Вы находитесь на странице: 1из 8

6092

Langmuir 2008, 24, 6092-6099

Nanoemulsions Prepared by a Two-Step Low-Energy Process


Lijuan Wang, Kevin J. Mutch, Julian Eastoe, Richard K. Heenan, and Jinfeng Dong*,
College of Chemistry and Molecular Science, Wuhan UniVersity, Wuhan 430072, China, School of
Chemistry, UniVersity of Bristol, Bristol BS8 1TS, U.K., and ISIS-STFC, Rutherford Appleton Laboratory,
Chilton, Oxon OX11 0QX, U.K.
ReceiVed February 27, 2008. ReVised Manuscript ReceiVed March 20, 2008
A simple low-energy two-step dilution process has been applied in oil/surfactant/water systems with pentaoxyethylene
lauryl ether (C12E5), dodecyldimethylammonium bromide, sodium bis(2-ethylhexyl)sulfosuccinate, sodium n-dodecyl
sulfate-pentanol, and hexadecyltrimethylammonium bromide-pentanol. Appropriate formulations were chosen for
the concentrate to be diluted with water to generate oil-in-water (O/W) emulsions or nanoemulsions. For the system
of decane/C12E5/water, bluish, transparent nanoemulsions having droplet radii of the order of 15 nm were formed,
only when the initial concentrate was a bicontinuous microemulsion, whereas opaque emulsions were generated if
the concentrate began in an emulsion-phase region. Nanoemulsions generated in the system decane/C12E5/water have
been investigated both by dynamic light scattering (DLS) and contrast-variation small-angle neutron scattering (SANS).
The SANS profiles show that nanodroplets exist as spherical core-shell (decane-C12E5) particles, which suffer
essentially no structural change on dilution with water, at least for volume fractions down to 0.060. These results
suggest that the nanoemulsion droplet structure is mainly controlled by the phase behavior of the initial concentrate
and is largely independent of dilution. A discrepancy between apparent nanoemulsion droplet sizes was observed by
comparing DLS and SANS data, which is consistent with long-range droplet interactions occurring outside of the
SANS sensitivity range. These combined phase behavior, SANS, and DLS results suggest a different reason for the
stability/instability of nanoemulsions compared with earlier studies, and here it is proposed that a general mechanism
for nanoemulsion formation is homogeneous nucleation of oil droplets during the emulsification.

1. Introduction
Nanoemulsions1 consist of small quite monodisperse droplets,
typically in the 20-200 nm size range. Although they are of
similar size to microemulsion droplets (1-100 nm), and appear
transparent or translucent, they are in fact distinctly quite different
from true microemulsions. Nanoemulsions are thermodynamically unstable, and their formation generally requires energy
input. The properties of nanoemulsions depend not only on
thermodynamic conditions (i.e., composition, temperature, and
pressure) but also on the preparation method, and, crucially, on
the order of component addition.2 However, the kinetic stability
of nanoemulsions, and their transparent or translucent appearance
due to the presence of nanometer-sized droplets, makes nanoemulsions of interest for fundamental studies and practical
applications (e.g., in chemical, pharmaceutical,3 and cosmetic4
fields). To obtain nanodroplet emulsions, significant amounts of
mechanical energy are needed, making high-energy preparation
methods unfavorable for industrial applications.5,6 Therefore,
the preparation of nanoemulsions with reproducible properties
* To whom correspondence should be addressed. E-mail: colloid@
whu.edu.cn.

Wuhan University.

University of Bristol.

Rutherford Appleton Laboratory.


(1) Forgiarini, A.; Esquena, J.; Gonzalez, C.; Solans, C. Langmuir 2002, 17,
20762083.
(2) Esquena, J.; Solans, C. Prog. Colloid Polym. Sci. 1998, 110, 235239.
(3) Nicolaos, G.; Crauste-Manciet, S.; Farinotti, R.; Brossard, D. Int. J. Pharm.
2003, 263, 165171.
(4) Sonneville-Aubrun, O.; Simonnet, J.-T.; LAlloret, F AdV. Colloid Interface
Sci. 2004, 108-109, 145149.
(5) Liedtke, S.; Wissing, S.; Muller, R. H.; Mader, K. Int. J. Pharm. 2000, 196,
183185.
(6) Shi, R.; Hong, L.; Wu, D.; Ning, X.; Chen, Y.; Lin, T.; Fan, D.; Wu, K.
Cancer Biol. Ther. 2005, 4, 218224.
(7) Solans, C.; Izquierdo, P.; Nolla, J.; Azemar, N.; Garcia-Celma, M. J. Curr.
Opin. Colloid Interface Sci. 2005, 10, 102110.
(8) Pons, R.; Carrera, I.; Caelles, J.; Rouch, J; Panizza, P. AdV. Colloid Interface
Sci. 2003, 106, 129146.

and small droplet sizes using low-energy methods has been a


field of growing interest.7,8
Low-energy methods7 make use of accessible phase transitions
occurring during the emulsification process as a result of changes
in surfactant film spontaneous curvature. This curvature transition
has been achieved through different routes: (a) partitioning of
alcohol from the oil to the aqueous phase or diffusion of water
into the initial droplet, both producing a shift from lipophilic to
hydrophilic conditions;9,10 (b) chemical reactions which convert
lipophilic surfactants to hydrophilic surfactants;11,12 (c) a sudden
decrease of ionic strength with ionic surfactant systems;13,14 and
(d) an increased hydration of poly(oxyethylene) chains of PEOtype nonionic surfactants.1522 The reversal of spontaneous
curvature from water-in-oil (W/O) to an oil-in-water configuration
(O/W) reduces the solubilization capacity for oil to such an extent
that supersaturation may occur, leading to oil droplet nucleation.
(9) Rang, M. J.; Miller, C. A. J. Colloid Interface Sci. 1999, 209, 179192.
(10) Rang, M. J.; Miller, C. A. Prog. Colloid Polym. Sci. 1998, 109, 101117.
(11) Nishimi, T.; Miller, C. A. J. Colloid Interface Sci. 2001, 237, 259266.
(12) Sole`, I.; Maestro, A.; Pey, C. M.; Gonzalez, C.; Solans, C.; Gutierrez,
J. M. Colloids Surf., A 2006, 288, 138143.
(13) Nishimi, T.; Miller, C. A. Langmuir 2000, 16, 92339241.
(14) (a) Bataller, H.; Lamaallam, S.; Lachaise, J.; Graciaa, A.; Dicharry, C.
J. Mater. Process Tech. 2004, 152, 215220. (b) Lamaallam, S.; Bataller, H.;
Dicharry, C.; Lachaise, J. Colloids Surf., A 2005, 270-271, 4451.
(15) Leaver, M. S.; Olsson, U.; Wennerstrom, H. J. Chem. Soc., Faraday
Trans. 1995, 91, 42694274.
(16) Izquierdo, P.; Esquena, J.; Tadros, Th. F.; Dederen, C; Garcia, M. J.;
Azemar, N.; Solans, C. Langmuir 2002, 18, 2630.
(17) Morales, D.; Gutierrez, J. M.; Garca-Celma, M. J.; Solans, C. Langmuir
2003, 19, 71967200.
(18) Izquierdo, P.; Esquena, J.; Tadros, Th. F.; Dederen, J. C.; Feng, J.; GarciaCelma, M. J.; Azemar, N.; Solans, C. Langmuir 2004, 20, 65946598.
(19) Izquierdo, P.; Feng, J.; Esquena, J.; Tadros, Th. F.; Dederen, J. C.; GarciaCelma, M. J.; Azemar, N.; Solans, C J. Colloid Interface Sci. 2005, 285, 388394.
(20) Morales, D.; Solans, C.; Gutierrez, J. M.; Garca-Celma, M. J.; Olsoson.,
U. Langmuir 2006, 22, 30143020.
(21) Forgiarini, A.; Esquena, J.; Gonzalez, C.; Solans, C. Prog. Colloid Polym.
Sci. 2000, 115, 3639.
(22) Forgiarini, A.; Esquena, J.; Gonzalez, C.; Solans, C. Prog. Colloid Polym.
Sci. 2001, 118, 184189.

10.1021/la800624z CCC: $40.75 2008 American Chemical Society


Published on Web 05/20/2008

Nanoemulsions Prepared by a Two-Step Process

In these cases, emulsions can be generated spontaneously.


Therefore, the low-energy methods can be considered to generate
emulsions with small droplet sizes through this spontaneous
emulsification.
Particular attention has been directed to nonionic surfactants,
for which PEO hydration can be changed both by temperature
(phase inversion temperature (PIT) method)1620 and composition
(emulsion inversion point (EIP) method).1,21,22 With the PIT
method, emulsions stabilized by appropriate nonionic surfactants
are obtained by rapid temperature changes passing through the
hydrophile-lipophile balance (HLB) temperature. Nanoemulsions can be obtained only when the oil is completely dissolved
in a single phase prior to the nanoemulsification.1620 The
formation mechanism has been investigated by 1H-pulsed-fieldgradient-spin-echo NMR (PFGSE-NMR), indicating that a
thermally induced disruption of a bicontinuous microemulsion
is necessary to generate nanoemulsion droplets.19 In the EIP
method, water is added dropwise to a mixture of surfactant and
oil at constant temperature.1,12,21,22 The formation of nanoemulsions is generally attributed to phase instabilities during
emulsification, where the presence of lamellar crystallites and/or
bicontinuous microemulsions are thought to play critical
roles.1,21,22 However, since the systems may pass through various
different phases on the route to nanoemulsion formation, it is far
from clear which phase (if any) is key.
Nanoemulsions (so-called miniemulsions) stabilized by ionic
surfactants have also been prepared by dilution of microemulsions
with excess water.8,14,23,24 In the work reported by Pons et al.8,
nanoemulsions can be obtained either by dilution of a bicontinuous, a O/W microemulsion, or multiphase mixtures with water
contents intermediate between the two microemulsion regions.
The flexibility of the surfactant film, its affinity for water, and
the interfacial tension were suggested as important factors in the
formation of nanoemulsions in two complex systems studied by
Bataller et al.14 However, no systematic study into the mechanism
of nanoemulsion formation has been conducted for ionic surfactant
systems. Therefore, further effort is required in order to better
understand the mechanisms of nanoemulsion formation under
isothermal conditions, with both nonionic and ionic surfactants,
this will help in optimizing nanoemulsification processes for
industrial and research applications.
An isothermal two-step process, which requires no heat or
energy input, has recently been demonstrated to generate
nanoemulsions with an industrially relevant system composed
of methyl decanoate/technical grade surfactant poly(oxyethylene)
7-lauryl ether AEO-7/water.25 In this system, initial concentrates
with different water weight fractions were rapidly diluted with
water to reach the same final composition. Although the oil and
surfactant are industrial grade, these previous results25 suggest
a relationship between the formation of nanoemulsions and the
initial concentrate equilibrium-phase behavior. Bluish transparent
O/W nanoemulsions with a narrow size distribution were formed
only when the concentrate was located in a microemulsion
(bicontinuous and/or oil-in-water) region. The formation of O/W
nanoemulsions was attributed to homogeneous nucleation of oil
from the microemulsion phase upon dilution. The use of this
two-step process allow not only generation of nanoemulsions, but also improved understanding of the mechanism
of nanoemulsion formation.
Therefore, this low-energy two-step process was employed in
the present work for several academically relevant systems,
(23) Taylor, P.; Ottewill, R. H. Colloids Surf., A 1994, 88, 303316.
(24) Taylor, P.; Ottewill, R. H. Prog. Colloid Polym. Sci. 1994, 97, 199203.
(25) Wang, L. J.; Li, X. F.; Zhang, G. Y.; Dong, J. F.; Eastoe, J. J. Colloid
Interface Sci. 2007, 314, 230235.

Langmuir, Vol. 24, No. 12, 2008 6093

formulated from research grade components and well-known


surfactants: decane/pentaoxyethylene lauryl ether (C12E5)/water,
decane/dodecyldimethylammonium bromide (DDAB)/water, decane/sodium bis(2-ethylhexyl)sulfosuccinate (AOT)/water (with
100 mM NaCl), dodecane/sodium n-dodecyl sulfate (SDS)/
pentanol/water, and dodecane/hexadecyltrimethylammonium
bromide (CTAB)/pentanol/water. In this work, both contrast
variation small-angle neutron scattering (SANS) and dynamic
light scattering (DLS) were employed to gain insight into the
structure and stability mechanisms of model nanoemulsions
formulated from decane/C12E5/water mixtures at 25 C. SANS
has been used previously to study silicone oil-in-water nanoemulsions, which were generated by a high-energy extreme
shear method and stabilized by the anionic surfactant SDS, over
a range of nanoemulsion volume fractions (0.008 < < 0.6).26
As increases, the primary structure factor peak increases in
size, indicating stronger interactions and deformation of the
droplets at high above the jamming point. The systems studied
here contrast with those published previously19,26 in three essential
ways: (a) they are formed by a low-energy dilution method,
rather than PIT method19 or high-energy shear;26 (b) a direct
method SANS was used to gain insight into the equilibrium
structure and stability mechanisms of the resultant nanoemulsions,
rather than the dynamic 1H-PFGSE-NMR technique;19 and (c)
nonionic, rather than ionic, surfactants have been used,26
minimizing electrostatic structure factor effects which could
complicate the detailed analyses of SANS data.
Therefore this new work contributes to the studies on
nanoemulsions in three important ways: (1) the generality of the
low-energy dilution method has been explored for systems with
ionic, as well as nonionic surfactants; (2) internal droplet structures
of nanoemulsions can be assigned with better confidence, based
on model fitting of contrast variation SANS data; and (3) studies
with complementary scattering techniques SANS and DLS, both
sensitive to different relevant length scales, provide new insight
into the instability mechanism, suggesting it is flocculation
dominated, at least up to 4 h after the initial nanoemulsion
preparation.

2. Experimental Section
2.1. Materials. C12E5 (>98% purity), AOT (>98% purity), and
DDAB (98% purity) were purchased from Sigma-Aldrich, and SDS
(99% purity) and CTAB (98% purity) from Alfa Aesar. n-Decane
(Alfa Aesar, >99% purity), n-decane-d22 (Fluka, >99% D atom),
n-dodecane (Aldrich, >99% purity), 1-pentanol (Aldrich, >99%
purity), and D2O (Aldrich, 99.9% D atom) were used as received.
Milli-Q ultrapure water with resistivity no less than 18.4 M cm
was used.
2.2. Phase Diagrams. All components were weighed, sealed in
ampules, and homogenized with a vibromixer. The samples were
equilibrated at 25 C. Optically anisotropic liquid crystalline phases
were identified by using polarizing light microscopy (OptipHot-2,
Nikon, Japan). The boundary lines were found by consecutive addition
of one component to mixtures of the other components.
2.3. Preparation Method. Emulsions were prepared for several
different systems. For the decane/C12E5/water system, the partial
ternary phase diagram is shown in Figure 1, and sample preparation
pathways, outlined below. The nanoemulsion volume fraction, given
by the oil volume fraction plus the surfactant volume fraction ( )
decane + C12E5) was varied from 0.006 to 0.120. Two formulations
A (decane ) 0.020, C12E5 ) 0.010) and B (decane ) 0.017, C12E5
) 0.013) were chosen as the final systems. There are two important
composition parameters in the work, the overall droplet volume
(26) (a) Graves, S.; Meleson, K.; Wiliking, J.; Lin, M. Y.; Mason, T. G J. Chem.
Phys. 2005, 122, 134703/1134703/6. (b) Mason, T. G.; Graves, S.; Wiliking,
J. N.; Lin, M. Y. J. Phys. Chem. B 2006, 110, 2209722102.

6094 Langmuir, Vol. 24, No. 12, 2008

Wang et al.

Figure 1. Partial ternary phase diagram of the decane/C12E5/water system at 25 C. Samples A and B, generated from A5 and B5, respectively, are
shown in the inset photographs. The preparation protocol II is outlined in section 2.3. (Om, inverse micellar solution or W/O microemulsion; Lc,
optically anisotropic phase; MLc, multiphase region including liquid crystal phase; ME, bicontinuous microemulsion or O/W microemulsion; Em,
multiphase region).

fraction (defined above) and the weight fraction of water in the


initial concentrate w. Note, with reference to Figure 1, sample A
ends up in an emulsion region, whereas sample B, at slightly higher
surfactant concentration, is in a microemulsion region. The following
low-energy dilution pathways were followed to arrive at a suite of
samples A and B: (I) the one-step process, where components needed
to generate final compositions A and B were mixed by gentle magnetic
stirring at 25 ( 0.1 C (LTD 6G, Grant, England); (II) the two-step
process, where, first, oil, surfactant, and an appropriate amount of
water were mixed to generate a sequence of concentrates containing
water weight fractions w ) 0, 0.1, 0.2, . . . , 0.9 (these concentrates
are marked in Figure 1 and denoted as A0, A1, A2, . . . , A9 and B0,
B1, B2, . . . , B9, respectively), and then, a certain amount of one
of these concentrates was injected into an appropriate amount of
water under gentle stirring to achieve the final composition, either
A or B. As indicated in Figure 1 the dilution steps were arranged
so that A0 w A, A1 w A, A2 w A, . . . , A9 w A and B0 w B, B1
w B, B2 w B, . . . , B9 w B. The temperature was kept constant
at 25 ( 0.1 C.
Other systems with different ionic surfactants, decane/DDAB/
water,27 decane/AOT/water (100 mM NaCl),28 dodecane/SDS/
pentanol/water,29 and dodecane/CTAB/pentanol/water,29 were formulated to prepare emulsions by the two-step dilution process as
described above. The appropriate formulations were chosen as
concentrates from literature2729 and then injected into water to
yield emulsions with final volume fraction of ) 0.030.
2.4. Dynamic Light Scattering. DLS ( 500 nm) is commonly
employed to study droplet size distributions in emulsions and
nanoemulsions.1,1225 Analysis of the autocorrelation function decay
obtained by DLS30 from a dilution series of nanodroplets yields
information on the effective diffusion coefficient D. In the limit of
infinite dilution, the apparent hydrodynamic radius rh can be estimated
from

rh )

kT
6D

(1)

where is the viscosity of the medium. The apparent hydrodynamic


radius includes solvation effects. DLS experiments were carried out
using a Malvern 4800 Autosizer (Malvern Instruments, England).
In self-assembling systems such as emulsions, nanoemulsions,
and microemulsions, DLS is particularly sensitive to colloidal
(27) Blum, F. D.; Pickup, S.; Chen, S. J.; Evans, D. F. J. Phys. Chem. 1985,
89, 711713.
(28) Chen, C. H.; Chang, S. L.; Strey, R.; Samseth, J.; Mortensen, K. J. Phys.
Chem. 1991, 95, 74277432.
(29) Sripriya, R.; Muthu Raja, K.; Santhosh, G.; Chandrasekaran, M; Noel,
M. J. Colloid Interface Sci. 2007, 314, 712717.

interactions on a length scale of 2-1000 nm; hence, the apparent


size can become obscured if the samples are highly polydisperse,
insufficiently dilute, or experiencing attractive/repulsive interactions.
Hence, there are potential limitations to structural studies of
nanoemulsions by DLS alone.
2.5. Small-Angle Neutron Scattering Experiments. In comparison to DLS, SANS covers length scales which are associated
with the internal structure of nanoemulsion droplets (1-10 nm).31
An additional advantage of SANS over DLS is that contrast
variation can be used to obtain detailed structural information on
the droplets by selective deuteration of the various components.
The time-of-flight LOQ instrument at ISIS, U.K., was used; the
data acquisition, treatment, and normalization procedures have been
outlined elsewhere.3234 Samples were held in Hellma quartz cells
and thermostated at 25 C. Following standard procedures, raw data
were treated to yield normalized scattering intensities I(Q) in cm-1,
where the momentum transfer is related to the scattering angle, ,
by Q ) (4/) sin(/2). The observed Q range was 0.007-0.23 -1.
Scattering data were also corrected for wavelength-dependent
transmission factors, as well as cell, background, and any incoherent
scattering. The overall nanoemulsion drop contrast was generated
with decane-h/C12E5-h/D2O, and the external surfactant shell
contrast was highlighted with decane-d/C12E5-h/D2O. Data were
analyzed using standard Guinier limiting laws, Porod analysis, and
the multimodel FISH fitting program.35 Details of these analyses
can be found in the Supporting Information.

3. Results and Discussion


3.1. Samples Prepared by Methods I and II for the Decane/
C12E5/Water System. The partial ternary-phase diagram of the
system decane/C12E5/water at 25 C is shown in Figure 1. As
can be seen from the figure, samples A and B with volume
fraction ) 0.030 are located in the emulsion (A) and
microemulsion regions (B) of the phase diagram, respectively.
As the actual phase behavior is quite complicated,15 the exact
(30) Pecora, R. Dynamic Light Scattering; Plenum: New York, 1985.
(31) Eastoe, J. Surfactants; Wuhan University Press: Wuhan, China, 2005; pp
96-134.
(32) Dupont, A.; Eastoe, J.; Murray, M.; Martin, L.; Guittard, F.; Givenchy,
E. T.; Heenan, R. K. Langmuir 2004, 20, 99539959.
(33) Eastoe, J.; Dominguez, M. S.; Wyatt, P.; Orr-Ewing, A. J.; Heenan, R. K.
Langmuir 2004, 20, 61206126.
(34) Summers, M.; Eastoe, J.; Davis, S.; Du, Z.; Richardson, R.; Heenan,
R. K.; Steytler, D. C.; Grillo, I. Langmuir 2001, 17, 53885397.
(35) Heenan, R. K. Fish Data Analysis Program; Report RAL-89-129,
Rutherford Appleton Laboratory, CCLRC: Didcot, U.K., 1989.

Nanoemulsions Prepared by a Two-Step Process

Figure 2. Appearance of A and B prepared by method I at different


times.

Figure 3. Droplet radii for samples A (0) and B ()) with volume fraction
of 0.030 at 25 C as a function of w, the weight fraction of water in
the initial concentrate. Inset: appearance of emulsion diluted from different
concentrates.

microstructures of the liquid crystal phases have not been fully


identified.
The appearance of samples A and B prepared by the gentle
stirring method I, at different times, are shown in Figure 2. It
can be seen that a coarse, white emulsion was obtained by gentle
stirring for sample A (emulsion region). On the other hand sample
B became transparent over time, which is consistent with eventual
microemulsion formation, as expected on the basis of the phase
diagram shown in Figure 1. Samples A and B prepared by method
I were studied by DLS 24 h after preparation. System A is a
white coarse emulsion with radii 290 nm, whereas the
transparent system B gave rh ) 9.3 nm, exactly the same as for
an equivalent B sample but formulated by the dilution method
II (see below). The apparent droplet size for the presumed
microemulsion sample B was constant over 6 months.
Samples A and B prepared by method II were investigated by
DLS. The relationship between the final droplet radius and the
water weight fraction w in the concentrate is shown in Figure
3. It can be seen that, with system A, the droplet radii vary with
water level w in the concentrates. Bluish transparent nanoemulsions with droplet radii as low as about 15 nm were obtained on
dilution of concentrates with w ) 0.4 and 0.5 (A4 and A5; see
Figure 1). For concentrates with w less than 0.3 (A0-A3; refer
to Figure 1) or more than 0.6 (A6-A9; refer to Figure 1), white
coarse emulsions with radii larger than 100 nm were obtained.
On the other hand, the droplet radii for systems B remain
essentially constant at about 9 nm, regardless of water level in
the concentrate. It can be seen that microemulsions form
spontaneously, and the scattering and visual properties are
independent of the preparation pathway. However, appearance
and the droplet radii of the emulsions from pathway A are clearly

Langmuir, Vol. 24, No. 12, 2008 6095

dependent on the preparation procedure, and crucially the water


level in the concentrate.
Finally, comparing the DLS results for system A, prepared by
the two separate routes I and II (rh ) 290 and 15 nm if generated
from A5), shows that different states are achieved. This indicates
that the formulation pathway to nanoemulsion A is an important
factor.
3.2. Relationship between the Formation of Nanoemulsions
and Equilibrium Phase Behavior in the Decane/C12E5/Water
System. Electrical conductivity measurements on concentrates
of systems A with initial water weight fractions w in the range
0.34-0.57 are consistent with the presence of bicontinuous
microemulsions (see Supporting Information). It can be seen
from Figure 1 that concentrates A4 and A5 are located in the
bicontinuous microemulsion region, while other concentrates
are located in different regions of the phase diagram. Nanoemulsions with small droplet radii are formed only when the
concentrate starts as a bicontinuous microemulsion, with w of
0.4 and 0.5 (samples A4 and A5) (Figure 3). This suggests a close
relationship between the equilibrium phase behavior of the initial
concentrate and the final droplet radius of the resulting emulsions
after they have been diluted to achieve A.
These different phases are consistent with changes in the
surfactant hydration and solubility of decane as a function of
dilution. It is well-known that the preferred curvature of nonionic
surfactant layers depends on hydration of the PEO, which with
surfactant-rich concentrates can be increased by dilution with
water15 (EIP method). At low water contents the preferred
curvature may be around water, generating reverse W/O
structures. With increased water dilution, hydration of the PEO
headgroups may drive the curvature about oil to favor O/W phases.
The solubilization capacity for oil is decreased during the phase
transition, leading to local supersaturation and then oil droplet
nucleation: therefore, spontaneous emulsification occurs. For low
water level systems (A0-A3), oil is the continuous phase in these
initial concentrates. The existing oil droplets may act as nuclei
and trigger heterogeneous nucleation,36 which result in droplets
with larger sizes and polydispersity. On the other hand in
concentrates A4 and A5, the oil was completely solubilized in a
bicontinuous microemulsion phase (see Supporting Information).
The dilution of these concentrates with excess water converts
them to oil-in-water systems and decreases the concentration of
surfactants, which also inevitably leads to a decrease in oil
solubilization. Similar to concentrates A0-A3, these systems
became supersaturated in oil, leading to nucleation of oil. The
point is that supersaturation in the single phase results in
homogeneous nucleation, which favors the formation of smallsized oil droplets with narrow polydispersity.36,37 Considering
concentrates located in a region with higher water, the surfactant
layer is saturated with water and the mean curvature can be
considered as water concentration independent.15 Concentrates
A6-A9 are located in the emulsion (Em) region, in which the
oil droplets coexist with other phase(s). Highly polydisperse
emulsions were obtained by dilution of water because no curvature
(structural) change is induced by added water; the emulsion just
become less concentrated.
Therefore, nanoemulsions were obtained by dilution of
bicontinuous microemulsion, as a result of homogeneous
nucleation during a spontaneous emulsification. These results
are consistent with previous reports that nanoemulsions were
(36) Vincent, B.; Kiraly, Z.; Obey, T. M. In Modern Aspects of Emulsion
Science; Binks, B. P., Ed.; The Royal Society of Chemistry: Cambridge, U.K.,
1998; pp 100-114.
(37) Morris, J.; Olsson, U.; Wennerstrom, H. Langmuir 1997, 13, 606608.

6096 Langmuir, Vol. 24, No. 12, 2008

obtained only when the oil was completely dissolved in a single


phase prior to the nanoemulsification through the PIT method.1620 Furthermore, similar results have also obtained by an
isothermal two-step process in a system with technical grade oil
and surfactant.25
3.3. Emulsions Prepared by Method II in Systems with
Ionic Surfactants. To explore the generality of the mechanism
of nanoemulsion formation, a range of concentrates were
formulated with different surfactants. Literature was used to
guide the formulation of appropriate concentrates with
DDAB,27 AOT,28 and CTAB-pentanol and SDS-pentanol.29
For systems with DDAB, all concentrates including
bicontinuous microemulsions do not disperse in water at all;
coarsely dispersed oil droplets were observed during stirring
which phase-separated as soon as stirring was stopped. This
is consistent with the published phase behavior of the decane/
DDAB/water system:27 addition of an appropriate amount of
water transforms the bicontinuous concentrate into a discontinuous discrete droplet water-in-oil microemulsion, and then
further dilution with water places the samples in an undefined
(emulsion) phase region. There is no spontaneous curvature
transition to an oil-in-water structure in this system. Therefore,
it may be reasonable to assume that spontaneous emulsification
is also required for nanoemulsion formation as well as the
complete solubilization of oil in a bicontinuous microemulsion.
Crash dilution of concentrates with AOT, CTAB-pentanol,
and SDS-pentanol in water, to yield final volume fractions of
) 0.030, resulted in emulsions having droplet radii of the order
70 nm, when the concentrates were located in a W/O or
bicontinuous microemulsion phase region (see the Supporting
Information, Table S1). Phase separation was observed in these
systems after 3 days. This means nanoemulsions can be prepared
by dilution of concentrates even when they are not bicontinuous
microemulsions. Due to the decrease of ionic strength13,14 and
partitioning of alcohol,9,10 as reported elsewhere, dilution of these
concentrates with water allows a change of spontaneous curvature
from water-in-oil to a bicontinuous microemulsion and then to
an oil-in-water configuration.
3.4. Nanoemulsion Droplet Structure and Stability Determined by SANS. 3.4.1. Nanoemulsion Droplet Structure 30
min after Preparation. Because the nanoemulsions prepared
with decane/C12E5/water have droplet sizes ideally matched
to SANS (up to 60 nm), with the additional benefit of an
absence of interdroplet electrostatic interactions, this neutron
scattering method can be used to generate detailed structural
information.
Figure S2 in the Supporting Information shows the SANS
profile from concentrate A5, exhibiting a correlation peak at Q
0.03 -1, consistent with a bicontinuous microemulsion at
) 0.50. Nanoemulsions (sample A prepared by dilution from
A5), at drop contrast decane-h/C12E5-h/D2O, and shell contrast
decane-d/C12E5-h/D2O were studied by SANS 30 min after
preparation. These two contrasts can be clearly distinguished on
a log-log plot, as can be seen in Figure 4, and there is an absence
of any S(Q)-type correlation peak.
The drop contrast data were initially analyzed by Guinier and
Porod approximations and then finally fitted using the multimodel
FISH program (details given in the Supporting Information).
The Guinier plots show significant linearity at low Q for spheres
(ln I(Q) vs Q2, Figure S3 in the Supporting Information); no
linearity in an appropriate region of Q space was discerned for
cylinders ln [I(Q)Q] vs Q2 or discs (ln [I(Q)Q2] vs Q2). Therefore,
a Guinier analysis suggests the nanoemulsions exist in the form
of spherical (globular) droplets. Estimates for the nanoemulsion

Wang et al.

Figure 4. SANS profiles of the drop and shell contrasts of the


nanoemulsion (sample A, at ) 0.030) at 30 min and 25 C. Error bars
are shown. Lines are fits to a polydisperse sphere model described in
the Supporting Information. Inset: Schematic outlines of the contrast
arrangements.
Table 1. Values Obtained from the SANS Data from
Nanoemulsions at Different Times (Sample A at O ) 0.030)a
sample
A, 30 min
A shell, 30 min
A, 120 min
A, 240 min

RGuinier/nm RPorod/nm Cdrop/(mol dm-3) RFit/nm ts/nm


12.7

11.9

15

12.7
12.3

11.9
11.6

15
15

10.2
10.1
10.0
9.9

1.1

a
/Rc ) 0.23 ( 0.03, hs ) decane + surfactant, Rhs ) RPorod ( 0.8 nm.
Parameters: RGuinier and RPorod are radii estimated by the Guinier and Porod
approximations; Cdrop is the calculated nanoemulsion droplet concentration
as described in the Supporting Information; RFit is the average radius given
by Schulz polydisperse spheres fitting routine; ts is the apparent shell thickness
given by the core-shell model; /Rc is the width of the Schultz distribution
function.

radii R were obtained from the Guinier plot, which is valid only
for very low Q (QR < 1); at high Q, the SANS intensity is
sensitive to scattering from local interfaces. In this regime,
the Porod approximation was used to estimate particle radii and
the droplet concentration Cdrop, via the total specific area (see the
Supporting Information). These values were used as starting
points for a more detailed analysis, using a model of Schulz
polydisperse spheres.38,39 The data are described very well by
this model, with radii quite close to those obtained by the Guinier
and Porod analyses (see Table 1). These analyses were consistent
with the view that the droplets exist in the form of oil spheres.
A core-shell model35 was employed to fit the shell contrast
decane-d/C12E5-h/D2O. This model described the profile very
well, which suggests that the nanoemulsion with an oil core is
surrounded by a surfactant shell of apparent film thickness, ts.
The respective model fits, and associated parameters, are shown
in Figure 4 and Table 1. It is shown that the overall radius of
the nanoemulsion (sample A) is 10.2 nm and the thickness of
the surfactant layer is 1.1 nm. This outer layer thickness reflects
reasonably well the dimension of a C12H25-h chain (1.7 nm) and
would be consistent with minimal contrast difference between
the outer hydrated EO-5 groups and the external D2O solvent,
as would be expected under highly solvated conditions. The
(38) Eastoe, J.; Hetherington, K. J.; Sharpe, D.; Dong, J.; Heenan, R. K.;
Steytler, D. C. Langmuir 1996, 12, 38763880.
(39) Nave, S.; Eastoe, J.; Heenan, R. K.; Steytler, D.; Grillo, I. Langmuir 2000,
16, 87418748.

Nanoemulsions Prepared by a Two-Step Process

Figure 5. SANS profiles for nanoemulsions at ) 0.030 at different


times t ) 30 (0), 120 ()), and 240 min (). Error bars are shown. Lines
are model fits to a polydisperse sphere model described in the Supporting
Information. Inset: shown is the corresponding Porod plot.

overall droplet radius determined by SANS matches quite the


value determined by DLS (15 nm) considering the need to
include solvation of the surfactant shell. The polydispersity
value is about 0.23, which reflects a narrow distribution of
nanoemulsion droplet radii, and typical of microemulsion
droplets using this model.38,39
Because of preexisting literature,40 the structure of dilute
microemulsion B was not investigated by SANS here. Comparison
of the composition for system B [ ) 0.030, and R ) 0.50 ) (mass
fraction wtdecane)/(wtdecane + wtC12E5)]40c) with those investigated
previously40a,b clearly suggests spherical decane-in-water microemulsion droplets of radius 9 nm. Therefore, the DLS radii
for the microemulsion system B shown in Figure 3 are consistent
with previously published data.40 Note the microemulsion droplet
radius is consistently lower than for the nanoemulsion A (Figure
3 and ref 40), suggesting different underlying structures for the
two systems A and B.
3.4.2. Effect of Time on Nanoemulsion Droplets. The evolution of droplet size is an important aspect of emulsion stability.
Nanoemulsions were characterized by SANS, at different times
after the initial dilution to generate system A. No obvious changes
in the shape or intensity of the SANS were observed. Similar
detailed analyses as outlined above in section 3.4.1 were
performed, and the resulting values are shown in Table 1 while
the fitting lines are displayed in Figure 5. It can be seen that no
notable changes in the structure or radius of the droplets was
observed over the test time of 4 h after the initial dilution. This
is most obviously seen in the inset to Figure 5, which magnifies
the supersensitive high Q Porod region: increases in droplet size
would result in shifts of the maxima/minima to lower Q values.
The constant size demonstrates that the nanoemulsion droplets
in this system are quite stable during the test period against
coalescence41 and Ostwald ripening,8,42 both of which would be
expected to give rise to increases in droplet size.
3.4.3. Structure of Nanoemulsions as a Function of Dilution.
Systems were prepared by dilution with different water contents
(40) (a) Menge, U.; Lang, P.; Findenegg, G. H.; Strunz, P. J. Phys. Chem. B
2003, 107, 13161320. (b) Menge, U.; Lang, P.; Findenegg, G. H. J. Phys. Chem.
B 1999, 103, 57685774. (c) The composition parameter used in refs 40a and 40b
is R is an oil mass fraction ) wt decane/(wt decane + wtC12E5). Translating now
to total droplet volume fraction used in this paper, spherical droplets were
observed in ref 40b for ) 0.0141-0.141 and above R ) 0.25: the composition
of system B studied here is ) 0.030 and R ) 0.5.

Langmuir, Vol. 24, No. 12, 2008 6097

Figure 6. SANS profiles for nanoemulsions at various volume fractions,


) 0.120 (0), 0.060 ()), 0.030 (), 0.015 (O), and 0.006 (b). Error
bars are shown. Lines are model fits to a polydisperse sphere model
described in Supporting Information. Inset: shown is the Porod plot
corresponding to the SANS profiles.
Table 2. Values Obtained from the SANS Data from
Nanoemulsions at Various Volume Fractions Oa

RGuinier/nm

RPorod/nm

Cdrop/(mol dm-3)

RFit/nm

0.120
0.060
0.030
0.015
0.006

7.3
11.2
12.7
13.0
12.5

11.3
12.2
11.9
11.3
11.3

69
28
15
7
2

9.4
9.6
10.2
10.3
9.7

a
/Rc ) 0.23 ( 0.03, HS ) decane + surfactant, RHS ) RPorod ( 0.8 nm.
Parameters: RGuinier and RPorod are radii estimated by the Guinier and Porod
approximations; Cdrop is the calculated nanoemulsion droplet concentration
as described in the Supporting Information; RFit is average radius given by
Schulz polydisperse spheres fitting routine; /Rc is the width of the Schultz
distribution function.

from concentrate A5, located in the microemulsion region of the


phase diagram.
Figure 6 shows SANS data for the nanoemulsions as a function
of volume fraction , and it can be seen that the shape of the
SANS profile changes little over this range. The increase in
intensity with is an indication of an increase in the nanoemulsion
droplet number density. This can be seen clearly from the nearlinear relationship between the droplet concentration Cdrop
(calculated from the limiting high-Q intensity on the Porod plot)
and (details can be seen in Supporting Information, Figure S7)
and further verified by the linear relationship between the fitted
scale factor (fitted as polydisperse spheres) and (can be seen
in Figure S8 in the Supporting Information). Fitted and derived
parameters are summarized in Table 2 along with the results
obtained from the Guinier and Porod approximations. As is shown
in Table 2, there is no significant change in the droplet size with
(except for the R obtained from the Guinier approximation at
) 0.120, which may be attributed to the more prominent S(Q),
which is exhibited by the downward slope at low Q). Nanoemulsions with radius of the order of 10 nm were obtained, independent
of , and polydispersity indices were about 0.23, indicating a
narrow size distribution. It can then be said that nanoemulsions
may be obtained when the initial concentrates are located in a
(41) Deminiere, B.; Colin, A.; Calderon, F. L.; Bibette, J. In Modern Aspects
of Emulsion Science; Binks, B. P., Ed.; The Royal Society of Chemistry: Cambridge,
U.K., 1998; pp 261-291.
(42) Tadros, T.; Izquierdo, P.; Esquena, J.; Solans, C. AdV. Colloid Interface
Sci. 2004, 108-110, 303318.

6098 Langmuir, Vol. 24, No. 12, 2008

Wang et al.

1
8
t
) r20
r
3

( )

(2)

where r is the average droplet radius after time t, r0 is the droplet


radius at t ) 0, and is the frequency of rupture per unit surface
of the film. If the mechanism is Ostwald ripening,43 the droplet
radius should vary with time as

Figure 7. Dependence of apparent droplet radii of nanoemulsions against


time, determined by DLS, at various volume fractions, ) 0.120 (0),
0.060 ()), 0.030 (), 0.015 (O), and 0.006 (b). Inset: the observed
eventual phase separation of a nanoemulsion (sample A) after 3 days.

microemulsion region, regardless of the water content used for


dilution. Eventual droplet sizes are mainly controlled by the
structure of the concentrate and are independent of dilution. This
is different from the results reported by Mason et al.26 in that
the structure of nanoemulsions deforms from hard sphere to a
glassy structure at high volume fraction (with larger than 0.3)
through screened surface charge repulsions, perhaps because
those systems were stabilized by anionic SDS. Due to the lower
volume fraction and zero effective surface charge in the system
investigated here, the droplets of nanoemulsions are discrete
spheres at all test volume fractions, and excess water appears to
act only as a dilution medium without having any effect on the
structure.
3.5. Stability of Nanoemulsions Determined by DLS. Phase
separation of nanoemulsion sample A (prepared from A5) was
observed after 3 days (see inset to Figure 7), whereas the
microemulsion samples B showed no phase change when left for
up to 10 weeks. These results suggest that nanoemulsions are
actually thermodynamically unstable.
To follow the breakdown process, nanoemulsion droplet sizes
were determined by DLS as a function of and time; these
results are shown in Figure 7. When measured 15 min after
preparation, similar droplet radii were obtained for the different
volume fractions, agreeing with the SANS results described above.
However, the DLS-determined droplet sizes show a marked
increase with time; the higher is, the faster the droplet size
increased. These results are contradictory to those obtained by
SANS which suggest no change in the underlying droplet
structure, at least 4 h after preparation from the stock concentrates
(see Figure 5).
These discrepancies are at least consistent with the length
scale resolutions of the two scattering techniques: SANS is ideally
matched to examine the discrete droplet structure for dimensions
up to 30 nm or so; DLS being particularly sensitive to collective
motions resulting from attractive interactions of particles, certainly
with dimensions 30 nm or so and above.
To explore the mechanism of instability with nanoemulsions,
the results of Figure 7 were replotted as shown in Figure 8. The
change in droplet size with time may follow eq 2 if the instability
mechanism is coalescence:41

dr3 8 C()VmD
) )
dt
9
FRT

(3)

where r is the average droplet size after time t, C() is the bulk
phase solubility (the solubility of the oil in an infinitely large
droplet), is the interfacial tension, Vm is the molar volume of
the oil, D is the diffusion coefficient of the oil in the continuous
phase, F is the oil density, and R is the gas constant.
Figure 8 suggests that there is no linear variation upon plotting
1/r2 or r3 as a function of time. These results indicate that neither
coalescence nor Ostwald ripening are the underlying mechanisms
for the nanoemulsion instability, which are consistent with the
results obtained from SANS in section 3.4.2.
Comparing the DLS and SANS results, it is reasonable to
assume that droplet-droplet hydrodynamic interactions exist
without droplet deformation or rupture of the surfactant film
during the test period. Therefore, the mechanism for the instability
in this system may be attributed to flocculation,43 where drops
cluster without significant rupture of the stabilizing interfacial
layer. The droplet radius determined by DLS may be attributed
to an effective cluster radius, which increases as more droplets
aggregate with time through flocculation. These aggregates rise
under gravity because of the density difference between the
dispersed oil and the continuous phase, resulting in creaming
and eventually phase separation, presumably with accompanying
coalescence. At least over the time scale studied here, up to 4 h,
it can be said that the individual droplet radius remains constant:
SANS data are inconsistent with growth of individual droplets,
which would be an inevitable consequence of both coalescence
and/or Ostwald ripening.
The rate of flocculation depends on the product of a frequency
factor (how often drops encounter each other) and a probability
factor (how long they stay in contact).43 Thus it is easy to
understand that the higher volume fraction (in which drops
have a higher frequency of encounters), the faster flocculation
occurs, which results in the faster increase in droplet radius as
determined by DLS (as can be seen in Figure 7). Tests were
performed by DLS, swapping H2O for the higher density D2O
to assess if this would affect the flocculation (Figure S10), but
no significant differences were observed using this technique.

4. Conclusions
An isothermal two-step process (method II) has been applied
to oil/surfactant/water systems with C12E5, DDAB, AOT,
SDS-pentanol, and CTAB-pentanol. Oil-in-water nanoemulsions have been obtained in the decane/C12E5/water system, by
crash-dilution of a bicontinuous microemulsion into a large
volume of water. The structure of nanoemulsions obtained this
way is mainly controlled by the structure of the initial concentrate,
being apparently independent of the dilution factor with water.
Coupled with the results gained with systems stabilized by DDAB,
AOT, SDS-pentanol, and CTAB-pentanol, a general mechanism for nanoemulsion formation may be postulated as
homogeneous nucleation of oil droplets during spontaneous
emulsification.
(43) Binks, B. P. In Modern Aspects of Emulsion Science; Binks, B. P., Ed.;
The Royal Society of Chemistry: Cambridge, U.K., 1998; pp 17-38.

Nanoemulsions Prepared by a Two-Step Process

Langmuir, Vol. 24, No. 12, 2008 6099

Figure 8. DLS derived radii, 1/r2 as a function of time (a) and r3 as a function of time (b) for nanoemulsions at various volume fractions, ) 0.120
(-0-), 0.060 (-)-), 0.030 (--), 0.015 (-O-), and 0.006 (-b-).

The droplet sizes of nanoemulsions have been determined


both by DLS and SANS. The two techniques show no striking
differences in the nanoemulsion radii if newly prepared. However,
droplet structure determined by SANS shows no change over
4 h, while the apparent DLS droplet size increases with time.
This discrepancy suggests flocculation may be responsible for
the nanoemulsion instability (at least up to emulsion ages of 4 h),
which is different from the mechanisms widely reported before,
such as coalescence and Ostwald ripening.8,41,42
This work gains insight into the general mechanism of
nanoemulsion formation, which can be used to guide the
preparation of nanoemulsions in a wide range of systems. This
two-step process is easy to scale up and has low energy
consumption, which is of great interest for practical applications.
The findings verify that the two-step process can be used for the
preparation of nanoemulsions, which are stable without an

increase in the radii of individual droplets over several hours.


This new nanoemulsification route paves the way to new potential
applications of these easy-to-prepare systems in fields such as
foods, pharmaceuticals, and agrochemicals.
Acknowledgment. We acknowledge the National Natural
Science Foundation of China (Grant NSFC 20573079) and the
Ministry of Science and Technology (Grant 2006 BAE01A075) for financial support. STFC (U.K.) is thanked for provision
of beam time at ISIS and for financial support for travel and
consumables.
Supporting Information Available: Details of the conductivity
data, SANS analysis, and theoretical background (pdf). This material
is available free of charge via the Internet at http://pubs.acs.org.
LA800624Z

Вам также может понравиться