You are on page 1of 28

NANOEMULSIONS

Table of Contents

1. INTRODUCTION` 1

2. GENERALIZED FORMULATION – PHASE BEHAVIOR 2

3. EMULSIFICATION 5

3.1. Classification of Emulsification Methods 5

3.2. Emulsification by mechanical Energy (brute force) 6

3.3. Emulsification by Phase Transition (persuasion methods) 10

3.3.1 Transitional Inversion due to HLD Change at Constant Cs and WOR 11


3.3.2 Combined HLD-WOR Direct Change 14
3.3.3. Indirect or Induced HLD change produced by a WOR Change 14
3.3.4 Catastrophic Inversion 17

4. STABILITY ISSUES 18

4.1 Emulsion decay and coalescence 18

4.2. Ostwald Ripening 19

CONCLUSIONS 21

REFERENCES 21
1. INTRODUCTION
Emulsions are dispersions of (at least) two immiscible liquids that are thermodynamically
unstable because their Gibbs free energy of formation is positive (1).

ΔGform = γ ΔA - TΔS [1]

The dominant term is the surface contribution γ ΔA in which γ is the interfacial tension and
ΔA, the created surface area. The entropic term, TΔS, is insignificant unless the drop size is
extremely small. However, a small drop size also increases the surface area, and consequently
the first term, because the specific surface area per unit volume of drop phase is inversely
proportional to the drop size.
Emulsions have to be kinetically stabilized by the some effect to delay or inhibit the
coalescence of drops. Many factors are likely to influence the stability of an emulsion, among
them the physicochemical formulation, in particular the surfactant role, the composition and the
drop size distribution and average.
Emulsions with drops large enough to settle in the gravity field are generally called
macroemulsions or simply emulsions, although their droplet size could be as small as 1 µm.
This value corresponds to the limit of observation through an optical microscope, and to the
limit of gravitational settling according to Stokes’ law. Below this level, the name “colloid” is
generally used to describe a multiphasic system, which is often referred to as a certain kind of
emulsion if two liquids are involved.
First of all, it is worth remarking that the term “microemulsion” is a misnomer because it
does not correspond to a two-phase system, but to a single phase, with some eventual structure
such as a percolation or bicontinuity configuration. Moreover, a microemulsion is
thermodynamically stable and has no interface, nor drops (2-4).
Liquid-liquid two-phase dispersions with drop size below 0.5 µm have been named ultrafine
emulsions (5), unstable microemulsions (6) or submicron emulsions (7, 8). These are legitimate
labels but there are not very much in use, and the commonly accepted naming for such tiny
droplet dispersions is miniemulsion (9-12) or nanoemulsion. (13-15).
It is worth noting that nanoemulsion is nevertheless a misnomer because the typical drop size
ranges from 500 nm down to 50 nm, which is still much larger than 1 nm. However, the term
“nano” has a much stronger appeal than “mini”, and it has prevailed in the literature, both for
scientific and commercial purposes.
Nanoemulsions exhibit the typical properties of all emulsions, but with some specifics. (13,
16, 17) Their submicron drops make them transparent or at least translucent, depending on the
actual drop size and internal phase content. Since the interaction with the light is very sensitive
to the drop size, changes in visual aspect are very likely to occur as soon as there is some
evolution in the drop size, e.g., some destabilization.
Nanoemuslions are not particularly viscous at very low internal phase ratio. However, it is
well known that emulsion viscosity increases as the drop size decreases, hence, and contrarily to
macroemulsions, nanoemulsions could become viscous at 40-50% of internal phase, specially if
they are monodispersed, and in such case they are often called gel-emulsions. (18,19)
Since they associate transparency or translucency with their high viscosity, nanoemulsions
are appealing for use in pharmaceuticals and cosmetics. Because of the extremely small drop
size, there are compatible with parenteral applications, and transfer through the skin with rapid

FIRP Booklet E237A 1 Nanoemulsions


penetration and hydrating power. Consequently, they are found in cosmetics and pharmaceuticals
such as transparent milk, crystal gel, hair care rinse, body and skin care products, A and E
vitamin formulas, vaccines, desinfectants, antifungi and antiviral preparations etc. An Internet
search with the key word “nanoemulsion” returned over 1000 references with a large majority of
biological applications (20-22).
The differences between nanoemulsions and macroemulsions are not fundamental, but quite
important in practice, because their formation process and main properties are often different.
Because of their smaller drop size, nanoemulsions have a much higher surface area, and
consequently a higher amount of surfactant is required to stabilize them, particularly to account
for the interfacial coverage. If a macroemulsion can be stabilized with about 0.5-1.0 %
surfactant, five or more times is required for nanoemulsions, althought this is still quite lower
than the typical 20 % needed to cosolubilize oil and water in a microemulsion.
Gravity has essentially no effect on nanoemulsions, but Brownian motion is likely to put into
motion the droplets and provide them with a driving force for flocculation and coalescence.
Because of the small size, the Laplace pressure excess inside the drops is quite large. Hence,
the drops are difficult to shear and elongate, and accordingly to break, unless the interfacial
tension is very low. The large Laplace pressure also results in a strong osmotic driving force
from smaller to larger droplets, in the so-called Ostwald ripening, which is typical of
nanoemulsions with drop size below 100 nm.
Most surfactants are conceivably good candidates to stabilize a nanoemulsion by electrostatic
or steric repulsion mechanisms. However, not all surfactant will provide the right conditions to
form a stable nanoemulsion, particularly through low or zero energy emulsification techniques,
in which the formulation and phase behavior are paramount characteristics. With brute force
stirring methods, the main issue is the drop breaking process and many surfactants have been
used, such as ethoxylated sorbitan or sucrose alkyl esters, alcohol or phenol ethoxylates,
phosphoric esters, as well as various ionic surfactants such as alkylsulfates, alkylarylsulfonates,
or alkylammonium salts.

2. GENERALIZED FORMULATION – PHASE BEHAVIOR

Formulation is of paramount importance because the properties of all kinds of emulsions


directly depend on formulation, as discussed in details elsewhere (63). On the other hand, the
result of the emulsification process, i. e., the emulsion type and the drop size, depends on
properties such as the value of the interfacial tension and the adsorption of surfactant that are
also clearly influenced by the formulation.

The formulation concept is based on the chemical potential of substances, i.e. the driving
force to produce changes, which is usually written as:

µ = µ* + RT ln (χ/χref) [2]

where µ* is the value of the chemical potential in some reference state, and "χ" is the activity
of the substance, which, for the sake of simplicity, may be replaced by its concentration as a first
approximation. This expression indicates that the chemical potential of a substance depend on
two terms: the first one (µ*) depends on the nature of the components of the system, as well as

FIRP Booklet E237A 2 Nanoemulsions


temperature and pressure, which are all intensive formulation variables, or field variables,
independent of the size of the system. On the other hand, the second term (ln χ/χref) depends on
composition, which in the case of a simple surfactant-oil-water ternary system is represented by
any two independent concentrations, e. g., the surfactant concentration (Cs) and the water-oil
ratio (WOR).
The enhanced oil recovery research drive during the 1970s has generated a large corpus of
experimental and fundamental work, which has resulted in a generalized formulation concept,
so-called surfactant affinity difference (SAD). SAD actually represents the free energy of
transfer of a surfactant molecule from the water to the oil phase (2-3), as discussed elsewhere (4).

SAD = µ*o - µ*w = Δµ*w→o = RT ln (χ/χref) [3]

SAD may be expressed as a series of contributions of all the formulation variables involved
in the system, such as a surfactant characteristic parameter σ for ionic surfactants. In the case of
ethoxylated nonionics the surfactant parameter has been split into a contribution of the head
group (number of ethylene oxide group EON per molecule), and another for the tail group (α). If
the oil is a n-alkane, its nature is characterized by its alkane carbon number (ACN). If the
aqueous phase is a sodium chloride brine its characteristic parameter is its salinity. Finally the
temperature (T), and eventually pressure (P) have a known influence on SAD.
A dimensionless expression of SAD, so-called HLD for “hydrophilic-lipophilic deviation”
quantitatively express the departure of the formulation from the so-called "optimum" formulation
in which the surfactant-oil-water system exhibits three-phase behavior, i.e., Winsor III type, and
a minimum interfacial tension (23). The numerical expressions for ionic and ethoxylated
nonionic surfactants systems come directly from the conditions for the attainment of optimum
formulation for enhanced oil recovery (2):

ionic HLD = (SAD-SADref)/RT = lnS - k ACN + φ(A) + σ - aT (T-Tref) [4]

nonionic HLD = (SAD-SADref)/RT = α - EON - k ACN + b S + φ(A) + cT (T-Tref) [5]

where aT, cT, k, b are constants.


It is worth noting that SADref is essentially zero for most ionic systems, but it can be quite
different for poyethoxylated nonionics that exhibit a very low critical micelle concentration in
water and a very high partitioning in monomolecular form in some oil phases.
Unlike HLB, HLD involves not only the characteristics of the surfactant, but also takes into
accounts all the other compounds, as well as temperature and pressure, just as Winsor’s R ratio,
but in a quantitative way. Whenever HLD < 0 (respectively > 0), the surfactant affinity for the
aqueous (respectively oil) phase dominates and at low surfactant concentration a Winsor I
(respectively II) phase behavior is attained. At HLD = 0 a three-phase, microemulsion/excess oil
/excess water Winsor III system is exhibited.
These expressions are very useful in practice because (a) they describe the physicochemical
conditions with a single variable instead of a score of them, (b) they show that formulation could
be altered in many ways with similar effects, and (c) they quantify the compensating effects of
the different variables. From now on, the formulation will be handled only through HLD,
whathever the actually used variable, temperature, EON or salinity, which is used for changing
the formulation of the system. More details on how to estimate HLD are provided elsewhere (4).

FIRP Booklet E237A 3 Nanoemulsions


In a ternary system, two independent variables are required to describe the composition, and
are generally selected as the surfactant concentration (Cs) and the water-to-oil ratio (WOR). It
means that the chemical potential (equation [2]) depends on three variables: the generalized
formulation variable HLD, and two composition variables (Cs and WOR). In practice these are
too many variables to plot any property, and one of them has to be discarded. Consequently,
there are three kinds of diagrams, in which the phase behavior could be plotted (see Fig. 1).

(a) Cs-WOR diagram (at constant HLD), usually a triangular diagram


(b) HLD-Cs diagram (at constant WOR) so-called "fish" diagram
(c) HLD-WOR diagram (at constant Cs) sometimes called "X" diagram

Cs

Formulation HLD
(Temperature)

WOR

triangular cut fish cut X cut

S 2!
2!
HLD (Temo.)
HLD (Temo.)

3! 1! 3!
1!
2!
2!
2!

O W O WOR W
Cs
constant HLD > 0 constant WOR constant Cs

Fig 1: Phase behavior of SOW systems in the different kinds of 2D diagrams resulting from a cut
in a 3D diagram. (a) Cs-WOR diagram at constant HLD, (b) HLD-Cs diagram at constant WOR,
(c) HLD-WOR diagram at constant Cs.

Depending of the kind of variables that is handled, some diagram would turn up to be more
convenient than other. For instance, in the case of a change of temperature, any of the two last
cases, which involve HLD, hence temperature, is useful. On the contrary, if the process to be
described involves the dilution of a concentrated microemulsion by water, the triangular or X
diagrams would be preferred.
It is worth noting that the aspect of the phase behavior might be more complex than indicated
in Fig. 1, in particular with mixtures of surfactants or surfactants and co-surfactants in which
partitioning takes place. In such cases, the formulation is altered whenever Cs or WOR change,
thus resulting in slanted optimum formulation lines (24) in the center and right graph in Fig. 2.
Mesophases, particularly lamellar liquid crystals (LC) that form at optimum formulation, are also
likely to complicate the phase behavior as in Fig. 2 left and center graphs.

FIRP Booklet E237A 4 Nanoemulsions


S 3! 2!

HLD (Temo.)
2!

HLD (Temo.)
1! 3!
LC 1!
LC
2!
2! 2!
3!
O W Cs O WOR W
constant HLD constant WOR constant Cs

Fig. 2. 2D-Diagrams in presence of liquid crystal (LC) or partitioning of amphiphilic mixture of


surfactants or surfactant and co-surfactant.

3. EMULSIFICATION

3.1. CLASSIFICATION OF EMULSIFICATION METHODS

The formation of an emulsion results in an increase in surface area, i. e., an increase in Gibbs
free energy, even in the so-called spontaneous emulsification processes. As a consequence, the
required energy has to be found somewhere, and there are essentially three cases.
In the first case, the energy is provided by an external input under the form of shear or
inertial disruption. This is the most common case of emulsification by mixing or stirring devices,
in which only a very small proportion of the energy, e. g., 1%, is used to create surface area.
Since smaller droplets mean higher surface area, the energy required to form a nanoemulsion can
be quite high, with all associated practical problems such as cost and cooling issues. In the case
of a a liquid-liquid system with an interfacial tension of 10 mN/m, it may be calculated that to
deform a 100 nm spherical droplet into a cylinder four times longer than its diameter, so that it is
likely to breakup according to Rayleigh instability, the Laplace pressure to oppose is 500,000 Pa,
i. e., 10 atmospheres. The stress to apply would be probably quite higher because of the poor
efficiency of the shear transfer to the drop surface.
The second type of emulsification takes advantage of a change in emulsion morphology,
often a phase inversion, during which the emulsion undergoes through a weak stability state
when its curvature is swapping sign. In these transient circumstances, the drop-breaking
conditions could be easily satisfied by some mechanical instability or physicochemical transition
through an ultra- low tension state. In this second class, the system under evolution is generally
blended for macro-mixing purpose, but the stirring energy is low and has essentially no effect on
the drop size. Consequently, these techniques are quite attractive to make nanoemulsions.
The third class, referred to as spontaneous emulsification, has been sometimes explained by
the occurrence of a negative interfacial tension, which is however difficult to include in the usual
context of surface phenomena. If the tension is extremely low, the free energy of emulsion
formation can become negative because of the increasing entropy term since the drop number
increase concomitantly with the decreasing drop size. However, this reasoning does not match
the case of nanoemulsification because it works only when the drops are already very small, that
is, after the formation of the nanoemulsion. It seems that in essentially all the case of
spontaneous emulsification, there is a mass transfer process in which a substance, often the
surfactant or cosurfactant, is transferred from one phase to another where it has a lower chemical

FIRP Booklet E237A 5 Nanoemulsions


potential. Hence, the free energy source is the mass transfer, which might be sufficient when the
interfacial tension is ultra low. Thus, most of these processes are associated with near-optimum
formulation conditions, which also prevail in many instances of the second class.
Consequently the emulsification processes will be separated according to brute force
methods and near-zero energy techniques, which will gather the two last classes. (25)

3.2. EMULSIFICATION BY MECHANICAL ENERGY INPUT (BRUTE FORCE)

Drop breakup under shear, as in a stirred vessel with a rotating impeller, is not very efficient
because the power density is low and broadly distributed, and is consequently associated to a
long residence time. Although batch mixer with impellers have some advantages such as a very
good control of heating or cooling, they are not suitable for making fine emulsions by “brute
force”, unless an extremely low tension is attained.
The production of very fine droplets by stirring thus requires a device in which the disruption
zone is small and well defined, and in which the power density can be high. This is not the case
of the stirred tank illustrated in Fig. 3 (left) unless the turbine rotation is extremely energetic.
Colloid mills are essentially high-speed rotor-stator machines, as illustrated in Fig. 3 (center),
in which the fluid is forced to pass through a narrow gap where there is a very high shear, often
complemented by some inertial effects. In spite of a very short residence time (<< 1 s) the
throughput is not very high because of the small size of the flow cross-section and because the
heat generation has to be controlled.

ROTOR

high shear

fine emulsion
d< 2µm

energetic
elongation

Fig. 3. Stirred tank and Colloid mill

Static mixer of different kinds, particularly with orifice geometry, are producing an
elongational flow which is more efficient than a simple shear according to the famous Grace plot
of critical Capillary number versus viscosity ratio (26) illustrated in Fig. 4. The Capillary (or
Weber) number is the dimensionless ratio of input energy to capillary energy (resistance to drop
deformation).

"C !! R [6]
Ca =
!

where ηC is the viscosity of the continuous phase, !! the shear rate, R the drop radius, and γ the
interfacial tension,

FIRP Booklet E237A 6 Nanoemulsions


Ca crit

103

Critical Capillary Number


Two Regimes
2
10

1
10 Simple Shear
Elongational Flow Flow
0
10

-1
10 -1 1
10-3 10-2 10 100 10
Viscosity ratio R

Fig. 4. Critical capillary number versus viscosity ratio for static mixers

If the critical Ca value is attained, a drop would extend enough to be broken into two or more
droplets and conversely. Fig. 4 shows that the critical Ca for simple shear disruption undergoes
through a minimum when both fluids have roughly the same viscosity, and that it increases
particularly quickly if the drop phase is more viscous than the continuous phase. Hence simple
shear may be considered as quite inefficient, particularly for viscous drops. The figure also
shows that for axisymmetrical flow, i. e., elongational flow, the critical Ca is almost independent
of the viscosity ratio.
Emulsification devices generally referred to as homogenizers are static mixers with a
considerable variation in the cross section of flow in a constricted geometry. This may be an
orifice or an almost closed needle valve that produces axial elongation (Fig. 5 left) or a choke
valve that results in a radial elongation (Fig. 5 right) through which a large pressure drop, e. g.,
200-400 atmospheres, is maintained.
In these devices, the residence time is very short so that heating can be controlled, and the
drop size is directly related to the energy density, i. e., to the pressure drop ∆P. As far as the
drop size distribution is concerned, it is relatively monodispersed, particularly when the gap is
relatively extended as in radial flow so that a fracture flow takes place all over the constricted
zone (Fig. 5 right)

spring force
reduces the gap
Rotating Cylinder
adjustable narrow gap
(Monodispersed)
Sheared Emulsion
Fixed Cylinder

P> 200 atm Premixed


(polydispersed)
elongational Emulsion
flow

Syringe Pump

Fig. 5. Static devices for extensional flow emulsification

FIRP Booklet E237A 7 Nanoemulsions


It is worth noting that if the fracture flow zone is thicker than the actual gap, as in a
concentric cylinder device (Fig. 5 center), then a perfectly monodispersed emulsion can be
attained (27,28). However, the required energy increases rapidly with the drop size reduction and
the cylindrical gap device technique has not been applied yet to the formation of submicron
emulsions.
The design of new homogenizers focuses on the nozzle geometry, that sometimes includes
the impingement of opposing jets like in the so-called microfluidizer high pressure homogenizer.
The emulsification process occurs in a special commingling valve, the design of which is the
heart of the homogenizing apparatus. The fluid passes through a small gap in the homogenizing
valve. This creates conditions of high turbulence and shear, combined with compression,
acceleration, pressure drop, and impact that produce the disintegration of particles and dispersion
throughout the emulsion. After homogenization, the particles are of a uniform size, typically
from 200 to 2000 nm, depending on the operating pressure and formulation (or product type).
New methods, with better mechanical efficiency, but still with low throughput, have been
introduced in the two past decades.
Ultrasonic disruption is mainly due to the formation of vacuum cavities followed by their
implosions and crisscrossing high-pressure shock waves that tear the drops of the initial coarse
emulsion. The ultrasonic energy deposition is limited to a small volume (one cubic cm in most
cases) in front of a mechanical whistle, or a piezoelectric o magnetostrictive crystal in resonance,
so that the emulsion flow has to be directed to pass through this volume as indicated in Fig. 6 left
and center illustrations.

Membrane
Emulsification
Continuous Phase
ultrasonic waves (cross flow)

membrane

Phase to be
dispersed

Fig. 6. Ultrasonic whistle (left), vibrating crystal (center) and membrane (right) emulsification

The conversion of energy to ultrasonic waves is relatively efficient and this method is used
for making products like O/W cosmetic emulsions with a low content of low viscosity oils; drop
size down 200 nm may be produced if the conditions are appropriate. A review on ultrasonic
emulsification has been recently available. (29)
Membrane emulsification (30) is essentially similar to the injection (through a syringe
needle) of a stream of the internal phase into the continuous phase. It consists in forcing the drop
phase through the pores of a membrane as illustrated in Fig. 6 right. This method produce drops
one by one, hence with a very well defined size if all pores are equal and if the internal phase
does not wet the membrane surface. It is worth noting that the pressure gradient should exceed
the corresponding Laplace pressure that could be quite high for nanodroplets (31). The drop size
distribution is monodispersed, and the size depends on many factors such as characteristics of the

FIRP Booklet E237A 8 Nanoemulsions


membrane (pore size, wettability) and of the process (applied cross-flow velocity of continuous
phase, flow rate of dispersed phase, interfacial tension, adsorption kinetics). For instance, an
increased cross flow of continuous phase help the detachment of the drop and results in smaller
droplets, sometimes smaller than the pore size (32, 33). If the membrane is placed on a spinning
cylinder, the outward gravity and rotational shear can result in even smaller and uniform droplet
size. Microchannel emulsification is the last example of this kind of process, in which the drop in
formation follows the pattern of a micro fabricated channel array (35). This method requires a
smaller energy input, but seems to be limited to drop size larger than a few micrometers, hence it
is not satisfactory for making nanoemulsions, at least for now, but may be feasible for ultra-low
tension systems.
The surfactant has an important role in emulsification because its presence affects the two
opposite mechanisms that make the actual drop size distribution: the breaking mechanism and
the recoalescence during emulsification (35, 36)
By lowering the tension, the surfactant reduces the free energy of the created surface, and
makes it easier to elongate drops because of the reduced Laplace pressure. The Capillary number
is directly altered so that the amount of required energy decreases as tension diminishes. The
tension between water and oil decreases when the oil becomes more polar, e. g., it is around 50
mN/m with hexane or octane, 35 mN/m with benzene, and 8 mN/m with octanol. As soon as
some surfactant is added at its critical micellar concentration or above, the interfacial tension can
go down to a few mN/m or even lower than 1 mN/m, which is the usual value range for
detergency applications. Exceptionally, the tension can be down to 0.001 mN/m but only close to
optimum formulation.
Such a wide change, e. g., two orders of magnitudes, and sometimes much more, is likely to
affect the emulsification efficiency. However, it is worth remarking that the lowest tension takes
place at optimum formulation, and is associated with a very high instability of the produced
emulsion, so that it is not possible to take full advantage of the low tension (37-39). Actually
there exist a best compromise between a formulation close enough to optimum to insure a low
tension, but not too close to avoid the extremely quick coalescence (40). The recoalescence
during emulsification is due to the fact that after a drop has been broken into two or more
droplets, the daughter droplets are likely to be close enough and coalesce again, unless an
adsorbed surfactant layer provides an effective repulsion mechanism (36).
This means that the surfactant, which is in the continuous phase (according to Bancroft’s
rule), must adsorb very quickly on the newly created surface. This kinetics is influenced by
several factors among them the tendency to adsorb and the molecular weight. However, a low
molecular weight surfactant might not be bulky enough to provide a long term stability against
drop approach. This is why many formulation recipes advise to mix different surfactants to attain
a good enough protection against coalescence. The small molecules act during the breakup-
coalescence dynamic equilibrium taking place during emulsification and the high molecular
weight molecules that adsorb later, insures the long term effective repulsion a few minutes or a
few hours after emulsification.
Since the emulsification efficiency, i. e., the ratio of the interfacial free energy γ ΔA to the
input energy, is in general extremely low, says 1% or less, any improvement in the use of energy
is able to enhance the result, e. g., to decrease the attained drop size.
This is the case with high internal phase ratio emulsions, which exhibit a morphology similar
to a foam, that is with polyhedral drops. If such “concentrated” emulsions are stable, as it is
likely to happens in both A regions close to the inversion line (see chapter 3 maps), and at some

FIRP Booklet E237A 9 Nanoemulsions


distance from optimum formulation, says 3 or more SAD units, they exhibits a highly non-
newtonian behavior, e. g., typically a pseudoplastic or viscoelastic rheology (41). This is for
instance the case of mayonnaise (O/W) and cold cream (W/O). In such a case the introduction of
a new (big) drop of internal phase results in a region of low viscosity in the bulk of a viscoelastic
medium. As soon as the medium is submitted to motion, the shear concentrates on the low
viscosity drop and elongates it, as indicated in Fig. 7.

ΔX ΔX

ΔV ΔV

Fig. 7. Elongation of a drop in a viscoelastic medium

The elongation increases the shear (ΔV/Δx) because it reduces Δx, which becomes more
efficient to elongate the drop, and so on until Δx tends to zero and the elongated drop breaks into
scores of tiny droplets. It is worth noting that in this kind of process, the impeller rotation is
generally very slow because of the high viscosity of the emulsion, but it is obvious that the value
of ΔV is not an issue since it is the denominator Δx that tends to zero.
This process, which has been called high-alpha range emulsification because it takes place
when the emulsion has a high internal phase content, has been used in the past 25 years for
making products such as cosmetics, foodstuff or heavy crude oil emulsions (42-44). Even if the
final emulsion product is not required to contain a high proportion of internal phase, the
emulsification at high internal phase ratio might be advantageously applied, before a dilution
with external phase is carried out, generally with a static mixer, to adjust the final WOR.

3.3. EMULSIFICATION BY PHASE TRANSITION (PERSUASION METHODS)

In brute force emulsification methods the critical step is a transient modification of the drop
shape from spherical to a flattened or elongated high surface area structure, which spontaneously
tends to return to its original minimum area shape under the effect of the Laplace pressure
difference (see Fig. 8)
The smaller drop size of nanoemulsions makes this ∆P quite high, unless the tension is
extremely low or some event results in an almost zero curvature of the interfacial surface. These
conditions are met in two circumstances. The first one is when the formulation is close to
optimum, in which case both the tension is ultra low and the natural curvature is close to zero.
The second one is when an emulsion inversion takes place, i.e. when the curvature swaps
somehow and the emulsion morphology changes from O/W to W/O or conversely.

FIRP Booklet E237A 10 Nanoemulsions


"P = 2!/R Laplace's Law
! is the interfacial tension
R is the curvature radius
high P
low P Lower curvature
Lower "P
P0
P0
P1
P2 > P1 results fluid Higher curvature
motion that counters P2 higher "P
deformation
P0

Fig. 8. Laplace pressure effect on drop deformation by shear

Even if it is still not completely clear what happens at the very moment the emulsion inverts,
it is logical to guess that at some moment the curvature which changes sign, has to pass through
some zero value. These two circumstances are associated with the so-called dynamic inversion
processes, the first one named transitional and the second one catastrophic, whose characteristics
have been discussed thoroughly elsewhere (45,46).
The different cases discussed in the following sections deals with these two basic cases
and some combinations with other changes, but always with a near-zero or very low energy
input. More detailed classifications may be found elsewhere (25,47).

3.3.1. Transitional inversion due to a HLD change at constant Cs and WOR


This type of dynamic inversion corresponds to a change in any of the variable affecting HLD,
mainly temperature as reported in the well-known Phase Inversion Temperature (PIT)
emulsification technique introduced by Shinoda 35 years ago and still in use to produce fine and
submicron emulsions (48,49). It is worth noting that although the temperature is the preferred
variable to change the HLD and induce the emulsion inversion, such process can be performed as
well by changing any other variable, which is a compulsory choice for systems which are
insensitive to temperature (11,50).
Fig. 9 shows the two-dimensional maps in which the HLD is one of the variables. The bold
line indicates the inversion frontier, which exhibits a (horizontal) branch essentially at HLD = 0,
i. e., at optimum formulation. In Fig. 9 right, regions A, B, C with positive (+) or negative (-)
values of HLD are defined by the HLD = 0 line and by the position of the vertical branches, as
discussed in details elsewhere (46,51).

W/O
B+
HLD (Temp.)

C+
HLD (Temp.)

W/O 2!
2! A+
1! 1 1
- -
A 2! C
O/W 2 ! -
B O/W
Cs O WOR W
constant WOR constant Cs

Fig. 9. Inversion frontier in fish and X maps

FIRP Booklet E237A 11 Nanoemulsions


The presently discussed transitional inversion is associated at low surfactant concentration
with a Winsor III system, i. e., a microemulsion in equilibrium with excess water and excess oil.
At higher surfactant concentration, a Winsor IV system consisting of a single-phase bicontinuous
microemulsion or a lamellar liquid crystalline phase, is found. During the (vertical path) crossing
of the HLD = 0 line, the phase behavior switches from WI to WII or vice-versa, and the
surfactant-rich microemulsion changes continuously from aqueous to middle to oil phase, or
vice-versa, through a mass transfer process.
The decreasing temperature path indicated in Fig. 10 makes use of the fact that optimum
formulation (HLD = 0) corresponds to the maximum solubilization of oil and water in the
microemulsion middle phase (52). Any slight displacement from optimum would result in a
decrease in solubilization of oil or water, which would produce the exudation of one of the
excess phases, under the form of small droplets in the nanosize range.
In Fig. 10 case of a decrease in temperature with a polyethoxylated surfactant system, oil
nanodroplets will be formed when the decreasing temperature attains the PIT. Nevertheless, it is
worth noting that in most cases these nanodroplets would coalesce at once since the near zero
HLD region is intrinsically associated with extremely unstable emulsions (39, 53-56).

phase behavior at equilibrium

Temperature decrease (nonionic)

W/O emulsion MOW O/W miniemulsion


normally unstable!

Fig. 10. Changes in equilibrium (top) and emulsified (bottom) system containing a
polyethoxylated surfactant as temperature is decreased.

If coalescence could be prevented, somehow a fine emulsion could result from this process.
This is known to happen either because liquid crystal layers wrap around the droplets (57-60) or
the liquid crystal is the continuous phase of the emulsion, thus increasing its stability (61-62).
Alternatively the emulsion can be stabilized by quickly changing the formulation or temperature
through a thermal or formulation quench, so that a HLD >> 0 or HLD << 0 situation is attained
(63). When temperature is changed, such a quench corresponds to a shift of typically 20 °C
above or below the PIT. The PIT emulsification technique has been used in the past 20 years
with a variety of protocols, sometimes involving a back and forth crossing of the PIT, (64) which
might not be necessary since the main issue is to pass through an optimum formulation in which
a liquid crystal is deposited around the exuded droplets. Nevertheless, temperature cycles around
the PIT could influence the kinetics of development of LC structure and could be worth trying.

FIRP Booklet E237A 12 Nanoemulsions


If the phase behavior close to optimum formulation is monophasic, hence all formed droplets
come from the exudation of an excess phase, i. e., of the transformation of the microemulsion
domains into droplets, and hence all formed droplets are nanosized, and could be kept as such if
a proper conservation process takes place, either LC wrapping or HLD quench or both.
Contrarily, if the phase behavior is Winsor III at optimum formulation, two kinds of drops are
formed during the transition. The droplets that are exuded from the microemulsion as the
formulation goes away from optimum, and the droplets which are formed by the stirring of the
ultra-low tension system that prevails close to optimum formulation. The second category of
drops is in general larger because the low energy stirring is not effective to produce submicron
droplets.
As a consequence, a polydispersed and sometimes bimodal drop size distribution is often
found (11, 65), and if a proper stabilization mechanism is applied, the resulting fine emulsion
will be far from monodispersed, which is known as an inconvenient vulnerability with respect to
Ostwald ripening (17). The solution to this drawback is to increase the surfactant concentration
to attain a single-phase behavior at optimum formulation. To do so without using an excessive
concentration of surfactant, the solubilization has to be improved by using some techniques that
are discussed elsewhere (52). In most cases the dynamic inversion produced by a transitional
process is reversible, i. e., it happens exactly at the same formulation or temperature, whatever
the direction of change (45), unless some mesophases are formed which would stay out of
equilibrium for long period of time and inhibit mass transfer to take place (66). Oddly enough,
the crossing of a liquid crystal zone is nevertheless determinant to obtain the protective wrapping
around the nano-droplet. Then the issue seems to be where and how the liquid crystal deposits on
the droplet surface.
If the liquid crystal forms when the shifting formulation enters the optimum formulation
region, it might not be able to evolve quickly enough and it could inhibit the inversion or at least
delay it. On the other hand, if the liquid crystal forms at the very moment droplets are exuded
from the microemulsion, it is likely to build up a protective layer around them. Consequently, it
might be better not to cross the whole optimum formulation region, but rather to start in it and to
slowly shift away to allow droplets to be exuded at the same time a liquid crystal layer forms at
their interface; then in a second stage, it is convenient to shift the formulation quickly away to
avoid the occurrence of fast coalescence. Since nucleation and other kinetic phenomena are
involved, the most advantageous formulation programming will have to be optimized in each
case.
Several examples are found in the literature. When emulsification starts from three-phase
system (Winsor III), the emulsion droplet size distribution is wider than when emulsification
begins from the single microemulsion phase. To analyze these results, emulsions with different
proportion of intermediate microemulsion phase were obtained from pre-equilibrated system at
the PIT. Emulsification was produced through a temperature quench and the smallest droplet size
in resulting emulsion was obtained at higher microemulsion proportion in the initial stage. When
the emulsification processes started from single microemulsion phase, smaller droplets with
narrower size distribution were produced and it was inferred that a spontaneous emulsification
occurred in the microemulsion phase (67, 68). Likewise, studies related with nanoemulsion
formation for cosmetic application verified the microemulsion and/or lamellar liquid crystal
importance when using the PIT method (5, 69-71). Further studies have corroborated the
importance of the microemulsion or lamellar liquid crystals in the emulsification processes by
PIT method (11, 72-75).

FIRP Booklet E237A 13 Nanoemulsions


3.3.2. Combined HLD-WOR direct change
This second type of inversion is essentially equal to the previous one, but the WOR is
changed concomitantly with the formulation or temperature. This is by the way almost
compulsory when the formulation is changed by adding some concentrated solution of some
component or on the contrary by producing a dilution to reduce the salinity or some
concentration.
In this case the change in WOR is assumed not to change the interfacial formulation, and
thus does not apply when the WOR is drastically altered.
A recent study (76) showed that the exact location of the HLD-WOR path is critical, and that
the smallest drops are attained when the crossing is at HLD slightly different from zero, not at
HLD = 0, nor far away from it. This slightly off optimum formulation indicated as path 3 in
figure 11 seems to correspond to the best compromise between a low tension which facilitates
the drop breakup and a not-so-fast coalescence (77, 78).
In the reported study only micron size emulsions are attained with a WIII region at optimum
formulation. Nevertheless nanoemulsions are likely to be produced if the small droplets are
prevented from growing by some thermal quenching or liquid crystal wrapping.

W/O
2 1
HLD (Temp.)

4
O/W

O WOR W

constant Cs

Fig. 11. Combined HLD-WOR dynamic inversion paths

3.3.3. Indirect or induced HLD change produced by a WOR change.


In most cases, when the WOR changes, a partitioning of the different amphiphilic species
takes place, that is the HLD changes because of an induced HLD-WOR coupling (24, 79, 80).
For instance, if the proportion of water phase increases, the surfactant species likely to partition
onto the oil phase because they are the most hydrophobic, have less “oil space” to go and
consequently are more likely to appear at interface, thus turning it more hydrophobic. If water is
added to a system containing aqueous brine, the salt will be diluted and the salinity, i.e., a
formulation variable, will decrease. If the system contains alcohol, a dilution with any of the
phases will result in a change in the alcohol effect which is linked to the alcohol concentration
through f(A) and φ(A) functions in equations [3-4].
In all such cases the WOR variation might induce a transitional inversion as if it were a HLD
change. This is one of the most commonly used, thought most complex way to make
nanoemulsions with systems containing commercial surfactants which produce liquid crystals
(81-83). As in the previously discussed combined WOR-HLD change, the followed path is
critical, and has to be finely tuned so that a region with L3, gel phase or lamellar liquid crystal
phase is crossed.

FIRP Booklet E237A 14 Nanoemulsions


About half a century ago Becher classified emulsification methods into four categories based
in the way the surfactant is added to the formulation: agent-in-water method, agent-in-oil-
method, nascent soap method, and alternative addition method (84). Among these methods, some
of them were applied to make cosmetic emulsions, and the agent-in-oil one was reported to result
in small droplets. The addition of water to an oil phase containing emulsifier at constant
temperature eventually produces the inversion to O/W at the so-called EIP (emulsion inversion
point) (85) and small droplets may be obtained by a process which is relatively similar to the
catastrophic inversion to be discussed later, in which the increase in internal phase ratio is the
driving force for inversion. The method has been improved by Lin in the so-called low energy
emulsification process (86). Later on, Sagitani, introduced the phase-inversion emulsification
method (87), and reported that the proper adjustment of the water addition protocol allowed the
attainment of fine emulsions (88), provided that the phase behavior undergoes through the
formation of a liquid crystal.
As an example, Sagitani (89) reports a complex three step procedure to make fine emulsions
with a system containing paraffin oil and a mixture of sorbitan esters and ethoxylated esters at
60°C. Starting with a surfactant solution in the oil phase, the first part of added water results in a
lamellar liquid crystal; the second water addition transform it in a gel emulsion (oil in surfactant
phase — i. e., a so-called D-phase microemulsion); finally the third water addition produces the
separation of the oil phase from the D phase as small droplets (82). All these studies indicate that
during the emulsification process, the formulation has to be adjusted through a programming of
HLD, so that at some time a bicontinuous microemulsion or/and a lamellar liquid crystal could
be formed. Because of its zero curvature, such phases requires a formulation which is coincident
with HLD = 0.
This dilution procedure has been reported has the second most important way (after the PIT
method) to make nanoemulsions, either of the O/W type by diluting with water an “oily”
microemulsion (61, 82, 83, 90-92), or of the W/O type by diluting with oil an “aqueous”
microemulsion (93-95).
S

2!
3! W/O
HLD (Temp.)

LC

O/W
2!
1! 2!

O W O WOR W
constant Formulation
constant Cs
Fig. 12. Coupled HLD-WOR variation produced by a WOR change.

Fig. 12 indicates various phase behavior transitions produced by a change in WOR. In the
left graph the addition of water to a oily microemulsion results in a dilution path that crosses a
liquid crystal region and then a two-phase region. (Fig 12 left ). The right graph in Fig. 12
indicate the phase behavior and inversion line in the case of a system containing a mixture of
surfactants which partition differently in the phases, e.g. a commercial polyethoxylated
alkylphenol. The slanting of the three-phase region and the associated slope of the central branch

FIRP Booklet E237A 15 Nanoemulsions


of the inversion frontier, result in a complex sequence of transitions, i. e., three different kinds of
inversion, when the WOR is changed from left to right of vice-versa along the arrows (96, 97). It
is worth remarking here that the conceptualization and design of this kind of emulsification
process requires a huge phase behavior data bank to make sure that the phase diagram to be used
contains the proper features, and this may be a though problem since phase behavior studies are
particularly tedious.
Nevertheless, the phase diagram appropriate features are not enough to warranty the
formation of a nanoemulsion. The dynamic phenomena play a quite important role to attain the
result. When a lamellar liquid crystal is to be formed by diffusion of one of the phase in a
surfactant microemulsion or gel phase, the time scale should be long enough to accommodate the
slow molecular mass transfer. On the contrary, when the dilution starts from a microemulsion,
that is likely to produce droplets by exudation, then a quick, quench-like process is likely to
perform better, as soon as the droplets are formed.
The dilution of a microemulsion does not always generate a nanoemulsion. When a single-
phase dilution produces a change in composition so that the binodal curve (i.e. the frontier
between single and two phase behaviors) is attained, a phase separation always takes place. This
can happens in different ways, either by diminishing the surfactant concentration, or by changing
WOR or HLD (particularly the temperature as in the so-called “cloud point” phenomenon) (92).
However, in general this phase splitting does not result in small drops, unless some
supersaturation region occurs, as in spontaneous emulsification by diffusion and standing (98).
Other mechanisms were recently reported to produce fine droplets by spontaneous
emulsification (99), but their rather complex classification is outside the scope of this chapter. In
most cases the droplets are initially small, often in the nanoemulsion range, but as in all
previously discussed cases, they end up coalescing unless a stabilizing treatment is quickly
applied.
In many cases of transitional inversions, the scale of the microemulsion domains is much
smaller than the resulting droplets found even in the optimal PIT emulsification. This means that
during the transition process from single-phase to two-phase behavior, tens or hundreds of
domains gather to form a single droplet as illustrated in Fig. 13 sequence. This is not a real
coalescence process since the domains are not individual droplets, but some kind of
rearrangement typical of ultra-low tension systems, in which the surfactant layer that separates
the domains shrinks and a droplet appears (25).

1 2 3 4 5

6 7 8 9 10

Fig. 13. Possible evolution of a portion of a bicontinuous microemulsion into a droplet


(courtesy J. L. Salager and R. E. Antón)

FIRP Booklet E237A 16 Nanoemulsions


The transition from microemulsion to (nano)emulsion is still a fairly unclear process and it is
not absolutely known what makes the final droplets smaller or larger. Because of the importance
of harnessing this low energy mechanism, this issue will probably be a hot topic in future
research. The first hint can be brought by other phase separation processes such as nucleation or
cloud point phenomena. In effect it is well known that the rate of formulation/temperature
change, which alters the molecular dynamics is what influences most the size of the cluster in
which the molecules are collected into the seed of the separated phase. This phase separation
case could be view to be similar to an aggregation mechanism of colloid particles into a fractal
whose shape and size depends on the limiting mechanism (reaction-balistic-diffusion) and the
aggregation mechanism (molecule-cluster or cluster-cluster) (100). The variables that are likely
to affect theses mechanisms could be tested as candidates to control the final emulsion drop size.

3.3.4. Catastrophic inversion from abnormal to normal morphology by WOR change at


constant HLD
The inversion line exhibits two almost vertical branches in the HLD-WOR plot at roughly 30
and 70% water or oil. In the central regions A+/A- (see Fig. 14), the so-called normal emulsion
obeys Bancroft’s rule (i. e., the formulation effect prevails). In the B- (respectively C+) so-called
abnormal regions there is not enough water (respectively oil) to make it the external phase, and
an abnormal situation occur in which the emulsion type does not comply with Bancroft’s rule,
being then the external phase the one in higher amount. Nevertheless, these abnormal emulsions
often exhibit a multiple morphology with small droplets (of external phase) entering the drops. It
can be said that the "inner" emulsion of the multiple morphology follows the formulation effect,
whereas the "outer" one obeys the composition requirement (WOR) (46, 51).
When a change in WOR shifts the representative points of the emulsion from one side to the
other of one of these vertical branches, a so-called catastrophic inversion takes place. The name
comes form the characteristics of the process, in particular irreversibility and delay, resulting in
hysteresis, which is fairly well described by catastrophe theory. (45) Fig. 14 left indicates that the
vertical branch in both types of diagrams is now replaced by two branches, each one
corresponding to a direction of variation of WOR. The region in between belongs to one side or
the other depending on the direction of change and is called a hysteresis zone (45, 46). The
change from abnormal to normal should be expected to be spontaneous, but it is not always the
case and it can be delayed considerably depending of many factors (101-103). Multiple
emulsions are often found to be an intermediate stage in this inversion process, in which the
"inner" emulsion tends to displace the original "outer" emulsion (104).
In some extreme cases starting from region B-, the addition of 10% aqueous phase to oil (in
direction toward region A-) with a slow and long lasting stirring, produces the inversion and
result in an ultrafine 90% internal oil in water emulsion which could be diluted afterward. The
formation of such an ultrafine emulsion, e.g. in the nanoemulsion range, comes from the growing
and prevailing of the most inner emulsion and the vanishing of the outer one by transfer of most
of the oil from external phase into tiny droplets that become included in water drops (see Fig. 14
right). This is a very early inversion, as far as the amount of added water is concerned, but a long
one if the time scale is considered. Scarce evidence indicates that the attainment of
nanoemulsions by this way might require some transfer of surfactant from oil to water or the
presence of a viscoelastic rheology that enhances the shear. Nevertheless, the fact is that in spite
of being still unclear, this method is currently used in making alkyd, silicon and epoxy
waterborne resins worldwide (47, 105, 106).
FIRP Booklet E237A 17 Nanoemulsions
W/O
+ B- A-
+ +
B A

HLD (Temp.)
C

- - -
B A C
O/W
W/O O/W/O O/W
O WOR W
constant Cs

Fig. 14 . Dynamic inversion frontiers as WOR is changed (left)


Evolution of the emulsion as water is added to oil under stirring at HLD < 0 (right)

In the normal to abnormal morphology change, e.g. from A- to B- or from A+ to C+, the
triggering of the inversion is found to depend on the rate of change, in particular to the way the
internal phase is added to the emulsion (107). If the internal phase is added slowly and
continuously, the inversion is often delayed so much that the emulsion becomes a viscoelastic
high internal phase ratio dispersion. In such viscoelastic media, a slow motion stirring produce a
fracture emulsification process in which the shear, i.e. the ratio of some velocity to some
distance, can become extremely high because the distance is reduced to the fracture gap (27, 42-
44, 108). It is the typical case of the drop of oil added to a bowl of mayonnaise, which is
emulsified with three or four twirls of a spoon, according to the mechanism described in Fig. 7.
This emulsification process is not a zero-energy one but the mechanical energy yield is so high
that it deserves to be classified as a near-zero energy case (25). The fracture emulsification
process tends to result in a monodispersed emulsion, (28) a characteristic that is welcomed to
reduce Ostwald ripening but which might result in unwelcome increased viscosity.

4. STABILITY ISSUES

4.1. EMULSION DECAY AND COALESCENCE

Nanoemulsions essentially evolve according to all mechanisms susceptible to affect any kind
of emulsion (see Fig. 15), notwithstanding with some particularities.
The typical first step in the coalescence of two drops is their approach from long distance to a
fraction of the drop size; say 0.1 µm, until a thin film forms between flattened drops. This
approach is essentially due to gravity action in macroemulsions according to Stokes’s law. In
nanoemulsions gravity has no effect, hence creaming is unlikely to take place, and the approach
is driven by Brownian motion with an energy level around 15 kT.
According to DLVO (Derjagin, Landau, Vervey, Overbeek) theory or its equivalent with no
electrical repulsion, the drop-drop interaction potential might exhibit (1) a repulsive barrier lower
than 15 kT, hence producing coalescence, (2) an energy barrier higher than 15 kT, thus with no
coalescence taking place, (3) an energy barrier larger than Brownian effect but with a secondary
minimum at “long distance” resulting in drop flocculation (100, 109, 110).

FIRP Booklet E237A 18 Nanoemulsions


original emulsion

creaming flocculation brownian motion

thin film drainage Ostwald ripening

Fig. 15 Mechanisms of emulsion decay

Because the drops are quite small the attraction force is relatively weak and the adsorption of
a “thick” layer of surfactant might be enough to insure steric repulsion with two main types:
osmotic (mixing of surfactant molecules adsorbed at approaching interfaces in spite of a good
solvent affinity) and entropic (elastic compression) repulsion at a range long enough to inhibit
the interdrop film thinning and hence to prevent coalescence and even flocculation. This is
attained by adding polymeric surfactants sometimes referred to as colloid protectors (111-113).
The presence of a liquid crystal structure wraping the droplet is of course likely to produce
the same kind of steric protection. However it has to be formed concomitantly with the droplet,
and this implies the proper phase behavior and kinetics.

4.2. OSTWALD RIPENING

Ostwald ripening is due to Laplace’s law effect on the solubility of the drop phase in the
continuous phase. Because the pressure inside a drop is higher than the pressure outside it by
∆P=2γ/r, the solubility S of the drop phase in the continuous phase just at the boundary of the
drop of radius r depends on the radius, as predicted by Kelvin equation

S(r) = S0 exp (2γ vm/rRT) [7]

where vm is the molar volume of the continuous phase and S0 is the solibility in the bulk of the
continuous phase, which is the same as the solubility close to a flat interface.
The term 2 γ vm/RT has the dimension of a length with a value of 1.5 nm for water at 25 °C
exhibiting an interfacial tension of 1 mN/m with the oil phase. Consequently the increase in
solubility close to a 100 nm droplet will be an insignificant 1.5%, whereas it will be 15% for a 10
nm droplet. This effect is thus limited to really small droplets and has no effect on
macroemulsions.

FIRP Booklet E237A 19 Nanoemulsions


The difference in solubility produces a concentration gradient from the surrounding of a
small drop to the surrounding of a larger one, which drives a mass transfer by diffusion. As a
consequence the small droplets tend to desinflate into the larger ones. Theoretically the process
wont stop until the phase separation takes place, but in practice it slows down with drop growing.
Sometimes it could end up with a bimodal distribution.
The phenomenon has been modeled by the so-called LSW (Lifshitz-Slyozov-Wagner) theory
(114) which predicts that the volume of the drops (i.e. the cube of the radius r) increases linearly
with time, i. e.,
r3 = 8/9 [S0γ vm D / ρ RT] t [8]

where D is the diffusion coefficient of the drop phase in the continuous phase, ρ is the density of
the continuous phase, and t the time.
One of the main issues with nanoemulsions is how to reduce or inhibit Ostwald ripening.
Obviously the solubility of the drop phase in the continuous phase (S0) is of paramount
importance since it appears in the proportionality constant. Thus the velocity could be reduced
by using less water-soluble oils, i. e., higher alkanes, less branched alkanes, less polar oils,
polymeric oils etc. Interfacial tension is also important as in capillary phenomena, and the
formulation can be used to reduce it considerably, says, down to 0.1 mN/m, in a quite easy way.
The diffusion coefficient D has to do with the molecular weight (MW), hence higher MW oil
would exhibit a reduced Ostwald ripening, e.g., polydimethylsiloxanes.
The effect of surfactant concentration on Ostwald ripening is not simple, and depends on the
association level of the surfactant molecules. Below the critical micelle concentration (CMC), an
increase in surfactant concentration results in a lower tension and hence smaller drops, but
generally with a more monodisperse distribution, which results with the decrease in tension in a
slow down of Ostwald ripening rate. Above the CMC the tension stays essentially constant, but
the micelles can play the role of oil carriers from drop to drop in a faster way than diffusion, and
Ostwald ripening rate might be increased (114-117), although the interpretation of the results is
still a matter of discussion (118).
At high surfactant concentration, say beyond 5-10%, the surfactant might form liquid
crystalline phases that wrap around the droplets and result in an essentially solid barrier to mass
transfer that inhibits Ostwald ripening. Another alternative is to produce the deposition of an
insoluble layer at the droplet interface (119), particularly of polymeric substances (120,122)
A clever trick to stabilize O/W nanoemulsions against Ostwald ripening is to use a mixture of
two oils with quite different solubilities in the aqueous continuous phase. The oil that migrates
first (and mostly) from the small to the large drops is the more soluble one. Large drops then
become enriched with the more soluble oil, while small ones are depleted from it and become
increasingly concentrated in the less soluble one. This difference builds up a chemical potential
of osmotic nature that eventually equilibrates the Ostwald ripening driving force (17).
According to Kabalnov (123), if the initial distribution of droplets is narrow and the
proportion of less soluble oil is high enough, the process would result in an increase in emulsion
polydispersity. However, if the initial distribution is wide and the proportion of less soluble oil is
not high enough a bimodal distribution will be achieved. Then the bigger drops evolve as an
emulsion containing essentially the more soluble oil, up to the point where they will disappear by
re-condensation of the more soluble oil in the smaller droplets. Some studies showed that
Ostwald ripening of W/O emulsions is less sensitive to the nature of oil used and slower to ripen
when compared to O/W emulsions including the same hydrocarbons (124).

FIRP Booklet E237A 20 Nanoemulsions


CONCLUSION
Nanoemulsions, also called miniemulsions, are submicron droplet emulsions, i. e., two phase
systems, and as such they follow the general phenomenology which has been reported for
emulsions in Chapter 3 of this book.
However, it is quite difficult to make them by brute force stirring, and extremely high
elongational shear devices have to be used at the expense of a huge waste of energy that often
results in overheating. Alternate low energy methods of nano-emulsification implies most of the
times a transitional phase inversion, which is the use of complex formulation programming
techniques that are still the domain of experts.
Because of their extremely small droplets nanoemulsions exhibits several specific
characteristics: (1) their droplets do not settle in the gravity field, (2) they are relatively easy to
stabilize against Brownian collision with a polymeric surfactant that produces a steric repulsion,
(3) they are sensitive to Ostwald ripening, i.e., a mass transfer from smaller to larger droplets,
which is essentially negligible in macroemulsions.

REFERENCES
1. Walstra P. In: Becher ed. Encyclopedia of Emulsion Technology. Vol. 1. New York: Marcel Dekker (1983)
57-127
2. Bourrel, M, Schechter RS. Microemulsions and Related Systems. Surfactant Science Series, Vol. 34. New
York: Marcel Dekker (1988)
3. Salager JL. Microemulsions. In: Broze G, ed. Handbook of Detergents—Part A: Properties. New York:
Marcel Dekker (1999) 253-302.
4. Salager JL. Formulation Concepts for the Emulsion Maker. In: Nielloud F and Marti-Mestres G, eds.
Pharmaceutical Emulsions and Suspensions, Chapter 2, New York: Marcel Dekker (2000) 19-72.
5. Nakajima H. Microemulsions in Cosmetics. In: Solans C, Kunieda H, eds. Industrial Application of
Microemulsions. Surfactant Science Series Vol. 66. New York: Marcel Dekker (1997) 175-197.
6. Rosano HL, Lan T, Weiss, et al. Unstable microemulsion. J Phys Chem 85, 468-473 (1981).
7. Hansrani PK, Davis SS, Groves MJ. The preparation and properties of sterile intravenous emulsions. J
Parenteral Sci Technol 37, 145-150 (1983).
8. Benita S, Levy MY. Submicron emulsions as colloidal drug carriers for intravenous administration:
comprehensive physicochemical characterization. J Pharm Sci 82, 1069-1079 (1993).
9. Chou YJ, El-Aasser MS. and Vanderhof JW. Mechanism of emulsification of styrene using
hexadecyltrimethylammonium brimide-cetyl alcohol mixtures. J Dispersion Sci Technol 1, 129-150 (1980)
10. Sudol D, El-Aasser MS. Miniemulsion polymerization. In: Lovell PA, El-Aasser MS, eds. Emulsion
Polymerization and Emulsion Polymers, John Wiley & Sons Ltd. Chichester (1997) 699-722.
11. Miñana-Pérez M, Gutron C, Zundel C, Andérez JM, Salager JL. Miniemulsion formation by transitional
inversion. J Dispersion Sci Technol 20, 893-905 (1999).
12. El-Aasser MS. Miniemulsion: An Overview of Research and Applications, J. Coating Technology Research
1, 21-31 (2004)
13. Solans C, Izquierdo P, et al. Nano-emulsions. Current Opinion Colloid Interface Sci 10, 102-110 (2005).
14. Sing AJF, Graciaa A, Lachaise J, Brochette P. Salager JL. Interactions and coalescence of nanodroplets in
translucent O/W emulsions. Colloids Surf A 152, 31-39 (1999).
15. Forgiarini A, Esquena J, González C, Solans C. Studies of the relation between phase behavior and
emulsification methods with nanoemulsion formation. Prog Colloid Polym Sci 115, 36-39 (2000)
16. Solans C, Esquena J, Forgiarini AM, Usón N., Morales D, Izquierdo P, Azemar N. Garcia MJ. Nano-
emulsions: Formation, properties and applications. In: Mittal KL. and Shah DO, eds. Adsorption and
Aggregation of Surfactants in Solutions, New York: Marcel Dekker (2003) 525-554.
17. Tadros T, Izquierdo P, Esquena J, Solans C. Formation and stability of nano-emulsions. Adv Colloid
Interface Sci 108-109, 303–318 (2004).

FIRP Booklet E237A 21 Nanoemulsions


18. Pons R, Carrera I, Erra P, Kunieda H, Solans C. Novel Preparation Methods of W/O Highly Concentrated
Emulsions, Colloids Surf A 91, 259-263 (1994).
19. Kunieda H, Fukui Y, Uchiyama H, et al. Spontaneous formation of highly concentrated water-in-oil
Emulsions (Gel Emulsions). Langmuir 12, 2136-2140 (1996).
20. Sonneville-Aubrun O, Simmonet JT, L’Alloret F. Nanoemulsions: a new vehicle for skincare products. Adv
Colloid Int Sci 108-109, 145-159 (2004).
21. Chepurnov AA, Bakulina LF, Dadaeva AA, et. al. Inactivation of ebola virus with a surfactant nanoemulsion.
Acta Trop 87, 315-320 (2003).
22. Myc A, Kukowska-Latallo JF Bielinska AU, et al. Development of immune response that protects mice from
viral pneumonitis after a single intranasal immunization with influenza A virus and nanoemulsion. Vaccine
21, 3801-3814 (2003).
23. Salager JL, Márquez N, Graciaa A. et al. Partitioning of ethoxylated octylphenol surfactants in
microemulsion-oil-water systems. Influence of temperature and relation between partitioning coefficient and
physicochemical formulation. Langmuir 16, 5534-5539 (2000).
24. Graciaa A, Lachaise J, Sayous J.G, ., Grenier P., Yiv S., Schechter R. S., Wade W. H.. The partitioning of
complex surfactant mixtures between oil-water-microemulsion phases at high surfactant concentration. J
Colloid Interface Sci 93, 474-486 (1983).
25. Salager JL, Forgiarini A, Lopez JC, et al. Dynamics of near-zero energy emulsification, Paper # 203, CD
Proceedings 6th World Surfactant Congress CESIO, Berlin, Germany, June 21-23, 2004.
26. Grace HP. Dispersion phenomena in high viscosity immiscible fluid systems and application of static mixers
as dispersion devices in such systems. Chem Eng Communication 14, 225-277 (1982).
27. Mason TG, Bibette J. Shear rupturing of droplets in complex fluids. Langmuir 13, 4600-4613 (1997).
28. Mabille C, Schmitt V, Gorria P, et al. Rheological and shearing conditions for the preparation of
monodispersed emulsions. Langmuir 2000, 16, 422-429.
29. Canselier JP, Delmas H, Wilhelm AM. Ultrasound emulsification — an overview. J Dispersion Sci Tecnol
23, 333-349 (2002).
30. Joscelyne SM, Trägårdh G. Membrane emulsification — A literature review. J Membrane Sci 169, 107-117
(2000)
31. Joscelyne SM,Trägårdh G. Food emulsions using membrane emulsification: conditions for producing small
droplets. J Food Engineering 39, 59-64 (1999).
32. Nakashima T, Shimizu M, Kukizaki M. Particle control of emulsion by membrane emulsification and its
applications. Adv Drug Delivery Reviews 45, 47–56 (2000).
33. Christov NC, Ganchev DN, Vassileva ND, et al. Capillary mechanism in membrane emulsification: oil-in-
water emulsions stabilized by Tween 20 and milk proteins. Colloids Surf A 209, 83-104 (2002).
34. Sugiura S, Nakajima M, Seki M. Prediction of droplet diameter for microchannel emulsification. Langmuir
18, 3854-3859 (2002)
35. Jeffreys GV, Davies GA. Coalescence of liquid droplets and liquid dispersions. In: Hansen C, ed. Advances
in liquid-liquid extraction, New York: Pergamon Press (1971) 495-584.
36. Taisne L, Walstra P, Cabane B. Recoalescence during emulsification Paper 1-2-014, 2nd World Congress on
Emulsion, Bordeaux, France, Sep. 23-26, 1997.
37. Bourrel M, Graciaa A, Schechter RS, Wade WH. The relation of emulsion stability to phase behavior and
interfacial tension of surfactant systems. J Colloid Interface Sci 72, 161-163 (1979).
38. Salager JL, Quintero L, Ramos E, Andérez J. Properties of surfactant-oil-water emulsified systems in the
neighborhood of three-phase transition. J Colloid Interface Sci 77, 288-289 (1980).
39. Antón RE, Salager JL. Emulsion stability in the three-phase behavior region of surfactant-alcohol-oil-brine
system. J Colloid Interface 111, 54-59 (1986).
40. Pérez M, Zambrano N, Ramirez M, et al. Surfactant-oil-water systems near the affinity inversion. Part XII:
Emulsion drop size versus formulation and composition. J Dispersion Sci Technol 23, 55-63 (2002).
41. Princen HM. The structure mechanics and rheology of concentrated emulsions and fluid foams. In: Sjöblom
J, ed. Encyclopedic Handbook of Emulsion Technology. New York: Marcel Dekker (2001) 243-278.
42. Lin TJ. Low-energy emulsification. Part I: Principles and applications. J Society Cosmetic Chemists 29, 117-
125 (1978).
43. Lin TJ, Akabori T, Tanaka S, Shimura K. Low-energy emulsification. Part II: Evaluation of emulsion
quality. J Society Cosmetic Chemists 29, 745-756 (1978).
44. Lin TJ, Akabori T, Tanaka S, Shimura K. Low-energy emulsification. Part III: Emulsification in high alpha
range. Cosmetics Toiletries 95, 33-39 (1980).

FIRP Booklet E237A 22 Nanoemulsions


45. Salager JL. Phase transformation and emulsion inversion on the basis of catastrophe theory. In: Becher P, ed.
Encyclopedia of Emulsion Technology. vol. 3, Chapter 2, New York: Marcel Dekker (1988) 79-134.
46. Salager JL, Márquez L, Peña AA, Rondón M, Silva F, Tyrode E. Current phenomenological know-how and
modeling of emulsion inversion. Ind Eng Chem Res 39, 2665-2676 (2000).
47. Salager JL, Forgiarini A, Marquez L, Pizzino A., Rodriguez MP, Rondón M. Using emulsion inversion in
industrial processes. Adv Colloid Int Sci 108-109, 259-272 (2004).
48. Shinoda K, Saito H. The stability of O/W type emulsions as a function of temperature and the HLB of
emulsifiers: The emulsification by PIT-method. J Colloid Interface Sci 30, 258-263 (1969).
49. Förster T, Schambil F, Von Rybinski W. Production of fine disperse and long term stable oil-in-water
emulsion by the phase inversion temperature method. J Dispersion Sci Technol 13, 183-193 (1992).
50. Miller D, Henning T, Grunbein W. Phase inversion of W/O emulsions by adding hydrophilic surfactant – a
technique for making cosmetic products. Colloids Surf A 183-185, 681-688 (2001).
51. Salager JL, Miñana-Pérez M, Pérez-Sánchez M, et al. Surfactant-oil-water systems near the affinity inversion
- Part III: The two kinds of emulsion inversion, J. Dispersion Sci Technol 4, 313-329 (1983).
52. Salager JL, Antón RE, Sabatini DA, et al. Enhancing solubilization in microemulsions – state of the art and
current trends. J Surfactants Detergents 8, 3-21 (2005).
53. Hazlett RD, Schechter RS. Stability of macroemulsions. Colloids Surf 29, 53-69 (1988).
54. Kabalnov A, Wennerström H. Macroemulsions stability: the oriented wedge theory revisited. Langmuir 12,
276-292 (1996).
55. Kabalnov A, Weers J. Macroemulsions stability within the Winsor III region: Theory versus experiments.
Langmuir 12, 1931-1937 (1996)
56. Taisne L and Cabane B. Emulsification and ripening following a temperature Quench. Langmuir 14, 4744-
4752 (1998).
57. Friberg S, Mandell L, Larsson M. Mesomorphous phases, a factor of importance for the properties of
emulsions. J Colloid Interface Sci 29, 155-156 (1969).
58. Friberg S, Jansson PO. Surfactant association structure and emulsion stability. J Colloid Interface Sci 55,
614-623 (1976).
59. Friberg, S.E. & Solans, C. Surfactant association structures and the stability of emulsions and foams.
Langmuir 2, 121-126 (1986).
60. Forgiarini A, Esquena J, González C, Solans C. Formation and stability of nano-emulsions in mixed nonionic
surfactant systems. Prog Colloids Polymer Sci 118, 184-189 (2001).
61. Suzuki T., Takei H., Yamazaki S. Formation of fine three-phase emulsions by the liquid crystal
emulsification method with Arginine β-branched monoalkyl phosphate. J Colloid Interface Sci 129, 491-500
(1989).
62. Suzuki T, Tsutsumi H, Ishida A. Secondary droplet emulsion: Mechanism and effects of liquid crystal
formation in O/W emulsion. J Disp Sci Technol 5, 119-141 (1984)
63. Salager JL. Emulsion properties and related know-how to attain them. In: Nielloud F & Marti-Mestres G,
eds. Pharmaceutical Emulsions and Suspensions, Chapter 3, New York: Marcel Dekker (2000) 73-125.
64. Förster T. Principles of emulsion formation in surfactants. In: Rieger MM, Rhein LD, eds. Surfactants in
Cosmetics, 2nd Ed., New York: Marcel Dekker (1997) 105-125.
65. Friberg S, Solans C. Emulsification and HLB-temperature. J Colloid Interface Sci 66, 367-368 (1978).
66. Márquez L, Graciaa A, Lachaise J, Salager JL, Zambrano N. Hysteresis behavior in temperature-induced
emulsion inversion. Polymer International 52, 590-593 (2003).
67. Friberg S, Solans C. Emulsification and the HLB-temperature. J Colloid Interface Sci 66, 367-368 (1978).
68. Izquierdo P, Feng J, Esquena J, et al. The influence of surfactant mixing ratio on nano-emulsion formation
by the PIT method. J Colloid Interface Sci 285, 388-394 (2005).
69. Förster T, Schambil F, and Tersmann H. Emulsification by the phase inversion temperature method: the role
of self-body agents and the influence of oil polarity. Int J Cosmetic Sci 12, 217-227 (1990).
70. Förster T, Rybinski W, Wadle A. Influence of microemulsion phases on the preparation of fine-disperse
emulsions, Adv Colloids Interface Sci 58, 119-149 (1995).
71. Förster T. Principles of emulsions formation In: Rieger M and Rhein LD, eds. Surfactants in Cosmetics.Vol
68 New York: Marcel Dekker (1997) 105-125.
72. Osawa K, Solans C, and Kunieda H. Spontaneous formation of highly concentrated oil-in-water emulsions. J
Colloid Interface Sci 188, 275-281 (1997).

FIRP Booklet E237A 23 Nanoemulsions


73. Allouche J, Tyrode E, Sadtler V, Choplin L, Salager JL. Evolution of emulsion properties along a transitional
phase inversion process driven by temperature variation. 3rd World Congress on Emulsion, Lyon, France
Sept. 23-27, 2002.
74. Morales D, Gutierrez JM, Garcia-Celma MJ, et al. A study of the relation between bicontinuous
microemulsions and oil/water nano-emulsion formation. Langmuir 19, 7196–7200 (2003).
75. Izquierdo P, Esquena J, Tadros T, et al. Phase behavior and nano-emulsion formation by the phase inversion
temperature method. Langmuir 20, 6594-6598 (2004).
76. Márquez L, Graciaa A, Lachaise J, Salager JL. A 3rd type of emulsion inversion attained by overlapping the
two classical methods: combined inversion. 3rd World Congress on Emulsion, Lyon, France Sept. 23-27,
2002.
77. Salager JL, Pérez-Sánchez M, Garcia Y. Physicochemical parameters influencing the emulsion drop size.
Colloid Polymer Sci 274, 81-84 (1996).
78. Tolosa L, Forgiarini A, Moreno P, Salager JL. Combined effects of formulation and stirring on emulsion
drop size in the vicinity of three-phase behavior of surfactant-oil water systems, Ind Eng Chem Res 45, 3810-
3814 (20006)
79. Graciaa A, Lachaise J, Bourrel M, et al. Partitioning of nonionic and anionic surfactant mixtures between
oil/microemulsion/water phases, SPE Reservoir Eng 2, 305-314 (1987).
80. Graciaa A, Andérez JM, Bracho C, Lachaise J, Salager JL, Tolosa L. The Selective Partitioning of the
Oligomers of Polyethoxylated Surfactant Mixtures between Interface and Oil and Water bulk Phases. Adv
Colloid Interfacial Sci. 123-126, 63-73 (2006)
81. El-Aasser MS, Lack CD, Vanderhoff JW, et al. The miniemulsification process – different form of
spontaneous emulsification. Colloids Surfaces 29, 103-118 (1988).
82. Sagitani H. Formation of O/W emulsions by surfactant phase emulsification and the solution behavior of
nonionic surfactant system in the emulsification process. J Dispersion Sci Technol 9, 115-129 (1988).
83. Forgiarini A, Esquena J, Gonzalez C, Solans C. Formation of nano-emulsions by low energy emulsification
methods at constant temperature. Langmuir 17, 2076-2083 (2001)
84. Becher P, Emulsions: Theory and Practice, 2nd ed. Reprint. New York: RE Krieger Pub. (1965) 267-268.
85. Marszall L. HLB of nonionic surfactants: PIT and EIP methods. In: Schick MJ ed. Nonionic Surfactants
Physical Chemistry. New York: Marcel Dekker (1987) 493-548.
86. Lin TJ, Akabori T, Tanaka S, et al. Low-energy emulsification. Part V: Mechanism of enhanced
emulsification. Cosmetics Toiletries 98, 67-78 (1983).
87. Sagitani H, Making homogeneous and fine droplet O/W emulsions using nonionic surfactants. J Am Oil
Chem Soc 58, 738–743 (1981).
88. Sagitani H. Phase Inversion and D Phase Emulsification. In Friberg SE and Lindman B, eds. Organized
Solutions. New York: Marcel Dekker (1992) 259-271.
89. Sagitani H, Hattori T, Nabeta K, Nagai M. Formation of O/W emulsion having fine and uniform droplets by
the surfactant (D) phase emulsification method. J Chem Soc Jpn 1399–1404 (1983).
90. Dai L, Li W, Hou X. Effect of the molecular structure of mixed nonionic surfactants on the temperature of
miniemulsion formation. Colloids Surf A 125, 27-32 (1997).
91. Greiner RW, Evans DF. Spontaneous formation of a water-continuous emulsion from a W/O microemulsion.
Langmuir 6, 1793-1796 (1990).
92. Pons R, Carrera I, Caelles J, Solans C. Formation and properties of miniemulsions formed by
microemulsions dilution. Adv Colloid Int Sci 106, 129-146 (2003).
93. Esquena J, Solans C. Study of low energy emulsification methods for the preparation of narrow size
distribution W/O emulsions. Progress Colloid Polymer Sci 110, 235-239 (1998).
94. Forgiarini A, Esquena J, González C, Solans C. The relation between phase behavior and formation of
narrow size distribution W/O emulsions. J Dispersion Sci Technol 23, 209-217 (2002).
95. Uson N, Garcia MJ, Solans C. Formation of water-in-oil (W/O) nano-emulsions in a water/mixed non-ionic
surfactant/oil systems prepared by a low-energy emulsification method. Colloids Surf A 250, 415-421 (2004)
96. Sajjadi S, Jahanzad F, Yiannekis M. Catastrophic phase inversion of abnormal emulsions in the vicinity of
the locus of transitional inversion. Colloids Surf A 240, 149-155 (2004).
97. Marfisi S, Rodríguez MP, Álvarez G, Celis MT, Forgiarini A, Lachaise J, Salager JL. Complex emulsion
inversion pattern associated with the partitioning of nonionic surfactant mixtures in presence of alcohol
cosurfactant. Langmuir 21, 6712-6716 (2005).
98. Ruschak KJ, Miller CA. Spontaneous emulsification in ternary systems with mass transfer. Ind Eng Chem
Fundamentals 11, 534-540 (1972).

FIRP Booklet E237A 24 Nanoemulsions


99. Lopez-Montilla JC, Herrera-Morales PE, Pandey S. Spontaneous emulsification: mechanisms,
physicochemical aspects, modeling and applications, J Dispersion Sci Technol 23, 219-268 (2002)
100. Hiemenz P. C., Rajagopalan R. Principles of Colloid and Surface Chemistry, New York: Marcel Dekker 3rd
Ed. 1997.
101. Zambrano N, Tyrode E, Mira I, Marquez L, Rodriguez MP, Salager JL. Emulsion catastrophic inversion from
abnormal morphology to normal one. Part 1: Effect of water-to-oil ratio rate of change on the dynamic
inversion frontier. Ind Eng Chem Res 42, 50-56 (2003)
102. Mira I, Zambrano N, Tyrode E, Marquez L, Peña AA, Pizzino A Salager JL. Emulsion catastrophic inversion
from abnormal morphology to normal one. Part 2: Effect of stirring intensity on the dynamic inversion
Frontier. Ind Eng Chem Res 42, 57-61 (2003)
103. Tyrode E, Mira I, Zambrano N, Marquez L, Rondón-González M, Salager JL. Emulsion catastrophic
inversion from abnormal morphology to normal one. Part 3: Conditions for triggering the dynamic inversion
and applications to industrial processes. Ind Eng Chem Res 42, 4311-4318 (2003)
104. Tyrode E, Allouche J, Choplin L, Salager JL. Emulsion catastrophic inversion from abnormal morphology to
normal one. Part 4. Following the emulsion viscosity during three inversion protocols and extending the
critical dispersed-phase concept. Ind Eng Chem Res 44, 67-74 (2005)
105. Rouvière J. Stabilisation d’émulsions huile-dans-eau par formation de multicouches de tensioactifs: cas des
huiles de silicone, Informations Chimie 325, 158-164 (1991).
106. Yang ZZ, Zhao D. Preparation of bisphenol A epoxy resin waterborne dispersion by the phase inversion
emulsification technique. Chin J Polymer Sci 18, 33-38 (2000)
107. Bouchama F, Aken GA van, Autin AJE, Koper G. J. M. On the mechanism of catastrophic phase inversion in
emulsions. Colloids Surf A 231, 11-17 (2003).
108. Mabille C., Schmitt V., Gorria P., Leal-Calderon F., Faye V., Deminiere B., Bibette J. Rheological and
shearing conditions for the preparation of monodispersed emulsions, Langmuir 16, 422-429 (2000).
109. Denkov ND, Petsev DN and Danov KD. Interaction between deformable Brownian droplets. Phys. Rev. Lett.
71, 3226–3229 (1993)
110. Petsev DN. Theoretical analysis of film thickness transition dynamics and coalescence of charged
miniemulsion droplets. Langmuir 16, 2093-2100 (2000).
111. Tadros ThF, Vandamme A, Levecke B, et al. Stabilization of emulsion using polymeric surfactants based on
inulin. Adv Colloid Interface Sci 108-109, 207-226 (2004).
112. Chern CS, Liou YC. Effects of mixed surfactants on the styrene miniemulsion polymerization in the presence
of an alkyl methacrylate. Macromolecular Chem Phys 199, 2051-2061 (1998).
113. Savicl S, Vuleta1G, Daniels R and Müller-Goymann CC. Colloidal microstructure of binary systems and
model creams stabilized with an alkylpolyglucoside non-ionic emulsifier. Colloid Polymer Sci 283, 439-451
(2005)
114. Taylor P. Ostwald ripening in emulsions. Adv Colloid Interface Sci 75, 107-163 (1998).
115. Capek I. Degradation of kinetically-stable o/w emulsions. Adv Colloid Interface Sci 107, 125-155 (2004)
116. Soma J, Papadopoulos KD. Ostwald ripening in sodium dodecyl sulfate-stabilized decane-in-water
emulsions. J Colloid Interface Sci 181, 225-231 (1996)
117. Weiss J, Canceliere C, McClements DJ. Mass transport phenomena in oil-in-water emulsions containing
surfactant micelles: Ostwald ripening. Langmuir 16, 6833-6838 (2000)
118. Kabalnov AS, Can micelles mediate a mass transfer between oil droplets? Langmuir 10, 680-684 (1994)
119. Meinders MBJ, van Vliet T. The role of interfacial rheological properties on Ostwald ripening in emulsions.
Adv Colloid Interface Sci 108-109, 119-126 (2004).
120. Baskar G, Landfester K, and Antonietti M. Comb-like polymers with octadecyl side chain and carboxyl
functional sites: scope for efficient use in miniemulsion polymerization, Macromolecules 33, 9228 (2000).
121. Aoki T, Eric A, Decker EA, McClements DJ. Influence of environmental stresses on stability of O/W
emulsions containing droplets stabilized by multilayered membranes produced by a layer-by-layer
electrostatic deposition technique. Food Hydrocolloids 19, 209–220 (2005).
122. Sadtler VM, Imbert P, Dellacherie E. Ostwald Ripening of Oil-in-Water Emulsions Stabilized by Phenoxy-
Substituted Dextrans. J Colloid Interface Sci 254, 355–361 (2002).
123. Kabalnov A. S., Pertzov A. V., Shchukin E. D. Ostwald Ripening in Emulsions. 1. Direct Observations of
Ostwald Ripening in Emulsions. J Colloid Interface Sci 118, 590-597 (1987).
124. Jiao J, Burgess DJ. Ostwald ripening of water-in-hydrocarbon emulsions. J Colloid Interface Sci 264, 509–
516 (2003)

FIRP Booklet E237A 25 Nanoemulsions


Text: NANOEMULSIONS
Author: Jean-Louis SALAGER, Ana FORGIARINI, Laura MARQUEZ
Reference: FIRP Booklet # E237A
Version # 1 (2006)
Edited and published by:
Laboratorio FIRP
Escuela de INGENIERIA QUIMICA,
UNIVERSIDAD de Los ANDES
Mérida 5101 VENEZUELA

Conditions of use

FIRP Booklets are offered free of charge to teachers


and students, and may be downloaded and reproduced for
individual use only.
Reproduction for sale or as a support of paid courses
is forbidden without an autorization from the editor
(firp@ula.ve)

Laboratorio FIRP, Telephone ++58-274-2402954- 2402815 Fax ++58-274-2402957


Email firp@ula.ve Web site http://www.firp.ula.ve
Escuela de INGENIERIA QUIMICA,
UNIVERSIDAD de Los ANDES Mérida 5101 VENEZUELA

FIRP Booklet E237A 26 Nanoemulsions

You might also like