You are on page 1of 16

Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Formation and stability of nano-emulsions


Tharwat Tadrosa,*, P. Izquierdob, J. Esquenab, C. Solansb
a
89 Nash Grove Lane, Wokingham, Berkshire RG40 4HE, UK
b
CSIC, Jordi Jirona 18-26, Barcelona 08034, Spain

Abstract

This review describes the principles of formation and stability of nano-emulsions. It starts with an introduction highlighting the
main advantages of nano-emulsions over macroemulsions for personal care and cosmetic formulations. It also describes the main
problems with lack of progress on nano-emulsions. The second section deals with the mechanism of emulsification and the
dynamic light scattering technique for measurement of the droplet size of nano-emulsions. This is followed by a section on
methods of emulsification and the role of surfactants. Three methods are described for nano-emulsion preparation, namely high
energy emulsification (using homogenisers), low energy emulsification whereby water is added to an oil solution of the surfactant
and the principle of the phase inversion temperature (PIT). A section is devoted to steric stabilisation and the role of the adsorbed
layer thickness. The problem of Ostwald ripening (which is the main instability process of nano-emulsions) is described in some
detail. The methods that can be applied to reduce Ostwald ripening are briefly described. This involves the addition of a second
less soluble oil phase such as squalene andyor addition of a strongly adsorbed and water insoluble polymeric surfactant. The last
part of the review gives some examples of nano-emulsions that are prepared by the PIT method as well as using high pressure
homogeniser. A comparison of the two methods is given and the rate of Ostwald ripening is measured in both cases. The effect
of changing the alkyl chain length and branching of the oil was investigated using decane, dodecane, tertadecane, hexadecane
and isohexadecane. The branched oil isohexadcecane showed higher Ostwald ripening rate when compared with a linear chain
oil with the same carbon number.
䊚 2003 Elsevier B.V. All rights reserved.

Keywords: Nano-emulsions; Phase inversion temperature (PIT); Emulsification

1. Introduction Unless adequately prepared (to control the droplet


size distribution) and stabilised against Ostwald ripening
Nano-emulsions are transparent or translucent systems (that occur when the oil has some finite solubility in
mostly covering the size range 50–200 nm w1,2x. Nano- the continuous medium), nano-emulsions may loose
emulsions were also referred to as mini-emulsions w3,4x. their transparency with time as a result of increase in
Unlike microemulsions (which are also transparent or droplet size.
translucent and thermodynamically stable) nano-emul- The attraction of nano-emulsions for application in
sions are only kinetically stable. However, the long-term personal care and cosmetics as well as in health care is
physical stability of nano-emulsions (with no apparent due to the following advantages:
flocculation or coalescence) make them unique and they (i) The very small droplet size causes a large reduc-
are sometimes referred to as ‘Approaching Thermody- tion in the gravity force and the Brownian motion may
namic Stability’. be sufficient for overcoming gravity. This means that no
The inherently high colloid stability of nano-emul- creaming or sedimentation occurs on storage. (ii) The
sions can be well understood from a consideration of small droplet size also prevents any flocculation of the
their steric stabilisation (when using non-ionic surfac- droplets. Weak flocculation is prevented and this enables
tants andyor polymers) and how this is affected by the the system to remain dispersed with no separation. (iii)
ratio of the adsorbed layer thickness to droplet radius as The small droplets also prevent their coalescence, since
will be discussed below.
these droplets are non-deformable and hence surface
*Corresponding author. Tel.: q44-1189-732621. fluctuations are prevented. In addition, the significant
E-mail address: tharwat@tadros.fsnet.co.uk (T. Tadros). surfactant film thickness (relative to droplet radius)

0001-8686/04/$ - see front matter 䊚 2003 Elsevier B.V. All rights reserved.
doi:10.1016/j.cis.2003.10.023
304 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

prevents any thinning or disruption of the liquid film solubility andyor incorporation of polymeric surfactants
between the droplets. (iv) Nano-emulsions are suitable that strongly adsorb at the OyW interface (which are
for efficient delivery of active ingredients through the also insoluble in the aqueous medium). (viii) Fear of
skin. The large surface area of the emulsion system introduction of new systems without full evaluation of
allows rapid penetration of actives. (v) Due to their the cost and benefits.
small size, nano-emulsions can penetrate through the Inspite of the above difficulties, several companies
‘rough’ skin surface and this enhances penetration of have introduced nano-emulsions in the market and
actives. (vi) The transparent nature of the system, their within the next few years, the benefits will be evaluated.
fluidity (at reasonable oil concentrations) as well as the Nano-emulsions were used in the pharmaceutical field
absence of any thickeners may give them a pleasant as drug delivery systems w7x.
aesthetic character and skin feel. (vii) Unlike microe- Acceptance of nano-emulsions as a new type of
mulsions (which require a high surfactant concentration, formulation depends on customer perception and accept-
usually in the region of 20% and higher), nano-emul- ability. With the advent of new instruments for high
sions can be prepared using reasonable surfactant con- pressure homogenizers and the competition between
centration. For a 20% OyW nano-emulsion, a surfactant various manufacturers, the cost of production of nano-
concentration in the region of 5–10% may be sufficient. emulsions will decrease and that may approach that of
(viii) The small size of the droplets allows them to classical macroemulsions.
deposit uniformly on substrates. Wetting, spreading and Fundamental research in investigation of the role of
penetration may be also enhanced as a result of the low surfactants in the process w5,6x will lead to optimized
surface tension of the whole system and the low inter- emulsifier systems and more economic use of surfactants
facial tension of the OyW droplets. (ix) Nano-emulsions will emerge. The importance of phase behaviour in the
can be applied for delivery of fragrants, which may be preparation of nano-emulsions is also very crucial. As
incorporated in many personal care products. This could we will see later the method of mixing of oil, water and
also be applied in perfumes, which are desirable to be surfactant is very important.
formulated alcohol free. (x) Nano-emulsions may be In this review, I will discuss the following topics: (i)
applied as a substitute for liposomes and vesicles (which Fundamental principles of emulsification and the role of
are much less stable) and it is possible in some cases surfactants. (ii) Production of nano-emulsions using:
to build lamellar liquid crystalline phases around the high pressure homogenizers. The phase inversion tem-
nano-emulsion droplets. perature (PIT) principle. (iii) Theory of steric stabili-
Inspite of the above advantages, nano-emulsions have zation of emulsions. The role of the relative ratio of
only attracted interest in recent years for the following adsorbed layer thickness to the droplet radius. (iv)
reasons: (i) Preparation of nano-emulsions requires in Theory of Ostwald ripening and methods of reduction
many cases special application techniques, such as the of the process: incorporation of a second oil phase with
use of high pressure homogenisers as well as ultrasonics. very low solubility. Use of strongly adsorbed polymeric
Such equipments (such as the Microfluidiser) became surfactants. (iv) Examples of recently prepared nano-
available only in recent years. (ii) There is a perception emulsions and investigation of the above effects.
in the personal care and cosmetic industry that nano-
emulsions are expensive to produce. Expensive equip- 2. Mechanism of emulsification
ments are required as well as the use of high
concentrations of emulsifiers. (iii) Lack of understand- To prepare an emulsions oil, water, surfactant and
ing of the mechanism of production of submicron energy are needed. This can be considered from a
droplets and the role of surfactants and cosurfactants. consideration of the energy required to expand the
(iv) Lack of demonstration of the benefits that can be interface, DAg (where DA is the increase in interfacial
obtained from using nano-emulsions when compared area when the bulk oil with area A1 produces a large
with the classical macroemulsion systems. (v) Lack of number of droplets with area A2; A24A1, g is the
understanding of the interfacial chemistry that is interfacial tension). Since g is positive, the energy to
involved in production of nano-emulsions. For example, expand the interface is large and positive. This energy
few formulations chemists are aware of the use of the term cannot be compensated by the small entropy of
phase inversion temperature (PIT) Concept and how dispersion TDS (which is also positive) and the total
this can be usefully applied for the production of small free energy of formation of an emulsion, DG is positive,
emulsion droplets. (vi) Lack of knowledge on the
mechanism of Ostwald ripening, which is perhaps the DGsDAg y TDS (1)
most serious instability problem with nano-emulsions.
(vii) Lack of knowledge of the ingredients that may be Thus, emulsion formation is non-spontaneous and
incorporated to overcome Ostwald ripening. For exam- energy is required to produce the droplets. The formation
ple, addition of a second oil phase with very low of large droplets (few micrometers) as is the case for
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 305

macroemulsions is fairly easy and hence high speed is induced in the droplets, thereby reradiating the light.
stirrers such as the Ultraturrax or Silverson Mixer is Due to the random position of the droplets, the intensity
sufficient to produce the emulsion. In contrast the of scattered light will, at any instant appears as a random
formation of small drops (submicron as is the case with diffraction or ‘speckle’ pattern. As the droplets undergo
nano-emulsions) is difficult and this requires a large Brownian motion, the random configuration of the
amount of surfactant andyor energy. pattern will, therefore, fluctuate such that the time taken
The high energy required for formation of nano- for an intensity maximum to become a minimum, i.e.
emulsions can be understood from a consideration of the coherence time corresponds exactly to the time
the Laplace pressure p (the difference in pressure required for the droplet to move one wavelength. Using
between inside and outside the droplet, a photomultiplier of active area about the diffraction
maximum, i.e. one coherence area, this intensity fluctu-
B 1 1E ation can be measured. The analogue output is digitised
psgC q F (2) using a digital correlator that measures the photocount
D R1 R2 G
(or intensity) correlation function of the scattered light.
The photocount correlation function G (2)(t) is given by
where R1 and R2 are the principal radii of curvature of the equation,
the drop.
For a spherical drop, R1sR2sR and, 2
GŽ2.Žt.sB 1 q g2 wygŽ1.Žt.z~
Ž x |

. (4)
2g
ps (3) where t is the correlation delay time. The correlator
R
compares G (2)(t) for many values of t. B is the
To break up a drop into smaller ones, it must be background value to which G (2)(t) decays at long delay
strongly deformed and this deformation increases p. times. g (1)(t) is the normalised correlation function of
This can be shown when a spherical drop deforms into the scattered electric field and g is a constant (approx.
a prolate ellipsoid. For a spherical drop, there is only 1).
one radius of curvature Ra, whereas for a prolate For monodisperse non-interacting droplets,
ellipsoid there are two radii of curvature Rb,1 and Rb,2.
Consequently, the stress needed to deform the drop is gŽ1.sexpŽy Gt. (5)
higher for a smaller drop. Since the stress is generally
transmitted by the surrounding liquid via agitation, where G is the decay rate or inverse coherence time,
higher stresses need more vigorous agitation, hence more that is related to the translational diffusion coefficient
energy is needed to produce smaller drops w8x. D by the equation,
Surfactants play major roles in the formation of nano-
emulsions: By lowering the interfacial tension, p is GsDK2 (6)
reduced and hence the stress needed to break up a drop
is reduced. Surfactants prevent coalescence of newly where K is the scattering vector,
formed drops.
Various processes occurring during emulsification, 4pn BuE
namely break up of droplets, adsorption of surfactants Ks sinC F (7)
and droplet collision (which may or may not lead to lo D2G

coalescence occur w8x. Each of these processes occurs


numerous times during emulsification and the time scale l is the wavelength of light in vacuo, n is the refractive
of each process is very short, typically a microsecond. index of the solution and u is the scattering angle.
This shows that the emulsification process is a dynamic The droplet radius R can be calculated from D using
process and events that occur in a microsecond range the Stokes–Einstein equation,
could be very important. To describe emulsion formation
one has to consider two main factors: Hydrodynamics kT
Ds (8)
and interfacial science. 6p ho R
To assess nano-emulsion formation, one usually meas-
ures the droplet size distribution using dynamic light ho is the viscosity of the medium.
scattering techniques (photon correlation spectroscopy, The above analysis is valid for dilute monodisperse
PCS). In this technique, one measures the intensity droplets. With many nano-emulsions, the droplets are
fluctuation of scattered light by the droplets as they not perfectly mono-disperse (usually with a narrow size
undergo Brownian motion w9x. When a light beam passes distribution) and the light scattering results are analysed
through a nano-emulsion, an oscillating dipole moment for polydispersity (the data are expressed as an average
306 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

size and a polydispersity index that gives information During emulsification an increase in the interfacial
on the deviation from the average size. area A takes place and this causes a reduction in G. The
equilibrium is restored by adsorption of surfactant from
3. Methods of emulsification and the role of the bulk, but this takes time (shorter times occur at
surfactants higher surfactant activity). Thus, ´ is small at small a
and also at large a. Because of the lack or slowness of
With macroemulsions, several procedures may be equilibrium with polymeric surfactants, ´ will not be
applied for emulsion preparation, these range from the same for expansion and compression of the interface.
simple pipe flow (low agitation energy L), static mixers In practice, surfactant mixtures are used and these have
and general stirrers (low to medium energy, L–M), high pronounced effects on g and ´: Some specific surfactant
speed mixers such as the Ultraturrex (M), colloid mills mixtures give lower g values than either of the two
and high pressure homogenizers (high energy, H), ultra- individual components. The presence of more than one
sound generators (M–H). The method of preparation surfactant molecule at the interface tends to increase ´
can be continuous (C) or batch-wise (B). With nano- at high surfactant concentrations. The various compo-
emulsions, however, a higher power density is required nents vary in surface activity. Those with the lowest g
and this restricts the preparation of nano-emulsions to tend to predominate at the interface, but if present at
the use of high pressure homogenisers and ultrasonics. low concentrations, it may take long time before reach-
An important parameter that describes droplet defor- ing the lowest value. Polymer-surfactant mixtures may
mation is the Weber number, We, which gives the ratio show some synergetic surface activity.
of the external stress Gh (where G is the velocity Apart for their effect on reducing g, surfactants play
gradient and h is the viscosity) over the Laplace major roles in deformation and break-up of droplets.
pressure w10x, Surfactants allow the existence of interfacial tension
gradients, which is crucial for formation of stable drop-
Ghr
Wes (9) lets. In the absence of surfactants (clean interface), the
2g interface cannot withstand a tangential stress; the liquid
motion will be continuous. If a liquid flows along the
The droplet deformation increases with increase in
interface with surfactants, the latter will be swept down-
the Weber number, which means that for producing
stream causing an interfacial tension gradient. A balance
small droplets one requires high stresses (high shear
of forces will be established,
rates). In other words, the production of nano-emulsions
costs more energy than that required to produce macroe- B dV E dy
hC F
x
mulsions w4x. sy (12)
Surfactants lower the interfacial tension g and this D dy Gys0 dx
causes a reduction in droplet size. The latter decrease
If the g-gradient can become large enough, it will
with decrease in g. For Turbulent Inertial regime, the
arrest the interface. If the surfactant is applied at one
droplet diameter is proportional to g3y5.
site of the interface, a g-gradient is formed that will
The amount of surfactant required to produce the
cause the interface to move roughly at a velocity given
smallest drop size will depend on its activity a (concen-
by,
tration) in the bulk which determines the reduction in
g, as given by the Gibbs adsorption equation,
Vs1.2 Žhrz.y1y3 )Dg)2y3 (13)
ydgsRTGdln a (10)
The interface will then drag some of the bordering
where R is the gas constant, T is the absolute temperature liquid with it. Interfacial tension gradients are very
and G is the surface excess (number of moles adsorbed important in stabilising the thin liquid film between the
per unit area of the interface). droplets, which is very important during the beginning
G increases with increase in surfactant concentration of emulsification (films of the continuous phase may be
and eventually it reaches a plateau value (saturation drawn through the disperse phase and collision is very
adsorption). The value of g obtained depends on the large).
nature of the oil and surfactant used. Small molecules The magnitude of the g-gradients and of the Maran-
such as non-ionic surfactants lower g more than poly- goni effect depends on the surface dilational modulus
meric surfactants such as poly(vinyl alcohol). ´, which for a plane interface with one surfactant-
Another important role of the surfactant is its effect containing phase, is given by the expression,
on the interfacial dilational modulus ´ w11–14x,
B dg E
dg ´sC y F Ž1 q 2zq2z2.1y2 (14)
´s (11) D dln G G
dln A
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 307

B dm E B DE
1y
2 Closely related to the above mechanism, is the Gibbs–
´sC FC F
C
(15) Marangoni effect. The depletion of surfactant in the thin
D dg G D 2v G
film between approaching drops results in g-gradient
without liquid flow being involved. This results in an
dln A
vs (16) inward flow of liquid that tends to drive the drops apart.
dt The Gibbs–Marangoni effect also explains the Bancroft
rule, which states that the phase in which the surfactant
where D is the diffusion coefficient of the surfactant is most soluble form the continuous phase. If the
and v represents a time scale (time needed for doubling surfactant is in the droplets, a g-gradient cannot develop
the surface area) that is roughly equal to tdef. and the drops would be prone to coalescence. Thus,
During emulsification, ´ is dominated by the magni- surfactants with HLB)7 tend to form OyW emulsions
tude of the denominator in Eq. (14) because z remains and HLB-7 tend to form WyO emulsions. The Gibbs–
small. The value of dmydG tends to go to very high Marangoni effect also explains the difference between
values when G reaches its plateau value. ´ goes to a surfactants and polymers for emulsification. Polymers
maximum when mC is increased. give larger drops when compared with surfactants. Pol-
For conditions that prevail during emulsification, ´ ymers give a smaller value of ´ at small concentrations
increases with mC and it is given by the relationship, when compared to surfactants.
Various other factors should also be considered for
dp emulsification, the most important is the disperse phase
´s (17)
dln G volume fraction w. Increase in w leads to increase in
droplet collision and hence coalescence during emulsi-
where p is the surface pressure (psgo yg). ´ is given fication. With increase in w, the viscosity of the emul-
by the slope of the line of plots of p vs. ln G. Such sion increases and could change the flow from being
plots show that a surfactant such as sodium dodecyl turbulent to being laminar (LV regime). The presence
sulfate (SDS) gives a much higher ´ value when of many particles results in a local increase in velocity
compared with a protein such as b-casein or lysozome gradients. This means that G increases. In turbulent
w15x. This is because the value of G is higher for SDS. flow, increase in w will induce turbulence depression.
The two proteins show difference in their ´ values, This will result in larger droplets. Turbulence depression
which may be attributed to the conformational change by added polymers tend to remove the small eddies,
that occur upon adsorption. resulting in the formation of larger droplets. If the mass
The presence of a surfactant means that during emul- ratio of surfactant to continuous phase is kept constant,
sification the interfacial tension need not to be the same increase in w results in decrease in surfactant concentra-
every where. This has two consequences: (i) the equi- tion and hence an increase in geq. This results in larger
librium shape of the drop is affected; (ii) any g-gradient droplets. If the mass ratio of surfactant to disperse phase
formed will slow down the motion of the liquid inside is kept constant, these changes are reversed.
the drop (this diminish the amount of energy needed to General conclusions cannot be drawn since several of
deform and break-up the drop). the above mentioned mechanism may come into play.
Another important role of the emulsifier is to prevent Experiments using a high pressure homogenizer at var-
coalescence during emulsification. This is certainly not ious w values at constant initial mC (regime TI changing
due to the strong repulsion between the droplets, since to TV at higher w showed that with increasing w ()
the pressure at which two drops are pressed together is 0.1) the resulting droplet diameter increased and the
much greater than the repulsive stresses. The counter- dependence on energy consumption became weaker. A
acting stress must be due to the formation of g-gradients. comparison of the average droplet diameter vs. power
When two drops are pushed together, liquid will flow consumption using different emulsifying machines,
out from the thin layer between them, and the flow will showed that the smallest droplet diameters were obtained
induce a g-gradient. This produces a counteracting stress when using the high pressure homogenizers.
given by,
4. Preparation of nano-emulsions
2 )Dg)
tDgf (18) Three methods may be applied for the preparation of
Ž1y2.d nano-emulsions (covering the droplet radius size range
50–200 nm). Use of high pressure homogenisers (aided
The factor 2 follows from the fact that two interfaces by appropriate choice of surfactants and cosurfactants),
are involved. Taking a value of Dgs10 mØNmy1, the use of low energy emulsification method at constant
stress amounts to 40 kPa (which is of the same order temperature or application of the phase inversion tem-
of magnitude as the external stress. perature (PIT) concept.
308 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Fig. 1. Schematic representation of the experimental path in two emul-


sification methods: Method A, addition of decane to waterysurfactant
mixture; method B, addition of water to decaneyBrij 30 solutions.
(Reproduced with permission from Ref. w15bx).

4.1. Use of high pressure homogenizers Fig. 2. Emulsion droplet diameters (circles) and rate constant for
attaining steady size (squares) as function of HLB-Cyclohex-
aneyNonylphenol Ethoxylate. (Reproduced with permission from Ref.
The production of small droplets (submicron) requires w15cx).
application of high energy. The process of emulsification
is generally inefficient as illustrated below. tant in the disperse phase rather than the continuous
Simple calculations show that the mechanical energy phase; this often leads to smaller droplets.
required for emulsification exceeds the interfacial energy It may be useful to emulsify in steps of increasing
by several orders of magnitude. For example, to produce intensity, particularly with emulsions having highly vis-
an emulsion at ws0.1 with a d32s0.6 mm, using a cous disperse phase.
surfactant that gives an interfacial tension gs10
mØNmy1, the net increase in surface free energy is Ags
4.2. Low energy emulsification methods
6wg yd32s104 Jmy3. The mechanical energy required
in a homogenizer is 107 Jmy3, i.e. an efficiency of
A study of the phase behaviour of wateryoilysurfac-
0.1%. The rest of the energy (99.9%) is dissipated as
tant systems demonstrated that emulsification can be
heat w10x.
achieved by three different low energy emulsification
The intensity of the process or the effectiveness in
methods w5x, as schematically shown in Fig. 1 (a)
making small droplets is often governed by the net
stepwise addition of oil to a water surfactant mixture;
power density (´(t)).
(b) stepwise addition of water to a solution of the
surfactant in oil; and (c) mixing all the components in
ps´Žt. dt (19) the final composition, pre-equilibrating the samples prior
to emulsification. In these studies, the system watery
where t is the time during which emulsification occurs. Brij 30 (polyoxyethlene lauryl ether with an average of
Break up of droplets will only occur at high ´ values, 4 mol of ethylene oxideydecane was chosen as a model
which means that the energy dissipated at low ´ levels to obtain OyW emulsions.
is wasted. Batch processes are generally less efficient The results showed that nano-emulsions with droplet
than continuous processes. This shows why with a stirrer sizes of the order of 50 nm were formed only when
in a large vessel, most of the energy applies at low water was added to mixtures of surfactant and oil
intensity is dissipated as heat. In a homogenizer, p is (method b).
simply equal to the homogenizer pressure.
Several procedures may be applied to enhance the 4.3. Phase inversion temperature (PIT) principle
efficiency of emulsification when producing nano-emul-
sions: One should optimise the efficiency of agitation Phase inversion in emulsions can be one of two types:
by increasing ´ and decreasing dissipation time. The Transitional inversion induced by changing factors
emulsion is preferably prepared at high volume faction which affect the HLB of the system, e.g. temperature
of the disperse phase and diluted afterwards. However, andyor electrolyte concentration. Catastrophic Inversion,
very high w values may result in coalescence during which is induced by increasing the volume fraction of
emulsification. Add more surfactant, whereby creating a the disperse phase.
smaller geff and possibly diminishing recoalescence. Use Transitional Inversion can also be induced by chang-
surfactant mixture that shows more reduction in g the ing the HLB number of the surfactant at constant
individual components. If possible dissolve the surfac- temperature using surfactant mixtures. This is illustrated
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 309

Fig. 3. The PIT concept. (Reproduced with permission from Ref. w15dx).

in Fig. 2, which shows the average droplet diameter and temperature. This is illustrated in Fig. 3, which shows
rate constant for attaining constant droplet size as a schematically what happens when the temperature is
function of the HLB number. increased w18,19x.
It can be seen that the diameter decreases and the At low temperature, over the Winsor I region, OyW
rate constant increases as inversion is approached. For macroemulsions can be formed and are quite stable. On
application of the phase inversion principle one uses the increasing the temperature, the OyW emulsion stability
transitional inversion method which has been demon- decreases and the macroemulsion finally resolves when
strated by Shinoda and coworkers w16,17x when using the system reaches the Winsor III phase region (both
non-ionic surfactants of the ethoxylate type. These OyW and WyO emulsions are unstable). At higher
surfactants are highly dependent on temperature, becom- temperature, over the Winsor II region, WyO emulsions
ing lipophilic with increasing temperature due to the become stable.
dehydration of the polyethyleneoxide chain. When an Fig. 4 shows the most clear-cut image of the macroe-
OyW emulsion is prepared using a non-ionic surfactant mulsion inversion as a function of temperature. Equal
of the ethoxylate type is heated, then at a critical volumes of oil and water are emulsified at various
temperature (the PIT), the emulsion inverts to a WyO temperatures. Five hours after preparation, macroemul-
emulsion. At the PIT the droplet size reaches a minimum sions completely sediment. Below the balanced temper-
and the interfacial tension also reaches a minimum. ature (HLB temperature), a stable OyW macroemulsion
However, the small droplets are unstable and they is formed, whereas above the balanced temperature a
coalesce very rapidly. By rapid cooling of the emulsion stable WyO emulsion is formed.
that is prepared at a temperature near the PIT, very Close to the balanced point (60–68 8C), a three phase
stable and small emulsion droplets could be produced. equilibrium is observed and neither OyW or WyO
A clear demonstration of the phase inversion that emulsions are stable.
occurs on heating an emulsion is illustrated from a study Near the HLB temperature, the interfacial tension
of the phase behaviour of emulsions as a function of reaches a minimum. This is illustrated in Fig. 5. Thus
310 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

consider the interaction forces between droplets contain-


ing adsorbed layers (Steric stabilization). This was
described in detail in several reviews and textbooks only
a summary is given here w20,21x.
When two droplets each containing an adsorbed layer
of thickness d approach to a distance of separation h,
whereby h becomes -2d, repulsion occurs as a result
of two main effects: (i) Unfavourable mixing of the
stabilizing chains A of the adsorbed layers, when these
are in good solvent conditions. This is referred to as the
mixing (osmotic interaction, Gmix and is given by the
following expression,

Gmix 4p 2 B 1 E B hE
Fig. 4. Macroemulsion stability diagram of cyclohexane-water- s f2 C yxF C3aq2dq F (20)
kT 3V1 D2 G D 2G
polyoxyethylene (9.7) nonyl phenol ether system. (Reproduced with
permission from Ref. w16x).
where k is the Boltzmann constant, T is the absolute
temperature, V1 is the molar volume of the solvent, w2
by preparing the emulsion at a temperature 2–4 8C
is the volume fraction of the polymer (the A chains) in
below the PIT (near the minimum in g) followed by
the adsorbed layer and x is the Flory–Huggins (poly-
rapid cooling of the system, nano-emulsions may be
mer–solvent interaction) parameter.
produced.
It can be seen that Gmix depends on three main
The minimum in g can be explained in terms of the
parameters: The volume fraction of the A chains in the
change in curvature H of the interfacial region, as the
adsorbed layer (the more dense the layer is the higher
system changes from OyW to WyO.
the value of Gmix). The Flory–Huggins interaction
For OyW system and normal micelles, the monolayer
parameter x (for Gmix to remain positive, i.e. repulsive,
curves towards the oil and H is given a positive value.
x should be lower than 1y2). The adsorbed layer
For a WyO emulsions and inverse micelles, the mono-
thickness d .
layer curves towards the water and H is assigned a
(ii) Reduction in configurational entropy of the chains
negative value. At the inversion point (HLB tempera-
on significant overlap. This referred to as elastic (entrop-
ture) H becomes zero and g reaches a minimum.
ic) interaction and is given by the expression,
5. Steric stabilization and the role of the adsorbed w z
VŽh.
Gels2n2lnx |
layer thickness
(21)
y VŽ`. ~
Since most nano-emulsions are prepared using non-
ionic andyor polymeric surfactants, it is necessary to where n2 is the number of chains per unit area, V(h) is

Fig. 5. Interfacial tensions of n-octane against water in the presence of various CnEm surfactants above the cmc as a function of temperature.
(Reproduced with permission from Ref. w19x).
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 311

6. Ostwald ripening

One of the main problems with nano-emulsions is


Ostwald ripening which results from the difference in
solubility between small and large droplets. The differ-
ence in chemical potential of dispersed phase droplets
between different sized droplets as given by Lord Kelvin
w22x,

B 2gV E
cŽr.scŽ`. expC F
m
(23)
D rRT G

where c(r) is the solubility surrounding a particle of


radius r, c(`) is the bulk phase solubility and Vm is
the molar volume of the dispersed phase.
The quantity (2gVm yrRT) is termed the characteristic
length. It has an order of ;1 nm or less, indicating that
Fig. 6. Variation of Gmix, Gel, GA and GT with h. the difference in solubility of a 1 mm droplet is of the
order of 0.1% or less.
the configurational entropy of the chains at a separation Theoretically, Ostwald ripening should lead to con-
distance h and V(`) is the configurational entropy at densation of all droplets into a single drop (i.e. phase
infinite distance of separation. separation). This does not occur in practice since the
Combination of Gmix, Gel with the van der Waals rate of growth decreases with increase of droplet size.
attraction GA gives the total energy of interaction GT, For two droplets of radii r1 and r2 (where r1-r2),

GTsGmixqGelqGA (22) w
B RT E cŽr1. z B 1E
C
D Vm G
F lnx | s 2g CD r1
cŽr2. ~
y F
r2 G
(24)
y
Fig. 6 gives a schematic representation of the variation 1

of Gmix, Gel, GA and GT with h. As can be seen form


Fig. 5, Gmix increases very rapidly with decrease of h Eq. (24) shows that the larger the difference between
as soon as h-2d, Gel increase very rapidly with decrease r1 and r2, the higher the rate of Ostwald ripening.
of h when h-d. GT shows one minimum, Gmin, and it Ostwald ripening can be quantitatively assessed from
increases very rapidly with decrease of h when h-2d. plots of the cube of the radius vs. time t (the Lifshitz–
The magnitude of Gmin depends on the following Slesov–Wagner, LSW, Theory) w23,24x,
parameters: The particle radius R. The Hamaker constant
A. The adsorbed layer thickness d. w z
8 cŽ`.gVmD
As an illustration, Fig. 7 shows the variation of GT rs x
3 |t (25)
with h at various ratios of d yR. It can be seen from Fig. 9 y rRT ~

7 that the depth of the minimum decreases with increas-


ing d yR. This is the basis of the high kinetic stability
of nano-emulsions. With nano-emulsions having a radius
in the region of 50 nm and an adsorbed layer thickness
of say 10 nm, the value of d yR is 0.2. This high value
(when compared with the situation with macroemulsions
where d yR is at least an order of magnitude lower)
results in a very shallow minimum (which could be less
than kT).
The above situation results in very high stability with
no flocculation (weak or strong). In addition, the very
small size of the droplets and the dense adsorbed layers
ensures lack of deformation of the interface, lack of
thinning and disruption of the liquid film between the
droplets and hence coalescence is also prevented.
The only instability problem with nano-emulsions is
Ostwald ripening which is discussed below. Fig. 7. Variation of GT with h with increasing dyR.
312 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Fig. 8. Pseudoternary phase diagram at 258 of the system water-C12EO4-hexadecane. (Reproduced with permission from Ref. w29bx).

where D is the diffusion coefficient of the disperse not desorb during ripening, the rate could be signifi-
phase in the continuous phase and r is the density of cantly reduced. An increase in the surface dilational
the disperse phase. modulus and decrease in g would be observed for the
Several methods may be applied to reduce Ostwald shrinking drops. The difference in g between the drop-
ripening w25–27x: lets would balance the difference in capillary pressure
(i) Addition of a second disperse phase component, (i.e. curvature effects).
which is insoluble in the continuous phase (e.g. squa- To achieve the above effect it is useful to use A-B-A
lene). In this case, significant partitioning between block copolymers that are soluble in the oil phase and
different droplets occurs, with the component having insoluble in the continuous phase. The polymeric sur-
low solubility in the continuous phase expected to be factant should enhance the lowering of g by the emul-
concentrated in the smaller droplets. During Ostwald sifier. In other words, the emulsifier and the polymeric
ripening in two component disperse phase system, equi- surfactant should show synergy in lowering g.
librium is established when the difference in chemical
potential between different size droplets (which results 7. Practical examples of nano-emulsions
from curvature effects) is balanced by the difference in
chemical potential resulting from partioning of the two Several experiments were recently carried to investi-
components. If the secondary component has zero sol- gate the methods of preparation of nano-emulsions and
ubility in the continuous phase, the size distribution will their stability w29x. The first method applied the PIT
not deviate from the initial one (the growth rate is equal principle for preparation of nano-emulsions. Experi-
to zero). In the case of limited solubility of the second- ments were carried out using hexadecane and isohex-
ary component, the distribution is the same as governed adecane (Arlamol HD) as the oil phase and Brij 30
by Eq. (16), i.e. a mixture growth rate is obtained which (C12EO4) as the non-ionic emulsifier. The phase dia-
is still lower than that of the more soluble component. grams of the ternary system water-C12EO4-hexadecane
The above method is of limited application since one and water-C12EO4-isohexadecane are shown in Figs. 8
requires a highly insoluble oil as the second phase, and 9. The main features of the pseudoternary system
which is miscible with the primary phase. are as follows: (i) Om isotropic liquid transparent phase,
(ii) Modification of the interfacial film at the OyW which extends along the hexadecane-C12EO4 or isohex-
interface: According to Eq. (24) reduction in g results adecane-C12EO4 axis, corresponding to inverse micelles
in reduction of Ostwald ripening. However, this alone or WyO microemulsions; (ii) La lamellar liquid crystal-
is not sufficient since one has to reduce g by several line phase extending from the W-C12EO4 axis toward
orders of magnitude. Walstra w28x suggested that by the oil vertex; (iii) The rest of the phase diagram
using surfactants, which are strongly adsorbed at the Oy consists of two- or three-phase regions: (WmqO) two-
W interface (i.e. polymeric surfactants) and which do liquid-phase region, which appears along the water–oil
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 313

Fig. 9. Pseudoternary phase diagram at 258 of the system water-C12EO4-isohexadecane. (Reproduced with permission from Ref. w29ax).

axis; (WmqLaqO) three-phase region, which consists polydispersity index decreases with increase in surfac-
of a bluish liquid phase (OyW microemulsion), a tant concentration. The decrease in droplet size with
lamellar liquid crystalline phase (La) and a transparent increase in surfactant concentration is due to the increase
oil phase; (LaqOm) two-phase region consisting of an in surfactant interfacial area and decrease in interfacial
oil and liquid crystalline region; MLC a multiphase tension, g. As mentioned above, g reaches a minimum
region containing a lamellar liquid crystalline phase at the HLB temperature. Therefore, the minimum in
(La). interfacial tension occurs at lower temperature as the
The HLB temperature was determined using conduc- surfactant concentration increases. This temperature
tivity measurements, whereby 10y2 mol dmy3 NaCl becomes closer to the cooling temperature as the surfac-
was added to the aqueous phase (to increase the sensi- tant concentration increases and this results in smaller
tivity of the conductibility measurements). The concen- droplet sizes.
tration of NaCl was low and hence it had little effect All nano-emulsions showed an increase in droplet
on the phase behaviour. size with time, as a result of Ostwald ripening. Fig. 11
Fig. 10 shows the variation of conductivity vs. tem- shows plots of r 3 vs. time for all the nano-emulsions
perature for 20% OyW emulsions at different surfactant studied. The slope of the lines gives the rate of Ostwald
concentrations. It can be seen that there is a sharp ripening v (m3 sy1) and this showed an increase from
decrease in conductivity at the PIT or HLB temperature 2=10y27 to 39.7=10y27 m3 sy1 as the surfactant
of the system. concentration is increased from 4 to 8 wt.%. This
The HLB temperature decreases with increase in increase could be due to a number of factors: (i)
surfactant concentration. This could be due to the excess Decrease in droplet size increases the Brownian diffu-
non-ionic surfactant remaining in the continuous phase. sion and this enhances the rate; (ii) Presence of micelles,
However, at a concentration of surfactant higher than which increases with increase in surfactant concentra-
5%, the conductivity plots showed a second maximum tion. This has the effect of increasing the solublisation
(Fig. 10). This was attributed to the presence of La of the oil into the core of the micelles. This results in
phase and bicontinuous L3 or D9 phases w30x. an increase of the flux J of diffusion of oil molecules
Nano-emulsions were prepared by rapid cooling of from different size droplets. Although the diffusion of
the system to 25 8C. The droplet diameter was deter- micelles is slower than the diffusion of oil molecules,
mined using photon correlation spectroscopy (PCS). the concentration gradient (dCy dX) can be increased
The results are summarised in Table 1, which shows the by orders of magnitude as a result of solubilisation. The
exact composition of the emulsions, HLB temperature, overall effect will be an increase in J and this may
z-average radius and polydispersity index. enhance Ostwald ripening; (iii) Partition of surfactant
OyW nano-emulsions with droplet radii in the range molecules between the oil and aqueous phases. With
26–66 nm could be obtained at surfactant concentrations higher surfactant concentrations, the molecules with
between 4 and 8%. The nano-emulsion droplet size and shorter EO chains (lower HLB number) may preferen-
314 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Fig. 10. Conductivity vs. temperature for a 20:80 hexadecane: water emulsions at various C12EO4 concentrations. (Reproduced with permission
from Ref. w29cx).

Table 1 Table 2
Composition, HLB temperature (THLB ), droplet radius r and polydis- Composition, HLB temperature (THLB ), droplet radius r and polydis-
persity index (pol.) for the system water-C12 EO4 -hexadecane at 25 8C persity index (pol.) at 258 for emulsions in the system water-
C12EO4-isohexadecane
Surfactant Water OilyWater THLB(8C) rynm pol.
(wt.%) (wt.%) Surfactant Water OyW THLBy oC rynm pol.
2.0 78.0 20.4y79.6 – 320 1.00 (wt.%) (wt.%)
3.0 77.0 20.6y79.4 57.0 82 0.41
3.5 76.5 20.7y79.3 54.0 69 0.30 2.0 78.0 20.4y79.6 – 97 0.50
4.0 76.0 20.8y79.2 49.0 66 0.17 3.0 77.0 20.6y79.4 51.3 80 0.13
5.0 75.0 21.2y78.9 46.8 48 0.09 4.0 76.0 20.8y79.2 43.0 65 0.06
6.0 74.0 21.3y78.7 45.6 34 0.12 5.0 75.0 21.1y78.9 38.8 43 0.07
7.0 73.0 21.5y78.5 40.9 30 0.07 6.0 74.0 21.3y78.7 36.7 33 0.05
8.0 72.0 21.7y78.3 40.8 26 0.08 7.0 73.0 21.3y78.7 33.4 29 0.06
8.0 72.0 21.7y78.3 32.7 27 0.12

tially accumulate at the OyW interface and this may


result in reduction of the Gibbs elasticity, which in turn As with the hexadecane system, the droplet size and
results in an increase in the Ostwald ripening rate. polydispersity index decreased with increase in surfac-
The results with isohexadecane are summarised in tant concentration. Nano-emulsions with droplet radii of
Table 2. 25–80 nm were obtained at 3–8% surfactant concentra-

Fig. 11. r 3 vs. time at 25 8C for nano-emulsions prepared using the system water-C12 EO4 -hexadecane. (Reproduced with permission from Ref.
w29cx).
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 315

tion. It should be noted, however, that nano-emulsions


could be produced at lower surfactant concentration
when using isohexadecane, when compared with the
results obtained with hexadecane. This could be attrib-
uted to the higher solubility of the isohexadecane (a
branched hydrocarbon), the lower HLB temperature and
the lower interfacial tension.
The stability of the nano-emulsions prepared using
isohexadecane was assessed by following the droplet
size as a function of time. Plots of r 3 vs. time for four
surfactant concentrations (3, 4, 5 and 6 wt.%) are shown
in Fig. 12. The results show an increase in Ostwald
ripening rate as the surfactant concentration is increased
from 3 to 6% (the rate increased from 4.1=10y27 to
50.7=10y27 m3 sy1). The nano-emulsions prepared
using 7 wt.% surfactant were so unstable that they Fig. 13. r 3 vs. time at 25 8C for nano-emulsions (OyW ratio 20y80)
showed significant creaming after 8 h. However, when with hydrocarbons of various alkyl chain lengths. System water-
C12EO4-hydrocarbon (4 wt.% surfactant). (Reproduced with permis-
the surfactant concentration was increased to 8 wt.%, a
sion from Ref. w29cx).
very stable nano-emulsion could be produced with no
apparent increase in droplet size over several months.
This unexpected stability was attributed to the phase not sufficient to produce the nano-emulsion droplets
behaviour at such surfactant concentrations. The sample with high surface area. Similar results were obtained
containing 8 wt.% surfactant showed birifringence to with isohexadecane. However, nano-emulsions could be
shear when observed under polarised light. It seems that produced using 30y70 OyW ratio (droplet size being 81
the ratio between the phases (WmqLaqO) may play a nm), but with high polydispersity index (0.28). The
key factor in nano-emulsion stability. nano-emulsions showed significant Ostwald ripening.
Attempts were made to prepare nano-emulsions at The effect of changing the alkyl chain length and
higher OyW ratios (hexadecane being the oil phase), branching was investigated using decane, dodecane,
while keeping the surfactant concentration constant at 4 tetradecane, hexadecane and isohexadecane. Plots of r 3
wt.%. When the oil content was increased from 40% to vs. time are shown in Fig. 13 for 20y80 OyW ratio and
50%, the droplet radius increased from 188 nm to 297 surfactant concentration of 4 wt.%. As expected by
nm, respectively. In addition, the polydispersity index reducing the oil solubility from decane to hexadecane,
also increased to 0.95. These systems become so unsta- the rate of Ostwald ripening decreases. The branched
ble that they showed creaming within few hours. This oil isohexadecane also shows higher Ostwald ripening
is not surprising, since the surfactant concentration is rate when compared with hexadecane. A summary of

Fig. 12. r 3 vs. time at 25 8C for the system water-C12 EO4 -isohexadecane at various surfactant concentration; OyW ratio 20y80. (Reproduced with
permission from w29ax).
316 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Table 3 tions of a less soluble oil, namely squalene, was added.


HLB temperature (THLB), droplet radius r, Ostwald ripening rate v The results using hexadecane did significant decrease in
and oil solubility for nanao-emulsions prepared using hydrocarbons
with different alkyl chain length stability on addition of 10% squalane. This was thought
to be due to coalescence rather than increase in Ostwald
Oil THLBy8C rynm vØ1027 m3 sy1 C(`) mlØmly1 ripening rate. In some cases, addition of a hydrocarbon
Decane 38.5 59 20.9 710.0 with a long alkyl chain can induce instability as a result
Dodecane 45.5 62 9.3 52.0 of change in the adsorption and conformation of the
Tetradecane 49.5 64 4.0 3.7 surfactant at the OyW interface.
Heaxadecane 49.8 66 2.3 0.3 In contrast to the results obtained with hexadecane,
Isohexadecane 43.0 60 8.0 –
addition of squalene to the OyW nano-emulsion system
based on isohexadecane showed a systematic decrease
the results is shown in Table 3 which also shows the in Ostwald ripening rate as the squalene content was
solubility of the oil C(`). increased. The results are shown in Fig. 14, which
As expected from the Ostwald ripening theory (LSW shows plots of r 3 vs. time for nano-emulsions containing
theory, Eq. (25), the rate of Ostwald ripening decreases varying amounts of squalene. Addition of squalene up
as the oil solubility decreases. Isohexadecan has a rate to 20% based on the oil phase showed a systematic
of Ostwald ripening similar to that of dodecane. reduction in the rate (from 8.0=1027 to 4.1=1027 m3
As discussed before, one would expect that the sy1). It should be noted that when squalene alone was
Ostwald ripening of any given oil should decrease on used as the oil phase, the system was very unstable and
incorporation of a second oil with much lower solubility. it showed creaming within 1 h. This shows that the
To test this hypothesis, nano-emulsions were made using surfactant used is not suitable for emulsification of
hexadecane or isohexadecane to which various propor- squalane.

Fig. 14. r 3 vs. time at 25 8C for the system water-C12EO4-isohexadecane-squalane (20y80 OyW and 4 wt.% surfactant). (Reproduced with
permission from Ref. w29ax).

Fig. 15. r vs. HLB number at two different surfactant concentrations (OyW ratio 20y80). (Reproduced with permission from Ref. w29ax).
T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318 317

polymeric surfactant that adsorbs strongly at the OyW


interface, one would expect reduction in the Ostwald
ripening rate. To test this hypothesis, an A-B-A block
copolymer of polyhroxystearic acid (PHS, the A chains)
and polyethylene oxide (PEO, the B chain) PHS-PEO-
PHS (Arlacel P135) was incorporated in the oil phase
at low concentrations (the ratio of surfactant to Arlacel
was varied between 99:1 and 92:8). For the hexadecane
system, the Ostwald ripening rate showed a decrease
with the addition of Arlacel P135 surfactant at ratios
lower than 94:6. Similar results were obtained using
isohexadecane. However, at higher polymeric surfactant
concentrations, the nano-emulsion became unstable.
As mentioned above, the nano-emulsions prepared
using the PIT method are relatively polydisperse and
Fig. 16. v vs. HLB number in the systems water-C12EO4-C12 EO6 - they generally give higher Ostwald ripening rates when
isohexadecane at two surfactant concentrations. (Reproduced with compared to nano-emulsions prepared using high pres-
permission from Ref. w29ax). sure homogenisation techniques. To test this hypothesis,
several nano-emulsions were prepared using a Micro-
The effect of HLB number on nano-emulsion forma- fluidiser (that can apply pressures in the range 5000–
tion and stability was investigated by using mixtures of 15 000 psi or 350–1000 bar). Using an oil:surfactant
C12EO4 (HLBs9.7) and C12EO4 (HLBs11.7) (Fig. ratio of 4:8 and OyW ratios of 20:80 and 50:50,
15). Two surfactant concentrations (4 and 8 wt.%) were emulsions were prepared first using the Ultturrax fol-
used and the OyW ratio was kept at 20y80. Fig. 15 lowed by high pressure homogenisation (ranging from
shows the variation of droplet radius with HLB number. 1500 to 15 000 psi). The best results were obtained
This figure shows that the droplet radius remain virtually using a pressure of 15 000 psi (one cycle of homogen-
constant in the HLB range 9.7–11.0, after which there isation). The droplet radius was plotted vs. the
is a gradual increase in droplet radius with increase in oil:surfactant ratio, R(OyS) as shown in Fig. 17.
the HLB number of the surfactant mixture. All nano- For comparison, the theoretical radii values calculated
emulsions showed an increase in droplet radius with by assuming that all surfactant molecules are at the
time, except for the sample prepared at 8 wt.% surfactant interface was calculated using Nakajima equation w1,2x,
with an HLB number of 9.7 (100% C12EO4). Fig. 16
B 3Mb E B 3aM E
shows the variation of Ostwald ripening rate constant v
rsC FRqC Fqd
b
(26)
with HLB number of surfactant. The rate seems to D ANra G D ANrb G
decrease with increase of surfactant HLB number and
when the latter is )10.5, the rate reaches a low value where Mb is the molecular weight of the surfactant, A
(-4=10y27 m3 sy1). is the area occupied by a single molecule, N is the
As discussed above, the incorporation of an oil soluble Avogadro’s number, ra is the oil density, rb is the

Fig. 17. r vs. R(OyS) at 25 8C for the system water-C12EO4-hexadecane. Wm smicellar solution or OyW microemulsion, Laslamellar liquid
crystalline phase; Osoil phase. (Reproduced with permission from Ref. w29ax).
318 T. Tadros et al. / Advances in Colloid and Interface Science 108 – 109 (2004) 303–318

Fig. 18. r 3 vs. time for nano-emulsion systems prepared using the PIT and Microfluidiser. 20:80 OyW and 4 wt.% surfactant. (Reproduced with
permission from Ref. w31x).

density of the surfactant alkyl chain, a is the alkyl chain w13x J. Lucassen, in: E.H. Lucassen-Reynders (Ed.), Anionic Sur-
weight fraction and d is the thickness of the hydrated factants, Marcel Dekker, NY, 1981.
w14x M. van Den Tempel, Proc. Int. Congr. Surf. Act. 2 (1960) 573.
layer of PEO. w15x (a) D.E. Graham, M.C. Phillips, J. Colloid Interface Sci. 70
In all cases, there is an increase in nano-emulsion (1979) 415
radius with increase in the R(OyS). However, when (b) A. Forgiarini, J. Esquena, C. Gonzalez,
´ C. Solans, Lang-
using the high pressure homogeniser, the droplet size muir 17 (2001) 2076
can be maintained to values below 100 nm at high (c) B.W. Brooks, H.N. Richmond, Chem. Eng. Sci. 49 (1994)
1053
R(OyS) values. With the PIT method, there is a rapid (d) A.S. Kabalnov, in: B.P. Binks (Ed.), Modern Aspects of
increase in r with increase in R(OyS) when the latter Emulsion Science, Royal Society of Chemistry, Cambridge,
exceeds 7. 1998.
As expected, the nano-emulsions prepared using high w16x K. Shinoda, H. Saito, J. Colloid Interface Sci. 30 (1969) 258.
pressure homogenisation showed a lower Ostwald rip- w17x K. Shinoda, H. Saito, J. Colloid Interface Sci. 26 (1968) 70.
w18x B.W. Brooks, H.N. Richmond, M. Zerfa, in: B.P. Binks (Ed.),
ening rate when compared to the systems prepared using
Modern Aspects of Emulsion Science, Royal Society of Chem-
the PIT method. This is illustrated in Fig. 18, which istry Publication, Cambridge, 1998.
shows plots of r 3 vs. time for the two systems. w19x T. Sottman, R. Strey, J. Chem. Phys. 108 (1997) 8606.
w20x D.H. Napper, Polymeric Stabilisation of Colloidal Dispersions,
References Academic Press, London, 1983.
w21x Th.F. Tadros, Polymer adsorption and colloid stability, in: Th.F.
w1x H. Nakajima, S. Tomomossa, M. Okabe, First Emulsion Con- Tadros (Ed.), The Effect of Polymers on Dispersion Properties,
ference, Paris (1993). Academic Press, London, 1982.
w2x H. Nakajima, in: C. Solans, H. Konieda (Eds.), Industrial w22x W. Thompson (lord kelvin), Phil. Mag., 42 (1871) 448.
Applications of Microemulsions, Marcel Dekker, 1997. w23x I.M. Lifshitz, V.V. Slesov, Sov. Phys. JETP 35 (1959) 331.
w3x J. Ugelstadt, M.S. El-Aassar, J.W. Vanderhoff, J. Polym. Sci. w24x C. Wagner, Z. Electrochem. 35 (1961) 581.
11 (1973) 503. w25x A.S. Kabalnov, E.D. Shchukin, Adv. Colloid Interface Sci. 38
w4x M. El-Aasser, in: J.M. Asua (Ed.), Polymeric Dispersions, (1992) 69.
Kluwer Academic Publications, The Netherlands, 1997. w26x A.S. Kabalnov, Langmuir 10 (1994) 680.
w5x A. Forgiarini, J. Esquena, J. Gonzalez, C. Solans, Prog. Colloid w27x J.G. Weers, in: B.P. Binks (Ed.), Modern Aspects of Emulsion
Polym. Sci. 115 (2000) 36. Science, Royal Society of Chemistry Publication, Cambridge,
w6x K. Shinoda, H. Kunieda, in: P. Becher (Ed.), Encyclopedia of 1998.
Emulsion Technology, Marcel Dekker, NY, 1983. w28x P. Walstra, Chem. Eng. Sci. 48 (1993) 333.
w7x S. Benita, M.Y. Levy, J. Pharm. Sci. 82 (1993) 1069. w29x (a) P. Izquierdo, Studies on Nano-Emulsion Formation and
w8x P. Walstra, in: P. Becher (Ed.), Encyclopedia of Emulsion Stability, Thesis, University of Barcelona, Spain (2002)
Technology, Marcel Dekker, NY, 1983. (b) A. Forgiarini, Estudio de la Relacion
´ entre comportamiento
w9x P.N. Pusey, in: J.H.S. Green, R. Dietz (Eds.), Industrial ´
fasico y formacion´ de nano-emulsiones de fase externa acusosa,
Polymers: Characterisation by Molecular Weights, Transcripta University of Barcelona, Spain (2001)
Books, London, 1973. (c) P. Izquierdo, J. Esquena, Th.F. Tadros, C. Dederen, M.J.
w10x P. Walstra, P.E.A. Smoulders, in: B.P. Binks (Ed.), Modern Garcia, N. Azemar, C. Solans, Langmuir 18 (2002) 26.
Aspects of Emulsion Science, The Royal Society of Chemistry, w30x H. Kuneida, Y. Fukuhi, H. Uchiyama, C. Solans, Langmuir 12
Cambridge, 1998. (1996) 2136.
w11x E.H. Lucassen-Reynders, in: P. Becher (Ed.), Encyclopedia of w31x P. Izquierdo, J. Esquena, Th.F. Tadros, C. Dederen, M.J. Garcia,
Emulsion Technology, Marcel Dekker, NY, 1996. N. Azemar, C. Solans, XXXI Jornadas del comite´ Espanol ˜ de
w12x E.H. Lucassen-Reynders, Colloid Surf. A 91 (1994) 79. la Detergencia, Barcelona, Spain (2001) 359.

You might also like