emulsion system. Equation 15 can be rewritten
as:
HLB 7
SHydrophilic group numbers
 ngroup number per  CH2  (16)
where n is the number of hydrophobic carbons in
the alkyl tail of the single chain surfactant, as

methyl and methylene groups are assigned the


same group number.
Lin (Lin et al. 1973) tried to extend Daviess
work by relating HLB to the cmc of the amphiphiles. It was noted by Klevens et al. (Klevens
1953) that
logcmc a  bn

(17)

Lin et al. (1973) combined Eqs. 16 and 17


leading to
logcmc a0 b0 HLB

(18)

where a and b are empirical constants for


a homologous series and from which a0 and b0
can be derived. Given a value for cmc, the HLB of
any member of a surfactant homologous series
can be calculated. Shinoda (Shinoda and Friberg
1986) however issued warning at this correlation
of the cmc to the HLB, as the latter can stay
constant with size of hydrophile and lipophile
while the cmc will increase geometrically. However, Lin only examined ionic surfactants and
their aqueous cmcs with the derived HLB numbers were in very reasonable agreement with
those predicted by previous methods.
Phase Inversion Temperature (PIT) or HLB
Temperature

HLB has been a widely used tool in the preparation of emulsions. However, it primarily deals
with the solubility of the surfactant in a general
water and oil emulsion. The HLB method, as
discussed in the previous section, does not
account for the behavior of the surfactant in different oils, polar solvents, temperatures, and
additives in the polar and nonpolar fluids.
Shinoda (Shinoda and Saito 1969) used
a characteristic property of the emulsions as an
indicator of the HLB of the surfactant. Nonionic
surfactants form microemulsions and emulsions
that are highly sensitive to temperature. At
a particular, system-specific point, known as the
phase inversion temperature (PIT) or HLB temperature, the surfactant film curvature (explained
below) changes from positive (around oil) to

696

negative (around water), and phase inversion


takes place from o/w to w/o. The origin of this
structural inversion is thought to be a result of
changes in the hydration of the hydrophilic surfactant moieties, which generally weakens with
increasing temperature. The reduced hydration
then effectively decreases the HLB value, as
pointed out by Shinoda et al. (Shinoda and Saito
1969):
If T < PIT, an oil-in-water microemulsion
forms (Winsor I)
I T > PIT, a water-in-oil microemulsion forms
(Winsor II)
At T PIT, a middle-phase microemulsion
exists (Winsor III)

Microemulsions

than fully extended equivalent hydrocarbon chain


length lmax. Therefore,
lc  lmax

(19)

According to Tanford (1978), the volume of


liquid hydrocarbon chains are essentially the
same as those for the same fragment in micellar
aggregates, as determined by X-ray scattering.
Assuming additivity and exclusion of water
from the hydrophobic core, the volume is proportional to the number of surfactants through Nagg,
and the volumes of the hydrocarbon tails further.
Hence,


v nm3 0:0274 0:0269nc

(20)

lmaxnm 0:154 0:126nc

(21)

Packing Parameter and Microemulsion Structures

Recall that the chemical nature of the surfactant is


one of the important factors determining the type
of (micro)emulsion formed. Israelachvili et al.
(1976) studied changes in film curvature that
followed certain geometrical parameters of
the surfactant molecules. When considering surfactant in aqueous systems, the interactions
at play are the hydrophobic effect of the surfactant tails promoting molecular aggregation and
the repulsive electrostatic and steric effects of
the head groups. As such, the hydrophobic
effect works toward decreasing the contact
between the polar environment, while the electrostatic and steric forces work toward increasing contact with the polar fluid. The most
favored geometry is formed when these interactions are balanced.
To introduce the concept of packing parameter, consider first of all an aqueous micelle,
comprising Nagg surfactant molecules. Since the
structure is built up of many individual molecules, it can be appreciated that the most favored
aggregation geometry will be governed by
a balance of intermolecular interactions, being
dominated by an optimal head group area ao
(representing repulsive electrostatic and/or steric
interactions) and the hydrocarbon tail volume v
(being the limiting volume of more weakly
interacting and incompressible hydrocarbon
chains). The tails are also assumed to be at
a maximum extension lc, which is slightly less

and

with nc being the number of carbon atoms in the


hydrocarbon chain and v and lmax having units of
nm3and nm, respectively. With these variables
considered, then the packing parameter can be
defined:
P

v
ao lc

(22)

Geometrical constraints dictate:


P  13 spherical micelles
1
1
3 < P  2 cylindrical or rod-like micelles
1
2 < P  1 vesicles and flexible bilayers
P 1 planar extended bilayers
P > 1 reverse micelles
The conditions that the aggregate must satisfy
irrespective of its shape are the following:
(a) No point in the aggregate can be further than
lc from the water interface.
(b) The total hydrocarbon volume or the aggregate and the total interface area must follow
the relationship Vv aAo N agg , where Nagg is
the aggregation number.
When discussing microemulsions, the above
conditions cease to hold. For instance, when oil is
introduced to a surfactantwater system where
the aggregates are cylindrical (and as a result

Microemulsions

stressed), the aggregate will become spherical,


thus
favoring
an
o/w
microemulsion
(Israelachvili et al. 1976).
If ao > v/lc, oil-in-water microemulsion.
If ao < v/lc, water-in-oil microemulsion.
If ao  v/lc, a middle-phase microemulsion is
the preferred system.
Becher (1984) argued that the packing parameter, HLB, and R ratio approaches to understating
microemulsions were all consistent. The packing
parameter could be considered as an inverse HLB
number expressed in terms of volume fractions of
the hydrophobic parts.
Surfactant Film Properties

A further approach to understanding the formation of a particular type of (micro)emulsion is to


consider the mechanical properties of a surfactant
film at an oilwater interface. The three phenomenological constants governing the film behavior
are the tension, bending rigidity, and spontaneous
curvature. The extent of influence of each of these
parameters depends on the constraints felt by the
film. These parameters are of importance since
the film behavior governs the microemulsion
behavior and response in terms of phase behavior
and stability, structure, and solubilization
capacity.
Ultralow Interfacial Tension In creating
a microemulsion, the interfacial area rises by 45
orders of magnitude, and (as stated above) to permit
this, the interfacial tension needs to be considerably
lowered. Microemulsion formation requires
ultralow interfacial oilwater tensions, go/w, usually
of the order of 102 to 104 mN m1. Typical
surfactants are incapable of reaching such low
interfacial tension values, and for this reason,
addition of cosurfactant as well as electrolyte
and variation of temperature, pressure, and oil
chain length are required, often in combination.
Several studies have been reported on the effects
of composition variables on go/w. In particular,
Aveyard and coworkers performed systematic
interfacial tension studies with both ionic
(Aveyard et al. 1986a, b) and nonionic surfactants
(Aveyard et al. 1989), varying oil chain length,
temperature, and electrolyte (NaCl) content. For

697

example, in the system waterAOTn-heptane, at


constant surfactant concentration (above its cmc),
a plot of go/w as a function of electrolyte concentration shows a deep minimum that corresponds
to the Winsor phase inversion, that is, upon addition of NaCl, go/w decreases to a minimum critical
value (Winsor III structure) and then increases to
a limiting value close to 0.20.3 mN m1 (Winsor II region). At constant electrolyte concentration, varying temperature (Aveyard et al. 1986a),
oil chain length, and cosurfactant content
(Aveyard et al. 1986b) have a similar effect.
With nonionic surfactants, a similar tension
curve and phase inversion were observed, but
instead with increasing temperature rather than
[NaCl] (Aveyard et al. 1989). In addition, when
increasing surfactant chain length, the interfacial
tension curves shift to higher temperatures and
the minimum in go/w decreases (Sottman and
Strey 1996). Ultralow interfacial tensions cannot
be measured with standard techniques such as Du
Nouy ring, Wilhelmy plate, or drop volume
(DVT), and spinning drop tensiometry (SDT) or
surface light scattering (Langevin 1992) must be
used.
Spontaneous Curvature Stability and structural transformations in microemulsions can
also be understood by considering the energy
associated with forming and bending of stabilizing interfacial films. In fact, this surface energy
contribution may be nonnegligible, because
interfacial areas in microemulsions are very
large (S/V  105 m1).
Spontaneous (natural or preferred) curvature
Co is defined as the curvature adopted a surfactant
film in the presence of equal amounts of water
and oil. This equilibrium condition imposes no
constraints on the film, which is then free to attain
the lowest free energy state. Relaxing the condition that the oil/water ratio is unity means
a deviation from Co. Every point on a surface
possesses two principal radii of curvature, R1
and R2, so that the associated principal curvatures
are C1 1/R1 and C2 1/R2 (see Fig. 2). The
separate mean and Gaussian curvatures are used
to define the bending of surfaces, defined as follows (Hyde et al. 1997):

698

Microemulsions

Microemulsions,
Fig. 2 Principal
curvatures of different
surfaces. (a) Intersection of
a surfactant film surface
with planes containing the
normal vector (n) to the
surface at the point p.
(b) convex curvature,
(c) cylindrical curvature,
(d) saddle-shaped
curvature (After Hyde et al.
1997)

R1

R2

R1 =
R2 > 0

R1 > 0
R2 > 0

Mean curvature : C 1=21=R1 1=R2 (23)


Gaussian curvature : k 1=R1  1=R2

(24)

If a circle is placed tangentially to a point p on


the surface and if the circle radius is chosen so
that its second derivative at the contact point
equals that of the surface in the direction of the
tangent (of normal vector, n), then the radius of
the circle is a radius of curvature of the surface.
The curvature of a surface is described by two
such circles chosen in orthogonal (principal)
directions as shown in Fig. 2a. For a sphere, R1
and R2 are equal and positive (Fig. 2b): for a cylindrical surface, R2 is indefinite (Fig. 2c); whereas,
for a plane, both R1 and R2 are indefinite. One
special case is a saddle structure, R1 R2, that
is, at every point the surface is both concave and
convex (Fig. 2d), so that both a plane and saddle
have the property of zero mean curvature.

R1 < 0
R2 > 0

In this model, the curvature of the surface Co


depends both on the composition of the phases it
separates and on surfactant type. Considering the
apolar side of the interface, it is possible for oil
molecules to penetrate to some extent between
the surfactant hydrocarbon tails. The more extensive the penetration, the more curvature is
imposed toward the polar side. This would result
in a decrease of Co since, by convention, positive
curvature is toward oil (negative toward water).
The longer the oil chains, the less they penetrate
the surfactant film and the smaller the effect on
Co. Eastoe et al. have studied the extent of solvent
penetration in microemulsions stabilized by
dichained surfactants, using SANS with selective
deuteration. Results suggested that oil penetration is only a subtle effect, depending on the
chemical structures and architecture of both
surfactant and oil. In particular, unequal surfactant chain lengths for d-chain compounds,

Microemulsions

(Eastoe et al. 1996a, b, 1997a, b) or the presence


of C C bonds (Bumajdad et al. 1998) result in
a more disordered surfactant/oil interface,
thereby providing a region of enhanced oil
mixing. For symmetric dichained surfactants
(e.g., didodecyldimethylammonium bromide
DDAB and the anionic AOT), however, no evidence for oil mixing into the interfacial region
was found (Eastoe et al. 1997a). The effect of
alkane structure and molecular volume on the oil
penetration was also investigated with n-heptane
and cyclohexane. The results indicate that heptane is essentially absent from the layers, but the
more compact cyclohexane has a slightly greater
penetrating effect (Eastoe et al. 1997b).
Surfactant type, and nature of the polar head
group, also influences Co through different interactions with the polar (aqueous) phase:
For ionic surfactants, electrolyte content and
temperature affect Co in opposite ways.
Increasing salt concentration screens electrostatic head group repulsions i.e., decreases
head group area so the film curves more
easily toward water, leading to a decrease in
Co. Increasing temperature has two effects:
(1) an increase in electrostatic repulsions
between head groups due to higher counterion
dissociation, so Co tends to increase, and
(2) more gauche conformations are induced
in the surfactant chains, which become more
coiled, decreasing Co. Therefore, the combined effects of temperature on the apolar
chains and on electrostatic interactions are
competitive. Although the electrostatic term
is believed to be slightly dominant, Co also
increases weakly with increasing temperature.
For nonionic surfactants, unsurprisingly, electrolytes have very little effect on Co, whereas
temperature is a critical parameter due to the
strong dependence on ethyleneoxide head group
hydration. For nonionic CiEj alkyl polyethyleneoxide surfactants, as temperature increases,
water becomes a poorer solvent for the hydrophilic units and penetrates less into the surfactant
layer. In addition, on the other side of the film, oil
can penetrate further into the hydrocarbon chains,
so that increasing temperature for this type of
surfactant causes a strong decrease in Co.

699

Thus, by changing parameters such as temperature and pressure, the nature of the oil, or electrolyte concentration, the spontaneous curvature
can be tuned, so driving transitions between Winsor systems.
Film Bending Rigidity The film bending
energy concept was first introduced by Helfrich
(1973) and is now considered as an important
model for understanding microemulsions. The
interfacial film is described by two elastic moduli
(Kellay et al. 1994), which are measures of the
energy required to deform the interfacial film
from Co:
The mean bending elasticity (or rigidity), K,
associated with the mean curvature, which
represents the energy required to bend unit
area of surface by unit amount. K is positive,
that is, spontaneous curvature is favored.
On the other hand, the factor K is associated
with Gaussian curvature, and hence accounting for film topology. K is negative for spherical structures or positive for bicontinuous
cubic phases.
Theoretically, it is expected that bending moduli should depend on surfactant chain length
(Safran and Tlusty 1996), area per surfactant
molecule in the film (Szleifer et al. 1990), and
electrostatic
head
group
interactions
(Winterhalter and Helfrich 1992).
The film rigidity theory is based on the interfacial free energy associated with film curvature.
The total free energy, F, of a surfactant layer at
a liquid interface can be considered a composite
of an interfacial energy term, Fi, a bending
energy term, Fb, and an entropic term, Fent. For
a droplet type structure, the total can be written as
(Gradzielski et al. 1996):
F Fi Fb Fent gA

Z 
K
2

C1 C2  2C0 kC1 C2 dA (25)

2
nkB Tf f
where g is the interfacial tension, A is the total
surface area of the film, K is the mean elastic
bending modulus, K is the Gaussian bending

700

Microemulsions

modulus, C1 and C2 are the two principal curvatures, Co the spontaneous curvature, n is the number of droplets, kB is the Boltzmann constant, and
f(f) is a function accounting for the entropy of
mixing of microemulsion droplets, where is the
droplet core volume fraction. For dilute systems
where f < 0.1, it was shown that f(f)(ln(f)1)
(Gradzielski et al. 1996). As noted above,
microemulsion formation is associated with
ultralow interfacial tension, go/w so the go/w A
term is small compared to Fb and Fent and can be
ignored as an approximation.
The curvatures C1, C2, and Co can be
expressed in terms of radii as 1/R1, 1/R2, and
1/Ro, respectively. For spherical droplets,
R1 R2 R, and the interfacial area is
A n4pR2 . Note that R and R0 are core radii
rather than droplet radii (Gradzielski et al.
1996). Solving Eq. 25 and dividing by area A,
the total free energy, F, for spherical droplets
(of radius R) is expressed as:




F
1
1 2 K
kB T
2K

2
f

(26)
A
R R0
R
4pR2
For systems at a solubilization boundary (WI
or WII region), the droplets have achieved
their maximum size Rav
max . Under this condition,
minimization of the total free energy leads
to a relation between the spontaneous radius,
R0, and the elastic constants K and K
(Safran 1992):
Rav
2K K kB T
max

f f

2K
8pK
R0

(27)

A number of techniques have been used to


measure K and K separately, in particular,
ellipsometry, X-ray reflectivity, and small-angle
X-ray scattering (SAXS) (Meunier and Lee 1991;
Kegel et al. 1995; Sicoli et al. 1993). De Gennes
and Taupin (1982) have developed a model
for bicontinuous microemulsions. For Co 0,
the layer is (Tondre and Derouiche 1991) supposed to be flat in the absence of thermal fluctuations. They introduced the term xK, the
persistence length of the surfactant layer that
relates to K via:

xK aexp2pK =kB T

(28)

where a is a molecular length and xK is the correlation length for the normals to the layer, that is,
the distance over which this layer remains flat in
the presence of thermal fluctuations. xK is
extremely sensitive to the magnitude of K. Experimental data show that K is typically between 100
kBT for condensed insoluble monolayers
(Daillant et al. 1989) and about 10 kBT for lipid
bilayers (Schneider et al. 1984; Engelhardt et al.
1985; Bivas et al. 1987) but can decrease below
kBT in microemulsion systems (Di Meglio et al.
1985; Farago et al. 1990).
An alternative, more accessible, method to
quantify film rigidities is to calculate the com which may be done
posite parameter (2K K),
via tensiometry and SANS measurements. Two
expressions can be derived from Eqs. 26 and 27.
1. From interfacial tension go/w and the maximum mean core radius Rav
max (measured by
SANS):
2K K
kB T
go=w 
  av 2 f f
av 2
Rmax
4p Rmax

(29)

which gives for the bending moduli:



2 k B T
f f
2K K go=w Rav

max
4p

(30)

2. Using the Schultz polydispersity width



r s Rav
max obtained from SANS data analysis:
Thermal fluctuations of the microemulsion
droplets drive droplet polydispersity, which
relates to the bending moduli. Safran (1983)
and Milner (Milner and Safran 1987b)
described the thermal fluctuations by an
expansion of the droplet deformation in
terms of spherical harmonics. In the case of
the two-phase equilibria at maximum solubilization (WI or WII), this polydispersity, p,

may be expressed as a function of K and K:
p2

u2o
kB T


4p 8p2K K 2kB Tf f

(31)

where uo is the fluctuation amplitude for the


I 0 mode. This polydispersity is given by the

Microemulsions

701

SANS Schultz polydispersity parameter



s Rav
max (Eastoe et al. 1997c), and Eq. 31 can
be written:
2K K

kB T
kB T
f f
 
2 
av
4p
8p s Rmax

(32)

Now, Eqs. 30 and 32 give two experimentally



accessible expressions for the sum (2K K)
using data from SANS and tensiometry. The
 has been determined for
composite (2K K)
nonionic films in WI systems (Gradzielski et al.
1996; Gradzielski and Langevin 1996) and also
cationic (Eastoe et al. 1997c) and zwitterionic
(Eastoe and Sharpe 1997) layers in WII
microemulsions, as a function of surfactant
alkyl carbon number n-C. These trends are in
line with statistical mechanical theories
(Szleifer et al. 1990), indicating K should vary
as n-C2.5 to n-C3, whereas there is only a small

effect on K.
Phase Behavior
Microemulsion phase stability depends on the
nature and concentration of the components and
also on the thermodynamic variables pressure
and temperature. Phase stability diagrams (or
phase maps), and location of the different structures formed within these mixed systems in terms
of variables of interest are, therefore, very important. Several types of phase diagram can be generated depending on the number of variables
under consideration. In using an appropriate
graphical representation, it is possible to describe
the limits of existence of the single and
multiphase regions and to characterize equilibria
between phases (tie-lines, tie-triangles, critical
points, etc.).
Microemulsion Equilibria and the Phase Rule

The phase rule (Eq. 1) enables the identification


of the number of variables (or degrees of freedom) depending on the system composition and
conditions. In the following, the phase rule is
applied strictly to the macroscopic appearance
of microemulsions (rather than the micro-/
nanoscopic structure).

Microemulsions, Fig. 3 The phase prism describing the


phase space for a system at constant pressure

A system is called invariant, monovariant,


bivariant, and so on, according to whether the
number of degrees of freedom F is zero, 1, 2,
and so on. For example, in the simplest case of
a system composed of three components and two
phases, F is unity at a fixed temperature and
pressure. This means that the mole or weight
fraction of one component in one of the phases
can be specified to define the equilibrium, but all
other compositions in both phases are fixed. In
general, microemulsions contain at least three
components: oil (O), water (W), and amphiphile
(S), and as mentioned previously a cosurfactant
(alcohol) and/or an electrolyte are usually added
to tune the system stability. These can be considered as simple OWS systems: if a cosurfactant
is used, and the ratio oil/alcohol is kept constant,
and assuming that alcohol does not interact with
any other component, then the mixture can be
treated (to a first approximation) as a threecomponent system. Hence, at constant pressure,
the compositiontemperature phase behavior can
be presented in terms of a phase prism, as illustrated in Fig. 3.
From an experimental point of view, the construction of 3-D phase maps is rather complex
and time consuming, so it is convenient to simplify the system by selecting only specific phase
cuts. Then, the number of variables F can be
reduced, either by keeping one term constant
and/or by combining two or more variables.
Any system with a composition within the
two-phase region (e.g., point o in Fig. 4a, c) will
exist as two phases with compositions linked at

702

Microemulsions,
Fig. 4 Ternary diagram
representations of two- and
three-phase regions formed
by simple wateroil
surfactant systems at
constant temperature and
pressure. (a) Winsor I type,
(b) Winsor II type,
(c) Winsor III type systems

Microemulsions

surfactant

surfactant

2
A
water

1
2 m

c surfactant

c
oil

p
n
oil

water

n
water

m 2
oil

Salinity

Winsor type

the ends of the tie-line, that is, a segment


formed by phases m and n. Therefore, every
point on a particular tie-line has identical
coexisting phases (m and n) but of different relative volumes. When the two conjugate phases
have the same composition (m n), this corresponds to a critical (or plait) point, p.
If three phases coexist (Fig. 4b) corresponding
to a WIII system, at constant T and P, there are no
independent degrees of freedom. This region of
three-phase invariant compositions is triangular
in form and called tie-triangle (Bourrel and
Schechter 1988). Any overall composition, such
as point q (Fig. 4b), lying within the tie-triangle
will divide into three phases having compositions
corresponding to the vertices A, B, and C of the
triangle.
Binary Phase Diagrams

As mentioned previously, ternary diagrams can


be further simplified by fixing some parameters
and/or combining two (or more) variables
together (examples are: defining water and electrolyte brine at a fixed concentration or combining water and oil into a fixed water-to-oil
ratio). The effect is to conveniently reduce the
degrees of freedom F, permitting determination
of a simplified phase diagram which reduces to
a study of a planar section through the phase

III

II

prism. Examples of typical pseudobinary diagrams are given in Figs. 57 for nonionic and
anionic surfactants.
The third example (Fig. 7) concerns the
anionic surfactant Aerosol-OT, which can be
conveniently studies in water-in-oil system by
maintaining a fixed surfactant concentration, giving F 2, for a single-phase region at constant
pressure. Then, the two experimental variables
are temperature and w, the water-to-surfactant
molar ratio defined as w [water]/[surfactant].

Characterization
Methods and Characterization
Electrical Conductivity
Percolation is a phenomenon where conductivity
(s) of a reversed (w/o) microemulsion increases
rapidly by orders of magnitude when a specific
threshold has been crossed. As such, percolation
can be induced by varying water content, ionic
strength of the dispersed water and temperature.
There are two proposed explanations of the phenomenon, static percolation and dynamic percolation. The static percolation model attributes the
increase in conductivity to the formation of system-spanning water channels above the percolation threshold. In dynamic percolation, ions cross

Microemulsions

Microemulsions, Fig. 5 shows the schematic phase diagram for a nonionic surfactantwateroil ternary system

the surfactant layer when the attractive


interdroplet interactions, which arise, become
enhanced. However, the exact mechanism is
still under debate. As such, the study of percolation can give insight into interfacial surfactant
film properties:
(a) Temperature has been found to induce percolation for systems which would otherwise not
exhibit significant conductivity (Borkovec
et al. 1988). For ionic surfactants, temperature increases the level of interdroplet interactions, and hence there is an increase in
conductivity. This however does not hold
for nonionic surfactants where the reverse is
observed: conductivity decreases with
increasing temperature (Eicke and Meier
1996). This behavior was also observed for
certain microemulsions with ionic surfactants
(John and Rakshit 1994).
(b) Salinity has been shown to decrease conductivity in microemulsions stabilized by ionic
surfactants. Tondre et al. (Tondre and
Derouiche 1991) examined water/AOT/decane microemulsions with increasing electrolyte concentration from 0 to 0.3 M. They
found that the maximum water content wmax

703

sharply increased and then dropped abruptly


leading to a Winsor II system, at which point
the conductivity dropped significantly. In
examining the effect of cation size, GarciaRio et al. (1994) found that the temperature of
percolation increased with increasing
hydrated cation radius.
(c) Organic salt additives are of particular interest since microemulsions may be employed
in drug delivery and extraction processes.
Furthermore, these actives often show
a degree of surface activity, and thus might
coadsorb into the surfactant film, thereby
affecting the properties. Hydrotropes are
such a case, and many drugs can be broadly
classed as hydrotropic. Moulik et al. (Ray
et al. 1993) studied the effect of a variety of
hydrotropes and of the bile salt sodium cholate on percolation dynamics and clustering
energetics in water/AOT/heptane (Ray et al.
1993), water/AOT/isooctane, and water/
AOT/decane w/o microemulsions (Hait
et al. 2001). The maximum gradient of the
sigmoidal s  f and s  y plots is considered
the transition point of the percolation process.
The points at which these occur are known as
the threshold volume fraction ft and temperature yt. Composition, pressure, and additives
can have an effect on the both ft and yt. The
threshold temperature yt was shown to be
water-content dependent. The yt decreased
with increasing w in both heptane and decane
microemulsions in the absence of additives
Aromatic solutes such as sodium salicylate
were shown to increase both yt and yt, hence
affecting the onset of percolation. On the
other hand, bile salts decrease ft and yt, by
assisting in channel formation. Hydroquinone (Hq), pyrogallol (Pg), resorcinol (Rc),
and catechol (Cc) reduced the yt, in the order
Hq < Pg < Rc < Cc, by assisting the
droplet fusion. Sodium salicylate, a-napthol,
and b-napthol blocked droplet fusion,
thereby increasing yt. Hait et al. (2002) investigated percolation in water/AOT/decane
microemulsions containing hydroxyl and
methoxy analogues of phenol. The monohydroxy compounds increased yt, whereas

704

Microemulsions

Microemulsions,
Fig. 6 Binary phase
behavior in ternary
microemulsion systems
formed with nonionic
surfactants. (a) Illustration
of a section at constant
surfactant concentration
through the phase prism.
(b) Schematic phase
diagram, plotted as
temperature versus volume
fraction of oil, jo, at
constant surfactant
concentration. Also shown
are various microstructures
found in different
microemulsion regions,
M. At higher temperatures,
a liquid phase is in
equilibrium with excess
water (M + W), at lower
temperatures with excess
oil (M + O). At
intermediate temperatures,
lamellar phases may be
stable at higher water
contents and higher oil
contents, respectively
(After Olsson and
Wennerstrom 1994)

dihydroxy compounds decreased the


value, with the relative efficiency at inducing
percolation following the order ortho < meta
< para. The threshold temperature was
shown to be hydrotrope concentration
dependent.
Spin-Echo Diffusion NMR

The study of self-diffusion provides information


on the connectivity of a system and of the size of
dispersed domains. Diffusion depends on factors
such as the size and shape of the diffusing entities, the viscosity of the medium and temperature.
In a dilute system of dispersed spherical
droplets of effective radius R, the limiting infinite
dilution diffusion coefficient Do is given by the
StokesEinstein equation (Eq. 33):

Do

kT
6pR

(33)

where k is the Boltzmann constant, T the


temperature, and  the viscosity of the medium.
Because of the size dependence of Do, diffusion
coefficient measurements can give an impression
of the underlying structure and domain size in
a system:
(a) In a w/o microemulsion, the diffusion coefficient of the confined water molecules will be
much lower than that of the external oil
medium.
(b) In o/w microemulsions, the reverse of (a)
holds.
(c) In a bicontinuous system, both oil and water
are expected to relatively have high diffusion

Microemulsions

705

7 6

100

D
1  2f
Do

6
7

9
9

10

11

50

12

0
0

(34)

15

30
T/C

45

Microemulsions, Fig. 7 Pseudobinary phase diagram in


ternary microemulsion systems formed with the anionic
surfactant Aerosol-OT (AOT) in various straight-chain
alkane solvents. The water-to-surfactant molar ratio, w,
is plotted versus temperature at constant surfactant concentration and pressure. Alkane carbon numbers are indicated; ringed numbers correspond to the lower
temperature (solubilization) boundary, TL, and unringed
numbers to the upper temperature (haze) boundary, TU.
The single-phase microemulsion region is located
between TL and TU. Below TL, the system consists of
a microemulsion phase in equilibrium with excess water
(WII type), and above TU, the single microemulsion phase
separates into a surfactant-rich phase and an oil phase
(After Fletcher et al. 1987)

coefficients, owing to the interpenetrating


and system-spanning nature of the underlying
structure.
(d) In a chaotic molecular solution, in the
absence of any colloidal-scale self-assembly
structures, the diffusion coefficients of the
components will be greater than any of
the previous three cases. Obviously in this
case, D should tend to that for the neat
solvent.
The collective diffusion domains will be
affected by collisions by other components. To
a first approximation, for spherical domains,
obstruction can be accounted for by Eq. 34:

With D being the diffusion coefficient at


a given volume fraction f.
Self-diffusion can be studied using spin-echo
NMR experiments. The technique relies on two
radio frequency pulses at phase angles of 90 and
180 with respect to the main magnetic field Bo
which are separated by a delay time t. At 2t, the
system generates an echo. Two gradient field
pulses of duration d and intensity g are added to
the spin-echo sequence. The first one labels the
spin by introducing a phase shift in the magnetization. The second one reverses the phase shift.
The separation between gradient pulses is known
as the diffusion time. If no diffusion occurs, the
amplitude of the echo will not change with
respect to the signal after the first 90o pulse. If
diffusion does occur, a reduction in the echo
amplitude will be detected as a function of gradient strength. The amplitude of the echo is related
to the self-diffusion coefficient by Eq. 35:


2t
A2t A0 exp
T 2i

 

d
2 2 2
 exp g G d D  Di
(35)
3
where A(2t) and A(0) are the echo amplitudes in
the presence and absence of the field gradient. T2i
is the transverse relaxation time for the nuclei i, g
is the gyromagnetic ratio, G is the gradient
strength, d is the length of the field gradient
pulse, D is the distance between gradient pulses,
and Di is the self-diffusion coefficient.
Lindman et al. (1981) examined an array of
microemulsions including those stabilized by
SDS, sodium octanoate, sodium octylbenzene
sulfonate, AOT, potassium oleate, and the nonionic C10E4 in the presence of a number of cosurfactants and in different oil media. In these
systems, the self-diffusion of added short chain
alcohols, butanol and pentanol, and in certain
cases the water and oil, approached values of
the neat components. This was attributed to the
alcohol bearing similar solubility in both oil and
water and absence of extended well-defined

706

domains. The nonionic surfactant systems


exhibited similar behavior to the ionic surfactant
systems in the presence of short chain alcohols.
The remainder of the systems, those with longchain cosurfactants and those not requiring the
presence of alcohol for microemulsion formation, showed distinct domains in the media,
judged by a distinct difference in the magnitudes
of the determined diffusion coefficients. Geuring
and Lindman (1985) continued probing such systems by looking into water/SDS/Butanol/Toluene microemulsions by scanning salt
concentration and replacing butanol for pentanol.
Low salinity levels showed that the self-diffusion
coefficients of water and toluene (Dw and Dtoluene
respectively) differed, with Dw > > Dtoluene,
while at higher salinity levels, the reverse
occurred, Dw < < Dtoluene. At intermediate salinity, Dwater  Dtoluene indicative of a bicontinuous
phase. The behavior of the self-diffusion coefficients of each of the components of the
microemulsions with increasing salinity is summarized in Fig. 8.
A further study (Lindman et al. 1983) focused
primarily on a quaternary system of water/
sodium octylbenzene sulfonate/alcohol/toluene,
wherein the effect of alcohol chain length was
investigated further. Nonanol and decanol systems showed behavior attributed to clear segregated domains. Once again, butanol and pentanol
exhibited what was referred to as a structurebreaking effect which paralleled the behavior of
the composition extent of the isotropic phase
region observed in the phase behavior. The
expected picture of distinct domains was in contradiction with the component self-diffusion
observations. It was argued that this indicated
the presence of disrupted structures. In a note
supporting these two studies, Stilbs and Rapacki
(1983) examined the effect of alcohol chain
length from methanol through to octanol on
different combinations of surfactant and oil.
The weight ratios of the components were
kept the same for all systems produced qualitatively the same picture as keeping the molar
ratios constant. It was noted that the selfdiffusion of hydrocarbons remained relatively
fast irrespective of the alcohol chain length.

Microemulsions

Surfactant diffusion increased with increasing


alcohol chain length but was always slower in
comparison to other components. The selfdiffusion of the cosurfactants appeared to
decrease with increasing tail length. The selfdiffusion of water showed a decreasing trend
with increasing alcohol length. The coefficient
decreased by a factor of 23 with every additional
methylene group in the alcohol. Overall, the
report was in agreement with the previous experiments which showed that lower chain alcohols
give rise to less structured solutions which they
described as in effect bicontinuous.
To examine the effect of molecular identity of
cosurfactant alcohol, Stilbs et al. (1984) examined medium- and long-chain alcohols as well as
branched and aromatic analogues. The water diffusion coefficient Dw differed appreciably with
different cosurfactant. In agreement with previous reports, Dw was smallest for the decanol
system. For phenol, butanol, and ethylene glycol
monobutyl ether, Dw was highest and
approaching that of neat water. The pentanol
system showed slightly decreased water mobility
compared to butanol, indicating a more structured system than was obtained when using the
smaller chain alcohols. Interestingly, branched
isomers of pentanol exhibited behavior similar
to the smaller alcohols. It was noted that the
alcohol mobility increased with dilution for the
aromatics, which was attributed to the lowering
of viscosity (Sjoblom and Henriksson 1983).
Water mobility on the other hand decreased, and
it was argued that this was due to alcohol being
redistributed resulting in the domain interfaces
becoming depleted of cosurfactant, which apparently hinders water diffusion.
Dynamic Light Scattering

Dynamic light scattering (DLS, also known as


quasielastic light scattering, QELS, and photon
correlation spectroscopy, PCS), measures temporal fluctuations in the intensity of laser light
which is scattered by a sample at a fixed angle;
this information can then be analyzed to provide
data on the diffusion of colloidal particles, such
as found in droplet microemulsions. Hence, DLS
is an important characterization method for

Microemulsions
Microemulsions,
Fig. 8 Self-diffusion
coefficients of components
of brine/SDS/butanol/
toluene microemulsion
(Guering and Lindman
1985)

707
Toluene

D/m2s1
Water
109

Butanol
Butanol

1010

SDS

1011
3

10

SALINITY ( g / 100 cm3)

microemulsions, providing initial (lab-based)


evidence for the presence of nanometer-sized
dispersed domains or droplets. The basic principle of DLS relies on static Rayleigh scattering,
but now, the intensity is studied as a function time
(more strictly a delay time). For an ensemble of
particles, the intensity measured at a given angle
will be a product of the constructive and destructive interference from light scattered by all of the
particles. As the particles experience Brownian
motion, the measured intensity fluctuates over
time: these fluctuations contain information
about the rate at which particles are diffusing in
the system. The fluctuating intensity data is handled by a computerized correlator, which compares I(t) (the intensity signal at time t) with that
a small time t + t later (where t is usually around
5  107 s). The rate at which the signal intensity
changes is proportional to the rate of diffusion of
the microemulsion droplets. The correlation
function, g(t), describes this rate of change
(Wu 2008):
gt hI t
I t ti

(36)

For an ensemble of microemulsion droplets


undergoing Brownian motion, this function

Microemulsions, Fig. 9 Schematic exponential decay


in correlation function for large (slow-diffusing) and
small (fast-diffusing) microemulsion droplets

takes the form of an exponential decay, with


faster diffusing (smaller) droplets showing
a more rapid decay (Fig. 9):

gt A 1 B
g1 t2

(37)

In Eq. 37, A is the baseline of the correlation


function as t ! 1, B is a parameter which
depends on the detection geometry in the system
(determined by calibration with a known

708

standard), and the function g1(t) contains contributions from all exponential decays contained in
the correlation function. Hence, the decay rate is
proportional to the effective droplet diffusion
coefficient. Information on size data is then
obtained by fitting a model function to the experimental correlation function data (Koppel 1972;
Provencher 1982; Roig and Alessandrini 2006).
These fitting methods are incorporated as algorithms in the control software for modern DLS
instruments, and hence, analysis is highly
automated.
Using the StokesEinstein equation (Eq. 33),
assuming monodispersity and the absence of
interdroplet interactions, this diffusion coefficient allows calculation of the effective hydrodynamic radius of the droplets:
In the case of concentrated or interacting systems, modifications to this approach must be
made (Cazabat et al. 1980). A limitation of DLS
(as presented) is that it does not explicitly separate the effect of interparticle interactions on diffusion. However, as samples become more dilute,
interactions should decrease, and hence, a series
of measurements at increasing dilution is often
performed. By extrapolating the obtained diffusion coefficient to zero particle concentration,
a value for the infinitely dilute (noninteracting)
case, and hence a better estimate of the particle
size, can be obtained (Ricka et al. 1991).
An early example of light scattering experiments on microemulsions was by Schulman et al.
(Schulman and Friend 1949), at a stage when
uncertainty reigned over the properties of transparent and ternary wateramphiphileoil mixtures. These experiments were carried out to
support findings from X-ray diffraction experiments contained in an earlier publication
(Schulman and Riley 1948): both concluded that
dispersed droplets surrounded by a monolayer of
amphiphiles were present in both o/w and w/o
apparently monophasic, transparent systems.
The difference between micelles and
microemulsions were long contested (points
discussed in section Thermodynamics) with
conventional techniques unable to provide
a clear picture. Zulauf et al. (Zulauf and Elcke
1979) systematically investigated the effect of

Microemulsions

surfactant and water concentration and temperature on the geometry of reverse water/AOT/isooctane microemulsions. AOT micelles in

isooctane showed a size invariance (15.0 A


radius) within the concentration range of
8200 mM and a temperature range
20.095.0 C. It was noted that the length of
which lead to
AOT molecules is in fact 12.0 A
the conclusion that, despite purification, water
molecules were still present and associated with
the surfactant head group. Karl Fischer titrations
indicated that there were sufficient amounts of
water to speculate that they might act as linking
agents between surfactant head groups. This, they
proposed, was the reason for the lack of a cloud
point even in the very low temperatures.
When water was added at w < 10, the apparent
radius determined by DLS increased, continuing
with increasing [H2O]. It is interesting to note
that when the samples were examined in open
cuvettes at atmospheric pressure, the dispersed
drops decreased in size down to the original
micelle size if given sufficient time. The rate of
this process increased with increased T.
In microemulsions where w > 10, an increase
in radii was observed in the range of
20.050.0 C, above which the solutions became
turbid. These observations were fully reversible.
Dilution with isooctane produced a pronounced
decrease in radius. Again when the solutions
were examined in open conditions, a rapid
decrease in sizes were observed down to the
values measured for the w < 10 regime. When
the temperature was lowered to 18 C and below,
the drops increased in size and eventually coalesced to produce a macroscopically biphasic
system.
Clarke and Nicholson (1984) examined the
behavior of water stabilized by AOT in a range
of alkanes and concluded that the radii of structures present increased linearly with increasing
w, a behavior almost unique to AOT.
Cazabat et al. (1980) studied a matrix of water/
SDS/pentanol/cyclohexane microemulsions at
SDS/water ratios of 0.69 (A), 0.345 (B), and
0.23 (C) by varying the volume fraction f.
They noted that the volumes of alcohol to
cyclohexane followed a linear relationship

Microemulsions
Microemulsions,
Fig. 10 Diffusion
coefficients D against
volume fraction f for A

709

D107 cm2/s

2
Microemulsion
A
B
C
C

.1

which was obeyed in the smaller f regions but


not for higher, which was attributed to
a discontinuity in the phase quaternary diagram.
Figure 10 shows that diffusion D initially
decreased with increasing f as would be
expected from Eq. 33. The particles increase in
size and hence R and to  increase, thereby
decreasing the value of D. At larger f (0.5),
D showed a maximum which was attributed to
a transition between inverse to regular
microemulsions.
Light scattering has been proven to be
a particularly useful technique and a first reference in attaining the size of dispersed systems;
however, it has a finite resolution. For example,
Tavacoli et al. (2008) investigated calcium carbonate nanoparticles (CaCO3) dispersed in
dodecane and cyclohexane stabilized by industrial surfactants. The purpose of the experiments
was to determine the locus of trace water in
engine oils, which commonly contain dispersed
nanoparticles. One fraction of dispersed CaCO3
was placed in a desiccator under vacuum in the
presence of silica gel, so that any water present
was removed, and a second fraction was placed in
a desiccator under vacuum in the presence of an
open H2O container. Both fractions were kept
under these conditions for 65 days and removed

.2

.3

.4

.5

only to carry out the measurements. DLS showed


that the particle radii in both fractions remained
the same. Fourier transform infrared spectroscopy (FTIR) showed an increasing water peak
with passing time. For the purpose of contrast
for SANS measurements, aliquots of D2O were
introduced to dry dispersed CaCO3 nanoparticles.
The oil phase was then removed. The SANS profiles were fitted with a Schultz polydisperse hardsphere model. Analysis showed that introducing
water to the system actually increases the particle
size. The core of the particles was 2.79 nm with
the fitting providing a shell (D2O) thickness of
0.54 nm. Contrast matching to include the surfactant layer in the particle shell thickness showed
that the shell was larger than the surfactant
length, reiterating that a D2O layer must be present on the particle surface. This increase of shell
thickness corresponded well with an increase of
the radii ascertained from the effective volume
fractions as determined from viscosity measurements. The slight increase in viscosity
corresponded to an increased volume fraction.
The disparity between DLS and SANS was
attributed to the errors in viscosity as determined
by DLS due to the presence of some base oil in
the overbased nanoparticles. Such issues do
not arise with SANS, as it is a firsthand report

710

Microemulsions

on the geometry of the scatterer. However, DLS


is still an important technique in studies of
microemulsions.
Small-Angle Neutron Scattering

Over recent decades, small-angle neutron scattering (SANS) has become one of the most valuable
techniques for probing structure and interactions
in colloidal dispersions. Like any scattering
experiment, it relies on the interaction of a beam
of radiation with the sample of interest; neutrons
are scattered by interaction with the atomic nuclei
in the sample. By measuring the intensity of
scattered radiation as a function of scattering
angle, information can be extracted on the size
and shape of structures present in the sample
(provided they are on an appropriate length
scale for the wavelength of radiation used)
and the interactions between them. Colloidal
structures such as small nanoparticles and
microemulsion droplets fall within a size range
that is ideal for study by neutrons (wavelength,
). Therefore, SANS is a very
l 0.130 A
important method for characterization of
microemulsions.
Basic Principles The magnitude of the interaction between the incoming neutron beam and the
nucleus depends on a property of the nucleus
known as scattering length, b. For bulk materials,
it is more convenient to use a summation of the
scattering lengths for all nuclei per unit volume.
This property is known as scattering length density, r, and is calculated as:
r

P
DN A X
i bi
b

i i
MR
Vm

(38)

In Eq. 38, D is the bulk density, MR is the


molar mass, NA is Avogadros number, and Vm
is the molecular volume. The scattering lengths
and scattering length densities of some commonly encountered nuclei and compounds are
shown in Table 1.
The substantial difference in scattering length
between hydrogen (1H) and deuterium (2H, D), is
a useful and much exploited property. By selectively deuterating parts of a microemulsion,

Microemulsions, Table 1 Scattering lengths of selected


nuclei and scattering length densities of some common
compounds




Compound r 1010 cm2
Nucleus b 1012 cm
1
H
0.374
H2O
0.560
2
2
H(D)
0.667
H2O(D2O)
6.356
12
C
0.665
Toluene
0.941
16
O
0.580
D-toluene
5.662
14
N
0.936
TX-100
0.519
32
S
0.285
AOT
0.542

contrast can be generated in order to show scattering from certain structures (e.g., the core of
a microemulsion droplet) by enhancing the difference in scattering length densities between
components. This is shown schematically in
Fig. 11.
The downside to this powerful tool is that
deuteration can be extremely expensive, it may
not be feasible to deuterate certain compounds,
and deuteration sometimes produces slight
changes in the phase behavior of a system.
An important experimental quantity is related
to both the scattering angle, y, and the incident
neutron wavelength, l:
Q

4p
y
sin
l
2

(39)

The scattered intensity, Is measured at the


detector, is simply the squared modulus of the
amplitude:


2


I s Q As Q

(40)

For a system of n particles, the scattering averaged over all orientations, o, and shapes, s, this
becomes:


2


Is Q
As Q
o

(41)
s

Data Analysis Many systems studied by SANS


(such as micelles and microemulsion droplets)

Microemulsions

711

Microemulsions,
Fig. 11 Schematic of the
different possible neutron
scattering contrasts for
a microemulsion droplet

1 2 3

core contrast: 1 2 = 3

comprise spherical particles dispersed in the solvent medium. A physical description of the pattern of scattered radiation from these structures is
therefore useful. Scattering from monodisperse
spheres can be described by:

2
IQ fp
rp  rs
V p
PQ; R

SQ Binc

(42)

In Eq. 42, fp is the sphere volume fraction


(calculated as the number density of particles, Np
multiplied by the volume of one particle Vp), rp
and rs are the scattering length densities of the
particle and solvent, P(Q,R) and S(Q) are the form
factor and structure factor, and Binc is the incoherent background. The proportionality factor which
relates the intensity to the form and structure
factors is known as the scale factor, SF:

2
SF fp
rp  rs
V p

(43)

This parameter is useful when modeling


SANS data, as it can separately calculated and
fitted, giving an estimate of the fit accuracy.
The form factor, P(Q), describes scattering
which arises from the nuclei within structures,
and hence, the interference patterns give information on the size and shape of structures. The
form factor for a homogeneous sphere is

PQ; R

3sinQR  QR
cosQR
QR3

2

1 2 3

shell contrast: 1 = 3 2

Systems such as microemulsion droplets stabilized by a surfactant layer may be seen as


a central core (the dispersed solvent component

1 2 3

drop contrast: 1 = 2 3

of a microemulsion) surrounded by a shell of


surfactant (Fig. 11). In this case, a more complex
form factor is needed to describe the scattering
which arises from such structures (Markovic and
Ottewill 1986).
In real systems such as micelles and
microemulsions, there is a distribution of sizes
around an average value. This polydispersity has
a significant effect on the scattering and needs to
be accounted for in modeling. The Schultz distribution function is a commonly used polydispersity factor, described by an average particle
radius and root mean square deviation
(Kotlarchyk et al. 1984):
s

av

Z 11=2

(45)

where Rav is the average particle radius and Z is


a width parameter.
The structure factor, S(Q), describes
interaggregate/particle scattering and hence
gives information on the interactions between
structures. It is dependent on the type of interactions which the structures experience (e.g.,
excluded volume, attractive, repulsive). For
water-in-oil (w/o) microemulsion droplets far
from any phase boundary, the hard-sphere interaction potential is often appropriate, as attractive
interactions are minimal (Ashcroft and Lekner
1966; de Kruif et al. 1988):
SQ 

(44)

1
 

1  N p
f r d ; fp

(46)

When a phase boundary is approached in


microemulsions (e.g., when droplets are close to
coalescence), attractive forces start to become

712

Microemulsions

Microemulsions,
Fig. 12 SANS profiles at
increasing pressure at
constant temperature of
AOT w/o microemulsions
at w 30 (Ornstein and
Zernike 1914)

92 bar
30

I(Q) (relative units)

20

120 bar

200 bar
10

500 bar

0
0.02

0.04

0.06

0.08

Q/1

important, and the hard-sphere potential is no


longer obeyed. In this case, the OrnsteinZernike
potential is commonly used (Ornstein and
Zernike 1914; Zemb 1991):


N p kB Tw
SQ 1
1 Q 2 x2


(47)

In Equation, kB is the Boltzmann constant, T is


the temperature, w is the isothermal compressibility and x is a correlation length.
Eastoe et al. (1990) studied AOT-stabilized
w/o microemulsions in alkanes, from propane to
decane. Microemulsions in propane to pentane
are considered to be low-density systems,
while microemulsions in hexane to decane are
referred to as high-density systems. Pressure
temperature phase diagrams were determined for
a range of w values. Viscosity and density of the
alkanes is a function of temperature and pressure.
Higher pressure at constant temperature causes
an increase in viscosity and density.
An increase of temperature at constant pressure results in a decrease of viscosity and density.
SANS revealed that the effect of pressure on the
microemulsion radius could be considered negligible. This appeared to be the case for all alkanes
and w values. It would have been expected that
a change in pressure might affect the equilibrium
between surfactant adsorbed at the water/oil

interface. However, if the concentration of


adsorbed surfactant was high and that of the free
surfactant was low, then changes in density and
consequently a shift of the equilibrium toward the
adsorbed surfactant would be small, as indeed
shown in the decreased scattering intensity with
increasing pressure in Fig. 12. Changing the oil
alkane length again was thought to have an influence on the droplet radius, as smaller alkanes
would be better at penetrating the adsorbed surfactant layer than the higher-chain-length
alkanes. However, once more there appeared to
be no change in the radii of the corresponding
w values across the alkane series. When the
microemulsions were forced close to the phase
boundary, by pressure variation, an effect on the
scattering profiles was observed. An increase in I
(Q) was observed with decreasing pressure. At
w 30 at 500 bar analysis showed a P(Q) consistent to a non interacting spherical model.
When pressure decreased, analysis showed
a progressively stronger S(Q), which indicates
that the droplets are experiencing long-range
attractive interactions. It was concluded that
near the phase boundary, large transient clusters
of droplets are formed.
It is readily appreciated that AOT-stabilized
microemulsions are suitable model systems for
understanding microemulsions in general and
their properties. Since microemulsion properties

Microemulsions

80

713

L2-H2O

o
60

o
o

40
1 L2-microemulsion

SO3 Na+

di-C6SS

20

10

20

30

40

50

60

70

T / C
o
3

80
60
w

o
1

o
o

SO3

Na+

o
o
o
o

o
5

o
o

40

SO3

Na+

o
10

20

30

40
50
T / C

60

SO3 Na+

o
5

o
o

o
o

20

2
SO3 Na+

SO3 Na+

o
o
o

6
SO3 Na+

70

Microemulsions, Fig. 13 Phase behavior and structural analogues of AOT (Zemb 1991)

are governed by the stabilizing film at the o/w


interface, investigating the effect of molecular
architecture variance proves fruitful. Nave et al.
(2000a) investigated eleven AOT-related compounds. This library of surfactants consisted of
linear dichain (di-C6SS) and branched dichain
analogues (AOTs) (Fig. 13).
The linear counterparts required the presence
of hexanol for the formation of microemulsions,
while the branched surfactants did not require
a cosurfactant. The phase behavior of the
branched analogue microemulsions differed on
in the locus of the phase boundary along the
temperature scale. SANS was performed on the
branched analogues, as comparison to the linear
analogues would not yield a direct comparison
due to the presence of cosurfactant. SANS on
water/AOTs/heptane reverse microemulsion systems were carried out at temperatures that were of
equivalent distance to the phase boundaries.
Analysis of the topologies of the microemulsion

was attained by carrying out core-shell-drop


contrast series experiments (CSD, Fig. 14).
The experiments revealed that the two most
highly branched analogues had smaller solubilization capacities with a decrease in droplet radii
relative to the rest. It was also noted that
of 36 A
the two branched analogues with the longest tails
also showed a thicker surfactant layer which is
testament toward the high-resolution capabilities
of the technique. The obtained radii scaled linearly with w value, a behavior shown to occur for
regular AOT microemulsions in previous publications (Kotlarchyk et al. 1982).
Following up on this investigation, Nave et al.
(2000b) examined AOT analogues with linear
and branched phenyl-tipped tails and compared
these to AOT itself (Fig. 15). Water in oil
microemulsion phase stability diagrams were
carried out in n-heptane and toluene.
While AOT formed stable single-phase
microemulsions in both media, the phenyl-tipped

714

Microemulsions,
Fig. 14 Contrast-matched
scattering data and fitting
lines for AOT analogue
(Zemb 1991)

Microemulsions
I(Q) / cm1
100

AOT 4

50

10
5
drop
1

core

0.5
shell
0.1
0.05
0.004

0.02

0.008

0.04

0.07

0.3

Q / 1

O
O
O

SO3 Na+
O

Aerosol-OT
O

SO3

di-PhC4SS

di-PhC5SS

Microemulsions,
Fig. 15 Regular and
phenyl-tipped AOT
analogues (Kotlarchyk
et al. 1982)

SO3 Na+

Na

O
SO3 Na+

O
O
br-di-PhC3SS

analogues only formed Winsor II systems in heptane. Single-phase systems were attained in toluene though at significantly lower w values than
typical AOT microemulsions in aliphatic media.
Interestingly, the phenyl-tipped analogues only

SO3 Na+

O
O
br-di-PhC5SS

showed Winsor II system formation at higher


w values. This boundary increased slightly with
temperature. This boundary was found at higher
w values for the shorter of the two linear phenyltipped analogues, while the reverse was observed

Microemulsions

50

10
5

I(Q) / cm1

Microemulsions,
Fig. 16 Scattering profiles
of di-PhC4SS (circles),
di-PhC5SS (crosses), and
br-PhC5SS (triangles) at
w 16 (intensities have
been multiplied by 3.5 for
di-PhC5SS and 20 for
br-PhC5SS for clarity)

715

1
0.5

0.1
0.05
0.01

0.02

0.03

0.05

0.08

0.28

Q / 1

for the branched ones. The boundary for AOT


itself lay between the two aforementioned classes. SANS measurements were carried out and
analyzed in the same manner as the previous
AOT analogue study. As was the case with the
previous set of experiments, at constant w values,
the droplet radius was shown to decrease slightly
with increasing alkyl tail chain length, while the
polydispersity was more or less constant
(Fig. 16).
This similarity in polydispersities follows the
argument of the film bending rigidity model, in
which the surfactant layer thickness scales
inversely with polydispersity. As the chain
lengths are not different in a pronounced way,
with chain length differences of 23 carbon
atoms in the alkyl tail, the approximately constant
polydispersities observed can be seen to be in line
with this argument. Reiterating the sensitivity of
this technique, surfactant layer thicknesses did
scale according to the increase in chain length.
Once more, the core radii were also shown to
scale linearly with increasing w values. Although
the w values were small, head group areas were
found to correlate well with the extent of
branching in the molecules. The two branched
molecules (AOT and the longest chain phenyltipped AOT analogue) were found to pack less
efficiently than the linear examples, producing
higher head group areas. This reduced packing
efficiency can account for the lower water

solubilization in toluene. The head group areas


of all compounds in microemulsions appeared
similar to those estimated by surface tension
measurements at the airwater interface,
suggesting little difference between the behavior
of the surfactants at the two interfaces.
Another example of the sensitivity and hence
suitability of SANS as a technique toward the
study of such systems is highlighted in the work
of Tabor et al. (2010). Here when water was
added to AOT-stabilized silica nanoparticles dispersed in toluene, the formation of reverse
microemulsion droplets occurred. An AOTstabilized silica sol was able to accommodate up
to 2.5 wt% water, remaining as an isotropic and
clear, homogeneous phase. Any further addition
resulted in a cloudy appearance that eventually
resolved into a sol solution coexisting with an
excess water phase. Tavacoli et al. (2008) previously showed that water uptake was indeed possible however at much lower amounts than that
observed in this study. DLS measurements
showed that addition of water yielded an increase
in the apparent effective diameter. While water
can be expected to form a film around the
nanoparticles as was the case for the study
(Tavacoli et al. 2008), this is not enough to produce the physical increase seen on this occasion.
Carrying out careful contrast matching variations
the scattering profiles exhibited two form factors
(P(Q)), one from the surfactant around the

716

particles at low Q and another at high Q owing to


the presence of micelles/microemulsions. The
silica sol particles did not appear to increase in
size with addition of water; however, the background microemulsions observed were of similar
size to microemulsions containing the same
amount of water, in the absence of the dispersed
nanoparticles. In other words, the microemulsion
droplets in the wet organosol are the same as in
a background silica-free microemulsion.
Hence, SANS indeed has proven itself over
the years as an essential and highly versatile
technique in the study of micellar and
microemulsion systems and of soft matter as
a whole. Hydrogendeuterium contrast variation
represents the main advantage of SANS, permitting a unique insight into internal structure of the
dispersed domains in microemulsions.

Applications
Microemulsions as Media for Synthesis and
Processing of Nanoparticles
Synthesis
The topic of nanoparticle synthesis has seen
a rapid rise in publications since 1980. Research
into inorganic nanoparticles has soared owing to
the special properties of nanomaterials compared
to bulk metals, especially with respect to their
photochemical and semiconductor properties.
Noble metal nanoparticles have attracted significant interest due to their application in electronics
(McConnell et al. 2000), catalysis (Xia et al.
2003), and in potential medical applications
(El-Sayed et al. 2005), as well as for their bactericidal properties. The properties that are so desirable in metal nanoparticles arise from their
particle size and size distribution, since nanoscale
sizes show remarkable optical properties as well
as lower melting points, making the engineering
of devices more feasible. Silver nanoparticles are
especially interesting due to both their optical
properties, their use in the fabrication of fineline electronic circuits.
Reverse microemulsions lend themselves as
suitable nanoreactors for the synthesis of
nanoparticles, as the water pool is of the same

Microemulsions

dimensions as the desired nanoparticle size.


Inorganic salts, the starting materials for
nanoparticles, can be dispersed in the
microemulsion medium by dissolution inside
water pools. Microemulsions are dynamic systems, and as described in the section discussing
percolation, the droplets collide as a result of
Brownian motion, hence facilitating exchange
of materials allowing the reactions to take place.
This mechanism is of fundamental importance in
the synthesis of nanoparticles, and there are two
ways in which it can be employed. In the first, two
microemulsions are prepared, one containing an
inorganic salt and another containing a reducing
agent, which are then mixed (Fig. 17). The reaction has been shown to start at the interface and
proceed toward the center of the droplet (Li et al.
2005). The rate limiting step appears to be the
droplet diffusion. Control of the exchange can be
achieved by tuning the surfactant film rigidity.
The second method, and perhaps the most common, involved the preparation of one
microemulsion containing the inorganic salt.
The reducing agent is then introduced directly
into the microemulsion medium.
A commonly encountered issue is that of particle size and shape control. Through the extensive research carried out in the field, parameters
that influence the size of the nanoparticles have
been identified as (a) the type of solvent, (b) the
surfactant and/or cosurfactants, (c) added electrolyte, (d) water content, and (e) concentration of
reagents. All of these parameters have an effect
on the film rigidity of the microemulsion. This is
commonly referred to as interfacial fluidity and
is a key concept in the exchange of materials
between microemulsions.
(a) The solvent effect is thought to play a lesser
role when compared to the other parameters
(Eastoe and Sharpe 1997), with noticeable
effects only for large changes in oil chain
lengths, with large chain lengths producing
more rigid films. Furthermore, solvents that
interact more with the surfactant tails are
better able to stabilize larger particles
(Lopez-Quintela et al. 2004b).
(b) The effects of surfactant and cosurfactant
type are still under considerable debate.

Microemulsions
Microemulsions,
Fig. 17 Proposed
mechanism for
nanoparticle preparation in
microemulsions
(Capek 2004)

717
Microemulsion I

Microemulsion II

Aqueous Phase
Metal Salt
(FeCl3, FeCl2,
CuCl2, etc.

Oil Phase

Aqueous Phase
Reducing Agent
(NH4OH, N2H4,
NaBH4, etc.)

Mix Microemulsions I and II

Collision and
Coalescence of
Droplets

Oil Phase

Percolation

Precipitate
Chemical Reaction Occurs
(Metal or Metal Oxide)

Despite this, some feint trends can be distinguished. Longer surfactant hydrophobic
chains make for more rigid films than shorter
chains (Szleifer et al. 1990; Gradzielski et al.
1996; Eastoe et al. 1997c; Eastoe and Sharpe
1997). Smaller surfactant head group molecular areas generally give rise to smaller particle
sizes. However, the opposite effect was noted
by Lee et al. (2005) for certain nonionic surfactant-stabilized microemulsions. The use of
cosurfactants appears to yield smaller particles
due to increased film fluidity (Uskokovic and
Drofenik 2005). The size of the nanoparticles
also decreases with increasing cosurfactant
chain length (Charinpanitkul et al. 2005).
(c) Added electrolyte does not seem to affect
the size of the final particles; however,
initial growth rates appeared to be greatly
enhanced. Both observations were attributed
to the destabilizing effect of the electrolytes
on the microemulsions (Uskokovic and
Drofenik 2005).
(d) Studies on the effect of water content
have failed to provide a generalized rule.

In many cases, the final particle size showed


a dependence on the initial w0 (Pileni 1997;
Lisiecki and Pileni 1995; Pileni et al.
1992); however, in many other cases, such
control was not observed. Irrespective of the
w0, when allowed to go to completion, the
reactions produced similarly sized particles.
Instead, it was argued that the rate of reaction
was affected by w0. The nanoparticle
growth rate was lower at lower w0 as
a result of high film rigidity with most of
the water molecules being bound to the surfactant head group. Increasing w0 allows
for greater exchange rate between
microemulsion droplets. However, after
a certain w0, the rate decreases or reaches
a plateau, attributed to a concentration effect.
At this point, the reagents are too dilute for
optimum reaction rates.
(e) Reagent concentration has been found to
have a significant role in particle size with
size increasing with increasing reagent
concentration (Lisiecki and Pileni 2003;
Eastoe et al. 1996c; Maillard et al. 2003).

718

Further still, polydispersity appeared reduced


when the concentration of one of the reagents
was increased (Eastoe and Cox 1995).
Despite extensive research into the effect of
these parameters, these observations serve more
as possible guidelines than governing laws
toward attaining specific particle geometries.
Another technique in producing nanoparticles
is the use of radiation and, in particular, light.
Luisa Marin et al. (2008) employed UV to
photoinduce free radicals from photosensitive
organic molecules (ketones) for the preparation
of Au nanoparticles in regular micelles. There
are now other examples (Oliveira et al. 2011),
where the surfactant also acts as a reducing
agent, being itself photoreactive. The generation
of nanoparticles in microemulsions using such
photoreactive surfactants holds promise for
future developments (Oliveira et al. 2011).
There are only a few publications employing
light to produce nanoparticles in microemulsions.
An early example was by Kurihara et al. (1983)
who used pulse radiolysis to generate Au
nanoparticles in water/pentaethylene glycol
dodecyl ether (PEGED)/n-hexane reverse
microemulsions. It was concluded that particle
growth was primarily governed by exchange of
the irradiated content between microemulsion
droplets.
Recovery

In the previous section, it was demonstrated that


microemulsions can serve as nanoreactors for
the production of nanoparticles. In making
nanoparticles, a crucial step is the nanoparticle
recovery and separation from by-products. Typical techniques are ultracentrifugation (Germain
et al. 2005), solvent evaporation (Steingerwald
et al. 1988), and the addition of appropriate
antisolvent to induce precipitation or phase
separation (Chen and Wu 2000). These are
high-energy, costly, and environmentally
unfriendly processes.
In recent times, physicochemical aspects of
microemulsions have been employed for the
recovery of nanoparticles. Abecassis et al.
(2009) used reverse microemulsions based on
the catanionic surfactant octylammonium

Microemulsions
100
t = 0.136 s
t = 15.476 s

Intensity (mm1)

101

102

103
0.1

1
q (nm1)

10

Microemulsions,
Fig.
18 SAXS
profile
of
a microemulsion during nucleation and growth of gold
nanoparticles at t 0.136 s (red) and t 15.476 s (blue)

octanoate in octane. Gold nanoparticles were


generated in this system by reducing gold hydrochloride with sodium borohydride. The reaction
and recovery were followed by small-angle X-ray
scattering (SAXS) using a stopped-flow device.
The SAXS profiles (Fig. 18) showed that the
microemulsion only changed slightly upon gold
nanoparticle formation, as evidenced by the
slight increase in intensity in the medium
Q-range. The nanoparticles generated were
spherical and the reaction did not appear to affect
the microemulsion geometry at any stage. Upon
completion of the reaction, the system was cooled
to 18 C from room temperature, causing a Winsor II phase separation (Fig. 19).
This gave preferential partitioning of the
nanoparticles into the top oil phase. Hollamby
et al. (2010) prepared Au nanoparticles by mixing
two microemulsion solutions consisting of water/
CTAB/butanol/octane, one of which contained
KAuCl4 and the other sodium borohydride,
respectively. After the formation of the Au
nanoparticles, further water was added forcing
a Winsor II transition. Employing NMR, compositions were determined to within a 1 wt% error.
The upper oil phase consisted of 96 % octane and
4 % butanol. Surfactant and water quantities were
too low to be detected. The lower phase had
a composition of 61 % water, 20 % octane, 9 %
butanol, and 10 % CTAB. The Au nanoparticles
again showed a strong preferential partitioning

Microemulsions

719

T = 25C

T = 15C

Microemulsions, Fig. 19 Appearance of a gold nanoparticle containing microemulsion Winsor IV system


(25 C) and a temperature induced a Winsor II system
(15 C)

toward the upper oil phase (82 % of the original


system). The difference to the work of Abecassis
et al. is in the method of inducing phase separation. Rather than energy-intensive cooling of the
system, environmentally benign water was introduced. This approach was extended, successfully
exploring the production/recovery and reuse of
nanoparticles by microemulsion physicochemical properties by adding nonadsorbing polymer
(Faizan et al. 2011) and fine tuning the quality of
the solvent with respect to microemulsion stability (Myakonkaya et al. 2011).
Soil Remediation
The approach of using microemulsions in soil
remediation (i.e., the removal of environmental
pollutants from topsoil) bears similarity with
approaches intended for oil recovery, which are
discussed below. This is partly due to the behavior of soil contaminants, which in most cases tend
to be aromatic and chlorinated organic molecules. Issues that arise in the use of
microemulsions in soil remediation are that
unlike oil wells, contaminant deposits have
much lower temperatures, obviously affecting
microemulsion phase behavior but also bulk
properties such as viscosity. Surfactants and
microemulsions will also adsorb onto soil particles, and hence their substituent parts need to be

biodegradable so as to avoid exchanging one


contaminant for another. Microemulsions can be
used in mobilizing the contaminants as well as
solubilizing them. Which one of the two processes dominates is system and contaminant
dependent. For example, adsorbed solid or viscous matter cannot be mobilized and solubilization is required, while low-density contaminants
can be mobilized and solubilized. In both cases,
ultralow interfacial tensions are required such as
those exhibited in Winsor III systems, as contaminants tend to be strongly adsorbed in porous
matter. Mobilization can be problematic as it
can lead to the contaminants being displaced to
greater a depth, which is speculated to have
occurred in at least one large scale field test
(Fountain et al. 1991; Oostrom et al. 1999). The
solubilization by microemulsions is seen as system dependent, with contaminants being solubilized in some microemulsions systems but not
others, and salinity scans and use of cosurfactants
are usually required, especially for polar contaminants such as chloroform, 1,2-dichlorobenzene
and trichloroethylene (Baran et al. 1994; Shiau
et al. 1996).
Soil remediation processes often involve abrasion of soil particles. However, washing is not
effective when the contaminants are organic molecules with low water solubility and hence strong
adsorption onto inorganic matter. Solvent extractions, though effective, normally are time consuming and require large solvent/soil ratios
(Khodadoust et al. 2000). Supercritical fluids
have also been employed, such as supercritical
carbon dioxide containing 10 % methanol, which
extracted up to 90 % polycyclic aromatic hydrocarbons (PAH) from soil at 50 C (Librando et al.
2004). However, it is readily understandable that
this is an energy-intensive process. Soil washing
with surfactant solutions showed that polycyclic
aromatic hydrocarbons can be removed up to
approximately 60 % by employment of commercially common nonionic surfactants (Ahn et al.
2008), though studies have shown that pollutant
swollen micelles can also adsorb into soil (Zhou
and Zhu 2007). An early report by Clemens et al.
(1994) demonstrated that surfactants can influence
the adsorption of pyrene (a model PAH).

720

The nonionic surfactant C12E4 below its cmc was


found to have no effect on the solubility of pyrene
in water, and only a minor effect on its adsorption
on layer silicates. Sodium dodecylsulfate (SDS)
and
dodecyltrimethylammonium
bromide
(DTAB) however enhanced adsorption leading to
accumulation at the solid/liquid interface. For the
nonionic surfactant above the cmc and in the presence of isooctane, solubilization of pyrene was
much higher than plain micellar solubilization,
leading to the conclusion that microemulsions
should be investigated for soil remediation applications. Suitable microemulsions should be composed of biodegradable surfactants so that
reintroduction of the soil avoids any toxicity
risks. Clemens et al. (1998) investigated three
types of microemulsions formed by nonionic
surfactants, which would take up a model contaminant and release it through the formation of a
Winsor I system as a function of temperature.
The microemulsion exhibited slight differences in
phase behavior, which played a detrimental role
in the selection of microemulsion for this process.
Calcium bentonite and a real contaminated soil
sample were used to test the efficiency of the
microemulsions in removing pyrene. Three
microemulsions were prepared: the first two types
contained two nonionic surfactants, and it was
found that the phase behavior of these systems
played a vital role in determining the optimal
extraction. In particular, avoidance of the formation of lamellar phases was crucial. All three
microemulsions showed remarkable extraction
efficiencies, surpassing Soxhlet extractions. Typically, the recovery of pyrene, in both model
and real samples, with microemulsions reached
approximately 100 % on the first step of extraction
with temperature increase offering little assistance
to the process.
A maximum contaminant solubilization was
observed in Winsor III systems. As discussed in
section Predicting Microemulsion Type, the
middle phase of the Winsor III system comprises
equal volumes of oil and water which coexists
with excess upper oil and excess lower water
phases. In the middle phase, interfacial tension
is at a minimum, resulting in optimal solubilization of contaminants therein. Winsor III systems

Microemulsions

can be produced by both ionic and nonionic


surfactants, by tuning water salinity and temperature respectively. This enhanced solubilization
by microemulsions is often referred to as
supersolubilization, and it is seen to occur in
Winsor III systems and near the transition from
Winsor I to Winsor III.
Graciaa et al. (1993a) investigated the effect
of the often necessary cosurfactant on the solubilizing capability of microemulsions. Ethoxylated
alkylphenols were employed in the formation of
brine/hexadecane and brine/ethyl oleate systems.
Oil and brine were present in equal amounts
(wateroil ratio, WOR 1). A scan of the alkyl
tail lengths for the hexadecane system showed
that in order to maintain maximum solubilization
(i.e., Winsor III system), the optimum HLB and
the optimum solubilization parameter, SP*
(Vsurfacant/Vwater), had increasing trends with
increasing tail length. In the ethyl oleate system,
although the optimum HLB followed the same
trend as for the hexadecane system, the SP* produced a minimum. The experiments were
repeated by using mixed surfactant systems
aiming at keeping the ethylene oxide number
(EON) constant at optimum (EON*) using surfactants of different ethylene oxide lengths. The
minimum in SP* was again observed, though
this time was shifted slightly higher than the
single surfactant systems. To further investigate
the causality of the effect, a model system of
brine/octylphenolethoxylates/ethyl oleate was
chosen. They found that although the apparent
EON was kept constant for all surfactant mixtures, the SP* varied considerably with the difference in EON for each of the two surfactant mixes.
In other words, solubilization increased with
increasing bimodality of the surfactants. As the
apparent EON and tail length were kept constant
at a constant brine/ethyl oleate ratio of unity, the
solubilization trend was thought to be due to
mixing effects. Also noted was that solubilization
was higher when stronger fractionation was taking place. The smaller EON surfactants become
strongly solubilized in the oil phase. These
smaller EON surfactants were given the name
lipophilic linkers. EO scans were carried out for
surfactants in the presence of a hydrophobic

Microemulsions

linker to ascertain the behavior of the EON*. The


results showed that the nonethoxylated
octylphenol linker and those bearing 0.5 and 1
EO units forced the EON* higher, which indicated that the oil medium has increased in polarity. The more lipophilic the linker was, the
stronger the effect observed. The presence of
1 % octylphenol in the oil doubled the amount
of oil solubilized in the microemulsion, which in
turn had to be balanced by watersurfactant interactions, leading to the higher apparent EON
required for the optimum phase to exist. It was
concluded that the linkers provided ordering and
enhanced oiloil and oilsurfactant interactions.
Though the interactions are extended, the ordering is not strong enough to allow for liquid crystal
formation. Graciaa et al. (1993b) went on to
examine the effect of lipophilic linkers in the
form of alcohols. For a WOR of unity, EON
scans were carried out in the presence of alcohol.
At the optimum formulation, the EON* should
shift to higher values to compensate for the
increased oil interactions; in other words more
hydrophilic surfactants are required for the formation of the Winsor II system. By measuring the
amount of alcohol in the upper oil and lower
water phase, the amount in the middle phase can
be determined. Only small amounts of alcohols
were found in the middle phase, partitioning
favorably and increasingly toward the oil phase
with increasing alcohol tail length. This also
suggested that the higher alcohols actually act as
cosolvents, a notion introduced in the first study.
The SP* varied most with the small alcohols (butanol to hexanol) which makes them also the most
surface active. Large increases in solubilization
were observed for medium length and long-chain
alcohols (hexanolhexadecanol) which is counter
intuitive as the HLB* was kept constant. The findings were interpreted as follows:
(a) C2C6 alcohols are more surface active and
adsorb at the water/oil interface at the
expense of surfactant. As a consequence, solubilization is low. This also explains the initial minima observed (also seen in the first
study (Graciaa et al. 1993a)).
(b) C6C10 alcohols are less surface active and
do not adsorb to the same extent as in case (a)

721

but provide additional surfactantoil interactions leading to higher required EON* values.
(c) C10C16 alcohols have negligible interfacial
presence. The EON* values showed invariance with increasing alcohol tail length. The
increase in solubilization is then attributed to
the lipophilic effect described in the previous
study. Furthermore, solubilization appeared
to be proportional to the alcohol length.
As mentioned previously, the nature of the
contaminant has a strong effect on its solubilization by microemulsions, with chlorinated contaminants posing the biggest problem. Acosta
et al. (2002) demonstrated that use of equimolar
amounts of hydrophilic and lipophilic linkers in
the form of sodium dimethylnapthalenesulfonate
and dodecanol in sodium dihexylsulfosuccinate
microemulsions produced the best results for
polar contaminants. This reinforces the concept
that strongly binary systems exhibit higher solubilization capabilities and demonstrating that the
linkers need not belong to the same family of
amphiphiles.
Oil Recovery
Oil fields contain, among many other substances,
a diverse mixture of hydrocarbons. These natural
resources have formed over many years from the
decomposition of biological organic matter
trapped inside the earth, under high temperatures
and pressures. Petroleum deposits (oil fields) can
span large areas and are usually within porous
rock enclaves which are surrounded by impermeable rock, thus confining oil within that area.
Initially, in certain deposits, oil may leak out
from the rock under its own pressure. In early
reports on oil, local populations especially
around the Caspian Sea would describe spontaneously occurring oil fountains which would
ignite, lasting even months (LeVine 2007). This
stage of oil recovery can release up to 20 % of the
reserve. When oil flow ceases, the next stage
involves well flooding with surfactantpolymer
solutions, injected under high pressure, mobilizing any easily accessible and displaceable oil in
the pores. This procedure can yield a further 20 %
of the total oil available in the well. The role of
the polymer is to increase the viscosity of the

722

medium, and importantly the polymer displaces


the surfactant from the bulk to the interface
and reduces its cmc. Prior to injecting the
surfactantpolymer solution, however, another
solution precedes it with the aim of removing
salts that would affect the surfactant. Such volumes of injected fluids irrespective of their nature
are referred to as slugs.
Thus the majority of the oil remains underground, adsorbed in the porous rock and too viscous to be mobilized by water. The oil exists held
in the pores in the form of ganglia trapped by
capillary action. The ganglia, which can be
thought of as large oil drops may not be mobilized due to bypassing of the water through least
resistance capillary paths or by an action known
as snap-off, where the ganglia become disconnected from the pore walls and unable to pass
through the capillary themselves, allow the less
viscous water to pass around it. This occurs when
the pore in which the oil is trapped is larger than
the pore throat. Hence, it can be seen that there
are two opposing forces, those of viscous forces
of the injected water and the capillary forces
trapping the ganglia. The dimensionless capillary
number (Nc) is used to describe the physical
properties of the particular site:
.
N c mv go=w
(48)
where m is the viscosity and n the velocity of the
injected fluid and go/w is the oilwater interfacial
tension. A high Nc is necessary to obtain a higher
extraction yield. Hence, research has focused on
increasing the viscosity of the fluid injected and
decreasing the interfacial tension. At this point,
efforts toward recovering the remaining oil are
referred to as enhanced oil recovery (EOR).
As mentioned in the previous section, ultralow
interfacial tensions are exhibited in Winsor III
systems which are obtained by increasing the
surfactant hydrophobicity, by raising the temperature for nonionic surfactants, or increasing salinity for ionic surfactants. Also at this point, equal
volumes of water and oil are solubilized in the
middle region microemulsion. In oil extraction,
the components (surfactants and brine) are usually injected into the well in which they can

Microemulsions

form microemulsions. Ideally, the components


should form Winsor III systems once in the
well. However, there are issues with the use
of microemulsions since in many cases a
cosurfactant is necessary which may separate
the solution as the mixture proceeds down
through the porous rock. As a result, research
efforts have been made toward attaining the
desirable behavior by use of a single surfactant.
Graciaa et al. showed that introducing a polar part
in the long alkyl tails of a sulfonate surfactant
allowed it to form Winsor III systems in both
polar and nonpolar oils, also exhibiting remarkable solubilization capability (Minana-Perez
et al. 1995).
Furthermore, the surfactants need to be soluble in the presence of brine and other salt ions and
to not adsorb strongly to the mineral surfaces.
Hence, sulfonate surfactants are often employed
for this purpose.
Doe et al. (1977a) varied the chain length, and
isomers were varied for alkylbenzenesulfonates
(ABS) to examine their effect on the ability of
these compounds to effect ultralow interfacial
tensions. The ABS studied are depicted below
(Fig. 20) and can be thought of as a compound
containing m + n carbon atoms with the benzene
ring attached to (m + 1)th carbon with m < n. The
compound can be visualized as a normal Cn ABS
with a Cm branch on the first carbon of the Cn
tail. Their results could be helpful in designing
surfactants depending on the desired application.
Measuring the interfacial tension of standard
surfactant solutions against a series of alkanes
provides important information on the minimum
interfacial tension (gmin) and the alkane carbon
number for minimum tension (nmin). Knowledge
of these variables can be of importance, especially in well flooding for the extraction of
petroleum.
The same authors investigated the linear
alkylbenzenesulfonates with additional alkyl
substituents for producing low interfacial tension
(Doe et al. 1978). The surfactants employed were
linear ABS having 8 to 16 carbon atom linear
alkyl chains. These molecules had one or two
alkyl substituents that could have methyl, ethyl,
propyl, i-propyl, butyl, or t-butyl groups on the

Microemulsions

723

H
H2m+1Cm

H
Cn1H2n1

SO3Na

H15C7

C8H17

SO3Na

Microemulsions, Fig. 20 Alkylbenzene sulfonates


(Doe et al. 1978)

benzene ring. The dialkylbenzenesulfonates were


grouped into two categories: those with a C12 or
C16 main chain length and those having C8 and
C10. For the first category, the general trend
followed was that the nmin increased with increasing surfactant molecular weight. For the second
category, the shift in nmin for an n-alkyl group on
the benzene ring shifted almost linearly with the
number of carbon atoms it contained, though an
added methyl group had less of an effect than
previously reported. This they attributed to the
first carbon atom effect being masked by the
benzene ring. The branched alkyl group had
a lesser effect than the linear groups. The authors
noticed that the shift in nmin for a linear butyl
group was larger than increasing the main alkyl
tail by four carbon atoms. Trialkylbenzenesulfonates were divided in the same way as the
dialkylbenzenesulfonates. For the high molecular
weight group the general trend again indicated
a nmin increase with increasing surfactant molecular weight. In the lower molecular weight group,
they observed that the introduction of two ethyl
groups had less of an effect than a single butyl
group (as in the case of dialkylbenzenesulfonates). Thus, they concluded that
dialkylbenzenesulfonates had the better molecular weight efficiency.
Doe et al. (1977b) revisited the study of
di- and trialkylbenzenesulfonates with an emphasis on the group in the ortho position on
the benzene ring relative to the sulfonate
group and molecular weight efficiency. The molecules investigated were p-di-n-alkylbenzenesulfonates, o-di-n-alkylbenzenesulfonates, and

tri-n-alkylbenzenesulfonates.
The
p-di-nalkylbenzenesulfonates had a butyl or higher
n-alkyl group in the ortho position with
respect to the sulfonate group; thus, the main
tail was in the para position. Similarly, o-di-nalkylbenzenesulfonates had the main tail in
the ortho position and the minor in the para
position. Of these three compounds, the p-di-nalkylbenzenesulfonates with the additional
n-alkyl chain proved to be the most molecular
weight efficient. Furthermore, the molecules
with relatively large ortho substituent gave their
lowest interfacial tensions at the lowest alkane
lengths a trend that held true even for the low
molecular weight structures.
In formulating water-based flooding fluids, the
parameters taken into account primarily revolve
around the surfactants employed. High temperatures pose a danger of surfactant decomposition
as the fluid may have a well residency time of up
to a year in some cases. It affects solubility and
phase behavior as well as electrolyte dissolution,
viscosity, density, and the equivalent alkane carbon number (EACN) of the oil in the well. Hence,
the oil tested under laboratory conditions may
have different EACN than that inside the well.
Other factors include salinity which has a primary
role when ionic surfactants are used as it
increases surfactant hydrophobicity and allows
the formation of Winsor III systems. However,
the presence of salts in the well may have adverse
effects when dissolved in the flooding fluid. Divalent ions may cause the precipitation of part of the
surfactant, rendering them ineffective. To avoid
such an occurrence, ion scavengers have been
employed, as well as sacrificial surfactant flushes
and salinity gradients in the injected fluid.
In summary, the overall process of EOR by
microemulsions would involve three steps:
preflushing, which conditions the well for the second flooding with the surfactants. The third step
requires a viscous polymer solution which will act
in almost a piston-like fashion to push the surfactant fluid through the pores. When this occurs, the
surfactant front forms Winsor III systems at the
frontier of the slug. This allows desorption of
the ganglia which coalesce to form an oil bank
that can then be pushed through the pore throats.

724

Increasing the rate of delivery and viscosity of


the slugs is very costly, and significant forces are
required to mobilize the ganglia in poorly accessible pores. Injecting viscous fluids at high rates
can also cause the cracking of the rock and hence
offer preferred paths of high porosity, hence lowering the efficiency of any additional input fluids.
As an alternative to oil/water-based systems,
supercritical carbon dioxide (scCO2) has been
evaluated as a displacing fluid, as it can reach
otherwise inaccessible pores. scCO2 is also miscible with an array of lower hydrocarbons, which
circumvents the issues found in water-based
flooding. As such, scCO2 has very little solubility
in water, and there is practically no loss of CO2
into brine. The disadvantage with the use of
scCO2 is that it is less viscous than the oil for
displacement, so that leakage and channeling
around the viscous oil deposits may occur with
a resulting loss of efficacy. Since the 1980s, academic attention has been directed toward increasing the viscosity of scCO2 (Cummings et al.
2011). A further issue is that the vast majority
of conventional surfactants do not dissolve in
scCO2 due to its low polarity and poor solvent
quality. Fluorinated surfactants have shown high
solubility in scCO2, but they are expensive and
environmentally persistent and damaging. This
can be avoided or partially avoided by lowering
the fluorine content in the molecules. Introducing
fluorinated groups into the surfactants is also
known to drastically decrease the cmc in water.
To examine the potential of self-assembly to provide enhanced viscosities, Eastoe et al. (Trickett
et al. 2009) employed a partially fluorinated AOT
analogue and examined the effect of three different surfactant counterions (Na+, Ni2+, Co2+).
Phase behavior significantly changed when the
divalent ions were used. The pressure instability
point (above which the system is stable, relating
to scCO2 bulk density) increased by 150 bar at
25 C. Introducing water to the micelles, thereby
increasing w, required an increase in temperature
to yield a stable dispersion. High-pressure SANS
measurements showed that the sodium analogue
formed near spherical microemulsions while the
divalent ions forced the formation of rigid, thin,
long, rod-shaped aggregates. High-pressure

Microemulsions

viscometry revealed an increase of up to 40 %


in scCO2 viscosity with the Ni2+ counterion surfactant. However, further research is required to
eliminate fluorine from such compounds and to
further increase their thickening capabilities. The
challenge has been to produce a fluorine-free
surfactant that is also soluble in scCO2 and
which also shows a workable solubility in
water. To address this problem, Mohamed et al.
(2010) designed a CO2-compatible surfactant,
TC14, being a fully hydrogenated tert-butyltipped trichain Aerosol-OT analogue. This
TC14 not only showed considerable solubility
in water, organic, and scCO2 media but also
formed micelles and microemulsions (Fig. 21
below).
In summary, the use of microemulsions in
EOR is still an emerging field with a steadily
increasing technology and promising results so
far that have found applications in detergency
and remediation processes.

Assessment
New and Future Developments
Microemulsions with Green and Novel Solvents
There is a need to reduce volatile organic solvents
and compounds (VOCs) employed in industrial
processes, since they pose environmental
threats. This has led to a drive to find suitable,
green, non-VOC solvents. Since they can be considered universal solvents, microemulsions
offer attractive prospects in this area. Two main
approaches have been explored: (a) supercritical
carbon dioxide (scCO2) (Eastoe et al. 2006)
and (b) room temperature ionic liquids (RTILs)
(Eastoe et al. 2005) as such replacements.
As such, scCO2 represents an excellent green
solvent due to the ease of solvent removal (by
reducing pressure to effect evaporation),
tuneability and recyclability of solvent quality
by temperature and pressure, the easily accessible
critical point (Pc 72.8 bar, Tc 31.1 C), and
the fact that it is nonflammable, nontoxic, environmentally benign, biocompatible, cheap, and
finally abundant. On the other hand, RTILs
are salts made of sterically mismatched ions.

Microemulsions

1.0

0.5

0.8

I(Q) / cm1

0.6

AOT
I(Q) / cm1I/CT

0.3
0.2
AOT
0.1

0.6

D2O

TC14

0.4

AOT4

0.0
0.05

0.4

0.10

Q /

0.15

0.20

TC14
0.2

AOT4
0.0
0.05

0.10

Q /1

0.15

12

10

0.20

d-heptane w = 5
AOT

I(Q) / cm1

Microemulsions,
Fig. 21 SANS profiles for
TC14, TC4, AOT4, and
AOT showing changes in
aggregate structure in three
different solvents:
(a) normal micelles in
water at 0.10 M surfactant.
The inset shows I(Q) scaled
by a reduced concentration
Cr (C cmc)/cmc to
account for differing levels
of micellized surfactants
owing to the widely
different cmcs and for
presentation purposes.
(b) Reversed micelles in
heptane for w [water]/
[surf] 5 and a surfactant
concentration of 0.10 M.
(c) Dry reverse micelles in
CO2 for w 0 and
[surf] 0.04 M obtained at
380 bar and 25 C
(Mohamed et al. 2010)

725

TC4
6

AOT4
4

TC14
2

0
0.05

0.10

Q /1

0.15

I(Q) / cm1

0.06

0.20

CO2 w = 0

TC14

AOT4
0.04

0.02

TC4

AOT
0.00
0.05

This structure hinders crystallization, and the


materials remain trapped in the liquid state.
RTILs are also green solvent candidates due to

0.10

Q /1

0.15

0.20

tuneability, polar solvation properties, and zero


volatility. By forming microemulsions with these
unusual liquids, their properties can be enhanced:

726

Microemulsions,
Fig. 22 SANS profiles of
TC14-stabilized dry
(w 0) and hydrated
(w 5) micelles obtained
in liquid CO2 at 360 bar,
25 C. The scattering
obtained from formulated
AOT4 w5 and dry AOT
w 0 systems is also
shown for comparison.
Smooth lines represent
model fits to a spherical
form factor scattering

model consistent with 11 A


radii ( 10 %) for both
TC14-stabilized micelles
(Hollamby et al. 2009)

Microemulsions

TC14 w5

TC14 w0
AOT4 w5
AOT w0

I(Q) 102/ cm1

2Rmic

0.05

0.10

0.15
Q

water-in-CO2 systems (Eastoe et al. 2006) serve


to extend the capability with polar solutes, and
oil-in-RTIL microemulsions provide compatibility of the highly polar RTIL with organic and
hydrophobic components. The development of
custom-made CO2-philic and RTIL-philic surfactants (which are not necessarily also very hydrophilic or hydrophobic) has greatly stimulated
research in these fields. As an example, SANS
data proving micelles of a custom-made CO2philic surfactant (TC14) are formed in dense
CO2, shown in Fig. 22. On the other hand, the
normal AOT-like compounds do not show any
evidence for aggregation in CO2.
Hydrofluorocarbon (HFC) solvents are recognized as attractive alternatives to fully fluorinated
solvents, owing to both low toxicity and flammability. A significant development is replacement
of CFCs by HFCs as refrigerants, as propellants
in metered-dose inhalers and drug delivery
devices for respiratory tract infections. However,
these HFC carrier solvents are hydrophobic
and, being partially fluorinated, are also
oleophobic, creating solubility problems for the
pharmaceutically active components. Using suitable HFC-compatible surfactants could allow for

0.20

0.25

/1

the dispersion of aqueous drug solutions as


microemulsions. Recently, progress has been
made (Patel et al. 2003) toward designing surfactants HFC microemulsions for potential drug
delivery applications.
With regard to drug delivery and encapsulation, the first commercialized oil-in-water
microemulsion formation is for the immunosuppressant drug cyclosporine, which is marketed as
Neoral (UK Patent No 2 222 770). The success of
Neoral points to a future role for microemulsions
in the pharma and allied medical sciences, an area
popularly termed nanomedicine.
To wrap up, microemulsion science is highly
developed: the initial intense phase of thermodynamic and phase behavior studies have paved the
way for detailed structural investigations by light,
X-ray, or neutron scattering, as well as NMR,
among other techniques. The underlying selfassembly structures are well understood and
documented, and the relationships between
chemical structure of the components (surfactants, oils) and phase stability, as well as colloidal-scale structure, are now well advanced.
The future research directions in this field will
be
toward
practical
applications
of

Microemulsions

microemulsions in diverse fields such as


nanomaterials, drug delivery, energy efficiency,
and environmental cleanup.

Cross-References
Emulsions
Micellar Systems
Phase Behavior of Surfactants
Surfactant Adsorption
Surfactants

References
Abe M, Nakamae M, Ogino K, Wade WH (1987) Chem
Lett 16(8):16131616
Abecassis B, Testard F, Zemb T (2009) Soft Matter
5:974978
Acosta E, Trans S, Uchiyama H, Sabatini DA, Harwell JF
(2002) Environ Sci Technol 36:46184624
Adamson AW (1960) Physical chemistry of surfaces.
Interscience, New York, p 393
Ahlnas T, Soderman O, Hjelm C, Lindman B (1983)
J Phys Chem 87:822828
Ahn CK, Kim YM, Woo SH, Park JM (2008) J Hazard
Mater 154:153160
Ashcroft NW, Lekner J (1966) Phys Rev 145:8389
Aveyard R, Binks BP, Clarke S, Mead J (1986a) J Chem
Soc Faraday Trans I 82:125142
Aveyard R, Binks BP, Mead J (1986b) J Chem Soc Faraday Trans I 82:17551770
Aveyard R, Binks BP, Fletcher PDI (1989) Langmuir
5:12101217
Bancroft WD (1913a) J Phys Chem 17:501519
Bancroft J (1913b) Phys Chem 17:514518
Baran JR, Popr GA, Wade WH, Weerasooriya V, Yapa
A (1994) Environ Sci Technol 28(7):13611366
Becher P (1984) J Disp Sci Technol 5:8196
Bivas I, Hanusse P, Botherel P, Lalanne J, AguerreChariol OJ (1987) Physique 48:855867
Borkovec M, Eicke H-F, Hammerich H, Gupta BD
(1988) J Phys Chem 92:206211
Bourrel M, Schechter RS (1988) Microemulsions and
related systems. Marcel Dekker, New York
Bumajdad A, Eastoe J, Heenan RK, Lu JR, Steytler DC,
Egelhaaf S (1998) J Chem Soc Faraday Trans
94:21432150
Capek I (2004) Adv Colloid Interface Sci 110:4974
Cazabat AM, Langevin D, Pouchelon A (1980) J Colloid
Interface Sci 73:112

727

Charinpanitkul T, Chanagul A, Dutta J, Rungsardthong U,


Tanthapanichakoon W (2005) Sci Technol Adv Mater
6:266271
Chen DH, Wu SH (2000) Chem Mater 12:13541360
Chen SJ, Evans FD, Ninham BW (1984) J Phys Chem
88:16311634
Clarke JHR, Nicholson JP (1984) In: Mittal K,
Lindman B (eds) Surfactants in solution III. Plenum,
New York
Clemens WD, Haegel F-H, Schwuger MJ (1994) Langmuir 10:13661369
Clemens WD, Haegel F-H, Schwuger MJ (1998) Micelles,
microemulsions and monolayers: science and technology. Marcel Dekker, New York, pp 215231
Clowes GHA (1916) J Phys Chem 20:407451
Cummings S, Trickett K, Enick R, Eastoe J (2011) Phys
Chem Chem Phys 13(4):12761289
Daillant J, Bosio L, Benattar JJ, Meunier J (1989)
Europhys Lett 8:453458
Davies JT (1957) Proceeding of the 2nd congress surface
activity. Butterworths, London
De Gennes PG, Taupin C (1982) J Phys Chem
86:22942304
de Kruif CG, Briels WJ, May RP, Vrij A (1988) Langmuir
4:668676
Di Meglio JM, Dvolaitzky M, Taupin C (1985) J Phys
Chem 89:871874
Doe PH, Wade WH, Schechter RS (1977a) J Colloid
Interface Sci 59(3):525531
Doe PH, El-Emary M, Wade WH, Schechter RS (1977b)
J Am Oil Chem Soc 55(5):513520
Doe PH, El-Emary M, Wade WH (1978) J Am Oil Chem
Soc 55:505512
Eastoe J, Cox AR (1995) Colloid Surf A Physicochem Eng
Asp 101:6376
Eastoe J, Sharpe D (1997) Langmuir 13:32893294
Eastoe J, Young WK, Robinson BH (1990) J Chem Soc
Faraday Trans 86(16):28832889
Eastoe J, Dong J, Hetherington KJ, Steytler DC, Heenen
RK (1996a) J Chem Soc Faraday Trans 92:6572
Eastoe J, Hetherington KJ, Sharpe D, Dong J, Heenan RK,
Steytler DC (1996b) Langmuir 12:38763880
Eastoe J, Stebbing S, Dalton J, Heenan RK (1996c) Colloid Surf A Physicochem Eng Asp 119:123131
Eastoe J, Hetherington KJ, Sharpe D, Dong J, Heenan RK,
Steytler DC (1997a) Colloid Surf A 128:209215
Eastoe J, Hetherington KJ, Sharpe D, Steytler DC,
Egelhaaf S, Heenan RK (1997b) Langmuir
128:24902493
Eastoe J, Sharpe D, Heenan RK, Egelhaaf S (1997c)
J Phys Chem B 101:944948
Eastoe J, Gold S, Rogers SE, Paul A, Welton T, Heenan
RK, Grillo I (2005) J Am Chem Soc 217:7302
Eastoe J, Gold S, Steytler DC (2006) Langmuir 22:9832
Eicke H-F, Meier W (1996) Biophys Chem 58:2937
El-Sayed IH, Huang X, El-Sayed MA (2005) Nano Lett
5:829834

728

Engelhardt H, Duwe HP, Sackmann E (1985) J Phys Lett


46:395400
Faizan M, Myakonkaya O, Shah SS, Eastoe J (2011)
J Colloid Interface Sci 354:624629
Farago B, Huang JS, Richter D, Safran SA, Milner ST
(1990) Prog Colloid Polym Sci 81:6063
Fletcher PDI, Howe AM, Robinson BH (1987) J Chem
Soc Faraday Trans I 83:9851006
Fletcher PDI, Clarke S, Ye X (1990) Langmuir
6:13011309
Fountain JC, Klimek A, Beikirch MG, Middleton TM
(1991) J Hazard Mater 28(3):295311
Garcia-Rio L, Leis JR, Mejuto JC, Pena ME (1994) Langmuir 10:16761683
Germain V, Brioude A, Ingert D, Pileni M-P
(2005) J Chem Phys 122(124707):18
Graciaa A, Lachaise J, Cucuphat C, Bourrel M, Salager JL
(1993a) Langmuir 9:669672
Graciaa A, Lachaise J, Cucuphat C, Bourrel M, Salager JL
(1993b) Langmuir 9:33713374
Gradzielski M, Langevin D (1996) J Mol Struct
383:145156
Gradzielski M, Langevin D, Farago B (1996) Phys Rev
E 53:39003919
Griffin WC (1949) J Cosmet Chem 1:311326
Griffin WC (1954) J Cosmet Chem 5:249256
Guering P, Lindman B (1985) Langmuir 1:464468
Hait SK, Moulik SP, Rodgers MP, Burke SE, Palepu
R (2001) J Phys Chem B 105:71457154
Hait SK, Sanyal A, Moulik SP (2002) J Phys Chem
B 106:1264212650
Hansen RS (1960) J Phys Chem 64:637641
Helfrich WZ (1973) Naturforsch 28c:693703
Hilfiker R, Eicke H-F, Hammerich H (1987) Helv Chim
Acta 70:15311536
Hoar TP, Schulman JH (1943) Nature (Lond)
152:102103
Hollamby MJ, Tricket K, Mohamed A, Cummings S,
Tabor S, Myakonkya O, Gold S, Rogers S,
Heenan RK, Eastoe J (2009) Angewandte Chem
48:49934995
Hollamby MJ, Eastoe J, Chemelli A, Glatter O,
Rogers SE, Heenan RK, Grillo I (2010) Langmuir
10:69896994
Hone DC, Robinson BH, Steytler DC (2000) Langmuir
16:340346
Hunter RJ (1994) Introduction to modern colloid science.
Oxford University Press, Oxford
Hyde S, Andersson K, Larsson K, Blum Z, Landh S,
Ninham BW (1997) The language of shape. Elsevier,
Amsterdam
Israelachvili JN (1994) Colloid Surf A 91:18
Israelachvili JN, Mitchell DJ, Ninham BW (1976) J Chem
Soc Faraday Trans 2(72):15251568
John AC, Rakshit AK (1994) Langmuir 10:20842087
Kahlweit M, Strey R, Busse G (1990) J Phys Chem
94:38813894
Kegel WK, Bodnar I, Lekkerkerker HNW (1995) J Phys
Chem 99:32723281

Microemulsions
Kellay H, Binks BP, Hendrikx Y, Lee LT, Meunier
J (1994) Adv Colloid Interface Sci 9:85112
Khodadoust AP, Bagchi R, Suidan MT, Brenner RC,
Sellers NG (2000) J Hazard Mater B80:159174
Klevens H (1953) J Am Oil Chem Soc 30:7480
Kling C (1949) Kolloid-Z 115:3744
Koppel DE (1972) J Chem Phys 57:48144820
Kotlarchyk M, Chen S-H, Huang JS (1982) J Phys Chem
86:32733276
Kotlarchyk M, Chen S-H, Huang JS, Kim MW
(1984) Phys Rev A 29:20542069
Kunieda H, Shinoda K (1980) J Colloid Interface Sci
75(2):601606
Kurihara K, Kizling J, Stenius P, Fendler J (1983) J Am
Chem Soc 105:25742579
Lang J, Zana R, Bauer R, Hoffman H, Ulbricht W (1975)
J Phys Chem 79:276283
Langevin D (ed) (1992) Light scattering by liquid surfaces
and complementary techniques. Marcel Dekker,
New York
Lee MS, Park SS, Lee G-D, Ju CS, Hong S-S (2005) Catal
Today 101:283290
LeVine S (2007) The oil and the glory. Random House,
New York
Li L, Quing-Sheng W, Ya-Ping D, Pei-Ming W (2005)
Mater Lett 59:16231626
Librando V, Hutzinger O, Tringali G, Aresta M (2004)
Chemosphere 54:11891197
Lin IJ, Friend JP, Zimmels Y (1973) J Colloid Interface
Sci 45(2):378385
Lindman B, Stilbs P, Moseley ME (1981) J Colloid Interface Sci 83(2):569582
Lindman B, Ahlnas T, Soderman O, Walderhaug H,
Rapacki K, Stilbs P (1983) Faraday Discuss Chem
Soc 76:317329
Lisiecki I, Pileni M-P (1995) J Phys Chem 99:50775082
Lisiecki I, Pileni M-P (2003) Langmuir 19:94869489
Lopez-Quintela MA, Tojo C, Blanco MC, Rio LG,
Leis JR (2004a) Curr Opin Colloid Interface Sci
9:264278
Lopez-Quintela MA, Tojo C, Blanco MC, Garcia Rio L,
Leis JR (2004b) Curr Opin Colloid Interface Sci
9:264278
Luisa Marin ML, McGilvray KL, Scaiano JC (2008) J Am
Chem Soc 130(49):1657216584
Lv FF, Li N, Zheng LQ, Tung CH (2006) Eur J Pharm Sci
62:288294
Maillard M, Giorgio S, Pileni M-P (2003) J Phys Chem
B 107:24662470
Markovic I, Ottewill RH (1986) Colloid Polym Sci
264:454462
McBain JW (1950) Colloid Science. D.C. Health, Boston
McConnell WP, Nowak JP, Brousseau LC,
Tenent RC, Feldheim DLJ (2000) Phys Chem
B 104:89258930
Meunier J, Lee LT (1991) Langmuir 46:18551860
Milner ST, Safran SA (1987a) Phys Rev A 36:
43714379
Milner ST, Safran SA (1987b) Phys Rev A 36:43714379

Miniemulsions
Minana-Perez M, Graciaa A, Lachaise J, Salager JL
(1995) Colloid Surf A 100:217224
Mohamed A, Trickett K, Chin SY, Cummings S,
Sagisaka M, Hudson L, Nave S, Dyer R, Rogers SE,
Heenan RK, Eastoe J (2010) Langmuir 26(17):
1386113866
Moulik SP, Paul BK (1998) Adv Colloid Interface Sci
78:99195
Myakonkaya O, Eastoe J, Mutch KJ, Grillo I (2011) Phys
Chem Chem Phys 13:30593063
Nave S, Eastoe J, Heenan RK, Steytler D, Grillo I (2000a)
Langmuir 16:87418748
Nave S, Paul A, Eastoe J, Pitt AR, Heenan RK (2000b)
Langmuir 21:1002110027
Oliveira RJ, Brown P, Correia G, Rogers SE, Heenan RK,
Grillo I, Galembeck A, Eastoe J (2011) Langmuir
27:92779284
Olsson U, Wennerstrom H (1994) Adv Colloid Interface
Sci 49:113146
Oostrom M, Hofstee C, Walker RC, Dane JH
(1999) J Contam Hydrol 37(12):179197
Ornstein LS, Zernike F (1914) Proceeding of the
Koninklijke Akademia van Weten-schappen te
Amsterdam, vol 17, pp 793806
Overbeek JTG (1978) Faraday Discuss Chem Soc
65:719
Patel N, Marlow M, Lawrence MJ (2003) J Colloid Interface Sci 258:345
Pileni M-P (1997) Langmuir 13:32663276
Pileni M-P, Motte L, Petit C (1992) Chem Mater
4:338345
Provencher SW (1982) Comp Phys Commun 27:229242
Ray S, Bisal SR, Moulik SP (1993) J Chem Soc Faraday
Trans 89(17):32773282
Ricka J, Borkovec M, Hofmeier U (1991) J Chem Phys
94(12):85038509
Rock PA (1969) Chemical thermodynamics. Macmillan,
London
Roig AR, Alessandrini JL (2006) Particle and particle
characterization 23:431437
Safran SA (1983) J Chem Phys 78:20732076
Safran SA (1992) In: Chen SH, Huang JS, Tartaglia P
(eds) Structure and dynamics of strongly interacting
colloids and supramolecular aggregates in solution,
vol 369, NATO Advanced Study Institute, Series C:
Mathematical and physical sciences. Kluwer, Dortrecht
Safran SA, Tlusty T (1996) Ber Bunsenges Phys Chem
100:252263
Schneider MB, Jenkins JT, Webb WW (1984) Biophys
J 45:891899
Schulman JH, Friend JA (1949) J Colloid Sci
4(5):497509
Schulman JH, Riley DP (1948) J Colloid Sci 3:383405
Shiau B-J, Sabatini DA, Harwell JH, Vu DQ (1996) Environ Sci Technol 30(1):97103
Shinoda K, Friberg S (1986) Emulsions and solubilization.
Wiley-Interscience, New York
Shinoda K, Saito H (1969) J Colloid Interface Sci
34:258263

729

Sicoli F, Langevin D, Lee LT (1993) J Chem Phys


99:47594765
Sjoblom E, Henriksson U (1983) In: Mittal KL,
Lindman B (eds) Surfactants in solution. Plenum
Press, New York, p 1867
Sottman T, Strey R (1996) Ber Bunsenges Phys Chem
100:237241
Steingerwald ML, Alivisatos AP, Gibson JM, Harris R,
Kortan R, Muller AJ, Thayer AM, Duncan DC, Douglas
DC, Brus LE (1988) J Am Chem Soc 110:30463050
Stilbs P, Rapacki K (1983) J Colloid Interface Sci
95(2):583585
Stilbs P, Warnheim T, Sjoblom E, Henriksson U (1984)
J Phys Chem 88:54205425
Szleifer I, Kramer D, Ben-Shaul A, Gelbart WM, Safran
S (1990) J Chem Phys 92:68006817
Tabor RF, Eastoe J, Dowding PJ, Grillo I, Rogers SE
(2010) J Colloid Interface Sci 344:447450
Tanford C (1978) The hydrophobic effect. Wiley,
New York
Tavacoli JW, Dowding PJ, Steytler DC, Barnes DJ, Routh
AF (2008) Langmuir 24:38073813
Tondre C (2005) Dynamics in microemulsion systems. In:
Zana R (ed) Dynamics of surfactant assemblies,
vol 125. Francis and Taylor, Boca Raton
Tondre CJ, Derouiche A (1991) J Disp Sci Technol
12:517530
Trickett K, Xing D, Enick R, Eastoe J, Hollamby MJ,
Mutch KJ, Rogers SE, Heenan RK, Steytler DC
(2009) Langmuir 26(1):8388
UK Patent No. 2 222 770 Pharmaceutical compositions
containing cyclosporine
Uskokovic V, Drofenik M (2005) Surf Rev Lett
12:239277
Vincent B, Kiraly Z, Obey TM (1998) In: Binks BP (ed)
Modern aspects of emulsion science. Royal Society of
Chemistry, London, p 100
Wei L, Sun P, Nie S, Pan W (2005) Drug Dev Ind Pharm
31:785794
Winsor PA (1948) Trans Faraday Soc 44:376398
Winterhalter M, Helfrich W (1992) J Phys Chem
96:327330
Wu C (2008) In: Seidel A (ed) Characterisation and
analysis of polymers. Wiley, New York
Xia Y, Yang P, Sun Y, Wu Y, Mayers B, Gates B, Yin Y,
Kim F, Yan H (2003) Adv Mater 15:353389
Zemb T (1991) In: Lidner P, Zemb T (eds) Neutron, X-ray
and light scattering: introduction to an investigative
tool for colloidal and polymeric scientists. North
Holland, Amsterdam, pp 177198
Zhou W, Zhu L (2007) Environ Pollut 147:6673
Zulauf M, Elcke H-F (1979) J Phys Chem 83(4):480486

Miniemulsions
Nanoemulsions

730

Mixed Film Theory


Tharwat Tadros
Wokingham, Berkshire, UK

Synonyms
Mixed film theory for microemulsion formation

Definition
The film (which may consist of surfactant and
cosurfactant molecules) is considered as a liquid
two dimensional third phase in equilibrium
with both oil and water. Such a monolayer
could be a duplex film, i.e., giving different properties on the water side and oil side. The initial
flat duplex film has different tensions at the oil
and water sides. This is due to the different packing of the hydrophobic and hydrophilic groups
(these groups have different sizes and crosssectional areas). It is convenient to define a twodimensional surface pressure p: p go g; go is
the interfacial tension of the clean interface,
whereas g is the interfacial tension with adsorbed
surfactant. One can define two values for p at the
oil and water phases, po and pw, which for a flat
film are not equal, i.e., p0 o p0 w. As a result of the
difference in tensions, the film will bend until
po pw. If p0 o > p0 w, the area at the oil side has
to expand (resulting in reduction of p0 o) until
po pw. In this case, a W/O microemulsion is
produced. If p0 w > p0 o, the area at the water side
expands until pw po. In this case, an O/W
microemulsion is produced. According to the
duplex film theory, the interfacial tension gT is
given
by
the
following
expression:
gT gO=W p, where (go/w)a is the interfacial
tension that is reduced by the presence of the
alcohol. The value of (go/w)a is significantly
lower than go/w in the absence of the alcohol
(e.g., for hydrocarbon/water go/w is reduced
from 50 to 1520 mNm1 on the additional of
a significant amount of a medium-chain alcohol

Mixed Film Theory

like pentanol or hexanol). Contributions to p are


considered to be due to crowding of the surfactant
and cosurfactant molecules and penetration of the
oil phase into the hydrocarbon chains of the interface. If p > (go/w)a, gT becomes negative and this
leads to expansion of the interface until gT
reaches a small positive value. Since (go/w)a is
of the order of 1520 mNm1, surface pressures
of this order are required for gT to approach
a value of zero. This above duplex film theory
can explain the nature of the microemulsion: The
surface pressures at the oil and water sides of the
interface depend on the interactions of the hydrophobic and hydrophilic portions of the surfactant
molecule at both sides, respectively. If the hydrophobic groups are bulky in nature relative to the
hydrophilic groups, then for a flat film, such
hydrophobic groups tend to crowd forming
a higher surface pressure at the oil side of the
interface; this results in bending and expansion at
the oil side forming a W/O microemulsion. An
example for a surfactant with bulky hydrophobic
groups is Aerosol OT (dioctyl sulfosuccinate). If
the hydrophilic groups are bulky such as is the
case with ethoxylated surfactants containing
more than five ethylene oxide units, crowding
occurs at the water side of the interface. This
produces an O/W microemulsion.

Cross-References
Interfacial Tension
Microemulsions
Surfactants

Mixed Film Theory for


Microemulsion Formation
Mixed Film Theory

Mixing Interaction Free Energy


Osmotic Repulsion

Mouthfeel and Food Texture

Mixtures of Suspensions and


Emulsions
Suspoemulsions

Molecular Mixing
Diffusion of Particles

Monolayer Phases
Surfactant Monolayers

Mouthfeel
Tharwat Tadros
Wokingham, Berkshire, UK

Synonyms
Mouthfeel and food texture

Definition
Mouthfeel is a sensory perception of food texture.
Food products are generally designed with an
optimum consistency for application in cutting,
slicing, spreading, or mixing. During eating and
mastication, the food loses its initial consistency, at least partially. The mouthfeel of food
products may be related to the loss of this initial
consistency. During the first stage of this mastication process, the food is comminuted by the
action of the teeth into particles (few millimeters
in size). At this stage, the food is close to its initial
consistency. Thus, in the first stages of

731

mastication, the mouthfeel may be related to its


rheological characteristics. It is, therefore, possible to relate the mouthfeel during the first stages
of mastication to the rheological parameters such
as yield value, creep compliance, and storage
modulus. After the initial stages of comminution,
the food particles soften as a result of temperature rise and moisture uptake in the oral cavity.
This results in significant reduction in consistency which may reach values of stresses comparable to the level encountered by the saliva
flow in the oral cavity. When these stresses are
reached, the food particles will be broken down to
much smaller size that is determined by the
hydrodynamics of the flowing saliva. The
flow in the saliva is rather complex, and calculation of shear stresses is not straightforward. When
the above stage is reached, the food product will
form a homogeneous mix with the saliva, and
the mouthfeel will appear smooth. It is clear that
if the consistency of the product does not
decrease to a sufficient degree (such that the
stresses are comparable to those encountered by
the saliva flow), the masticated food will remain
thicker and the mouthfeel becomes unacceptable to the consumer (feel of graininess, stickiness, or waxiness). Control of the
consistency (rheological characteristics) of
food products is essential for consumer acceptability, and this may require sophisticated measurements and interpretation of the results
obtained.

Cross-References
Food Colloids
Rheology

Mouthfeel and Food Texture


Mouthfeel

S-ar putea să vă placă și