doi:10.1016/j.cis.2007.02.002Advances in Colloid and Interf
Reverse micelles: Inert nano-reactors or physico-chemically active
guides of the capped reactions
Vuk Uskokovi , Miha Drofenik
Available online 27 February 2007
Abstract
Reverse micelles present self-assembled multi-molecular entities
formed within specific compositional ranges of water-in-oil
microemulsions. The structure of a reverse micelle is typically
represented as nano-sized droplet of a polar liquid phase, capped
by a monolayer of surfactant molecules, and uniformly distributed
within a non-polar, oil phase. Although their role in serving as
primitive membranes for encapsulation of primordial
self-replicating chemical cycles that anticipated the very origins
of life has been proposed, their first application for
‘parent(hesis)ing’ chemical reactions with an aim to produce
‘templated’ 2D arrays of nanoparticles dates back to only 25 years
ago. Reverse micelles have since then been depicted as passive
nano-reactors that via their shapes template the growing
crystalline nuclei into narrowly dispersed or even perfectly
uniform nano-sized particles. Despite this, numerous examples can
be supported, wherefrom deviations from the simple unilateral
correlations between size and shape distribution of reverse
micelles and the particles formed within may be reasonably implied.
A rather richer, dynamical role of reverse micelles, with potential
significance in the research and design of complex, self-assembly
synthesis pathways, as well as possible adoption of their
application as an aspect of biomimetic approach, is suggested
herein. © 2007 Elsevier B.V. All rights reserved.
Keywords: Colloids; Microemulsion; Nanomaterials; Reverse micelles;
Review
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . 23 2. The need to reevaluate the functional representation of
reverse micelles . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . 24 3. Examples of chemical ‘butterfly effects’ in
reverse micelle-assisted syntheses . . . . . . . . . . . . . . . .
. . . . . . . . . . . . 26 4. The example of nickel–zinc ferrite .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . 27 5. The example of
lanthanum–strontium manganite . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . 28 6. Correlations
with the biological context . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 7.
Future directions in the application of reverse micelles . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. 30 8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . 31 References . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . 32
1. Introduction
Corresponding author. E-mail address:
[email protected] (V.
Uskokovi).
0001-8686/$ - see front matter © 2007 Elsevier B.V. All rights
reserved. doi:10.1016/j.cis.2007.02.002
Boutonnet et al. first reported synthesis of a material via using
reverse micelles [2]. Numerous other nanostructured materials,
ranging from metallic catalysts [3–8] to semiconductor quantum dots
[9–11] to various ceramic materials [12–16], silica and gold coated
nanoparticles [17–21], latexes and polymer composites [22–24],
double-layered nanoparticles [25] and even superconducting
materials [26,27] have been prepared since then by means of reverse
micelle technique. However, the
Fig. 1. A drawing of a reverse micelle (a) and a computational
model (b) of reverse micelle [28]. Blue spheres represent
surfactant head groups, whereby smaller yellow spheres denote
counterions. Note that the surfactant head groups do not completely
shield aqueous interior of the modeled reverse micelle (b). (For
interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
24 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
explanation models that are typically invoked in the frame of such
experiments ordinarily refer to purely ‘templating’ role of reverse
micelles. As inert ‘nano-cages’, they are conceived as only
limiting the growth of precipitated nuclei, so that the initial
narrow dispersion of micelles in relation to their sizes and shapes
becomes reflected on a similar uniformity of the eventually
produced nanoparticles. Through referring to numerous deviations
from such an oversimplified picture, this review will challenge an
idea according to which the only role that reverse micelles have in
the processes of preparation of nanoparticles is their ‘templating’
effect and superimposition of spherical shapes upon the growing
nuclei.
2. The need to reevaluate the functional representation of reverse
micelles
Reverse micelles are typically depicted as spherical nano-
droplets, uniformly capped with a monolayer of surfactant molecules
(Fig. 1a), and isotropically distributed within an oil phase.
However, a recent attempt to model the structure of a reverse
micelle resulted in an image of a multi-molecular aggregate wherein
surfactant head groups did not completely shield aqueous interior
of the modeled micelle (Fig. 1b), indicating the need to reevaluate
the typical representations of micelles as perfectly
surfactant-capped and overstatically configured molecular
aggregates [28].
The field of reverse-micellar synthesis of nanostructured materials
is permeated by representations of passive and solely templating
role of the micelles in the course of particle formation processes.
Simple, parametric correlations are routinely used to predict and
explain the particle size out of the initial microemulsion
structure. Most notably, Pileni et al. proposed that the size of
particles obtained by precipitation in reverse-micellar
microemulsions based on sodium bis(2- ethylhexyl) sulfosuccinate
(AOT) as a surfactant ought to be equal to 1.5 times the
water-to-surfactant molar ratio in nanometers [29,30]. Carpenter et
al. suggested that the size of precipitated particles in reverse
micelles that comprise cetyl- trimethylammonium bromide (CTAB) as a
surfactant should be
equal in nanometers to the water-to-surfactant molar ratio of the
parent microemulsion [31]. Although the former relationship was
verified only for certain compositions of specific AOT- based
microemulsions and particles prepared within [30], it has been
frequently mistaken as corresponding to all types of microemulsions
and particles [32].
As a response to such an oversimplification, numerous cases of
experimental deviations from the proposed correlations were
reported [5,6,33,34]. It is not only that water-to-surfactant molar
ratio in reverse-micellar ranges of the given microemulsion phase
diagrams does not correspond to micellar sizes in direct proportion
in all cases, but the very same small-angle X-ray scattering (SAXS)
characterization technique that was relied upon in defining the
mentioned relationship between water-to- surfactant molar ratio and
the size of produced particles [30,35– 37], has shown that micellar
radii in the same AOT/isooctane/ water system change in response to
an addition of small amounts of compounds solubilized in the
microemulsion [29]. Experimental results indicate that the size of
reverse micelles depends not only on water-to-surfactant molar
ratio, but also on identity of all included microemulsion
components, their respective concentrations, pH, temperature and
ionic interac- tions caused by introduced electrolytes or
inherently dissociated molecular species [1]. Also, the particle
formation processes necessarily affect the structure of a parent
emulsion, resulting in a feedback interaction that ends as either a
form of phase segregation or a metastable state in cases when
isotropic colloidal dispersion structure is preserved.
It has been known that phase diagrams of microemulsions derived
with and without the presence of the prepared material or any other
additional component may be drastically different [38]. Therefore,
in light of such mutual transformations, the concept of
‘templating’ as translation of shapes and sizes of self-assembled
organic species onto the structure of nucleated and grown
crystallites looks as if it needs to be reevaluated, particularly
in the area of reverse-micellar preparation of materials where the
phrases like ‘nano-cages’, ‘nano-templates’ or ‘nano-reactors’ seem
to dominate the explanations of particle formation
mechanisms.
Table 1 Macroscopic and nanoscopic variables in the
microemulsion-assisted and particularly reverse-micellar synthesis
of nanoparticles
Macroscopic parameters Nano-sized parameters
Static, size and shape distribution of micelles
Microemulsion composition
Aggregation number
Water-to-surfactant molar ratio
Dynamic interaction, rates and types of merging and dissociation of
micelles
pH
Dissolved species concentrations Surfactant film curvature and
head-group spacing
Method and rate of introduction of species Effective Coulomb
repulsion potential
Temperature and pressure Aging times Van der Waals, hydrogen
and
hydrophobic interactionsMethod and rate of stirring Homogeneous
or
heterogeneous nucleation Screening length
25V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
For the most surfactant-mediated syntheses, connection between
morphology of the surfactant aggregates and the resulting particle
structure is more complex (than simply relating the average size
and shape of the micelles to size and shapes of the precipitated
particles) and affected by almost irreducible conditions that exist
in the local microenvironments that surround the growing particles
[39]. These molecular-level variables are subject to change with
macroscopically manipu- lated experimental conditions, as is shown
in Table 1. Composition, pH, concentration of the reactants, ionic
strength and heat content are some of the experimentally modified
variables that co-influence this local environment. As chemical
reactions and physical transformations caused by aging take place
within a colloid and its corresponding microenviron- ments, many of
these factors are subject to change. The decoupling of effects that
belong to each specific macroscopic modification of the system
presents one of the biggest challenges in the practical field of
colloid science.
Another oversimplified idea in the area of reverse-micellar
synthesis of nanoparticles is that the size of the produced
particles is supposed to be equal to the size of the micelles that
cap and limit the growth of individual crystallization nuclei.
Despite such a picturesque representation of the processes of
particle formation inside the so-called ‘nano-reactors’ (i.e.
‘water pools’) of reverse micelles, numerous cases wherein the
variations in the produced particle sizes could not have been
correlated with sizes of the reverse micelles, were reported
[35,40]. The size of colloid units or any other relevant property
of a colloid system can be considered as dependent not upon any
single internal variable, but only on the complex inter- actions
that are conditional for their existence. Many cases support the
idea that the reasons for the frequent mismatch of the properties
and quantities derived by different experimental methods do not
result from errors inherent in the experiments, but are evidential
of a fundamental shortcoming in the single parameter models [41].
Such a situation is highly reminiscent of
numerous attempts to infer hydrophobic interactions from
molecular-scale surface areas alone, even though bulk driving
forces and interfacial effects compete in determining hydro- phobic
effects in any particular case [42]. As a more reasonable
explanation, the dynamic interaction among colloid aggregates has
since lately been generally considered as the most important factor
that influences the morphology and the properties of the final
reaction products [43]. However, since dynamic interac- tion of
colloid multi-molecular aggregates, such as micelles, cannot be yet
directly observed in real-time conditions, indirect techniques are
usually applied in order to evaluate both static and dynamic
properties of the corresponding media. In the approximations
(introduced in order to overcome the limitations of
characterization techniques in terms of sampling, experiment time
scales, etc.) and different implicit presuppositions of various
such techniques are present the reasons for a frequent mismatch
[44,45] between the concluded properties attributed to the same
systems by using different experimental methods.
Unlike some of the surfactant-templating syntheses that can be
considered as structurally transcriptive (a copying or casting as
in the cases of some porous inorganics [46]), ‘templating’ of
crystallization processes within fine and sensitive, advanced
colloid systems such as microemulsions and particularly reverse
micelles can be regarded first as synergistic and only then as
reconstitutive [47]. Despite the fact that only spherical or
elliptical micelles have been detected and theoretically predicted
so far, beside spherically shaped particles, various other exotic
morphologies, including nanorods, nanofilaments, acicular
particles, star-shaped patterns etc, were prepared by relying on
this method. When a microemulsion-assisted synthesis of copper
nanocrystals was performed in the presence of sodium fluoride,
sodium chloride, sodium bromide or sodium nitrate, small cubes,
long rods, larger cubes and variety of shapes resulted,
respectively [48]. Variations of salt identities and concentrations
in another case of preparation of copper nanocrystals also resulted
in drastic morphological changes [49].
Although most of the particles produced in reverse-micellar,
AOT-based microemulsion systems were spherical in nature [50],
crystallization of barium sulfate resulted in extended crystalline
nanofibers aligned to form superstructures, whereby a precipitation
of barium chromate in the same microemulsion system resulted in
primary cuboids aligned to linear ‘cater- pillars’ or rectangular
mosaics [47]. In the case of synthesis of calcium phosphates,
variations in relative concentrations of the microemulsion
components resulted in various different morphologies, ranging from
co-aligned filaments to amorphous nanoparticles, hollow spheres,
spherical octacalcium phosphate aggregates of plate-shaped
particles, and elongated plates of calcium hydrogen phosphate
dehydrate [51]. Moreover, in the first historical report on the
synthesis of materials in reverse- micellar media [2], it was
observed that size of the prepared platinum, rhodium, palladium and
iridum particles was always in the range of 2–5 nm, independently
on surfactant, water and reactant concentrations applied in the
experiments [52].
Far from being only inert constraints to the growth of
crystallites, microemulsions were shown to be
physico-chemically
26 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
active in defining the reaction pathways that take place in their
presence, thus influencing the very chemical identities of the
final products [53]. Specific intermolecular interactions at the
hydrophilic sides of surfactants surrounding the aqueous cores,
intense local electric fields and significant level of cooperative
weak molecular forces that modify the local microenvironments
comparing to bulk conditions, as well as specific structure and
solvent properties of water at close interfacial distances, are
proposed to have catalytic effects on the rates of chemical changes
[54,55].
The behavior of liquid molecules confined in nano-sized spaces or
at solid–liquid interfaces in general, due to surface- induced
structuring, significantly differs from their behavior within a
bulk system [56]. Fourier-Transform Infrared (FTIR) spectroscopic
studies have indicated that the water interior of a reverse micelle
has a multilayered structure, consisting of interfacial,
intermediate and core water. The interfacial layer is composed of
water molecules that are directly bounded by polar head groups of a
surfactant; the intermediate layer consists of the next few
nearest-neighbor water molecules that can exchange their state with
interfacial water; and the core layer is found at the interior of
the ‘water pool’ and has the properties of bulk water [57].
Depending on the size of reverse micelles, available water may have
significantly different solvent properties, ranging from highly
structurized interiors with little molecular mobility [58,59] to
free water cores that approximate bulk water solvent
characteristics.
Different water structures may also dissolve different amounts of
gases, which can drastically influence the reaction pathways,
particularly in the cases where oxidation or reduction reactions by
means of dissolved gases comprise crucial steps in preparation
procedures, as the numerous cases of ferric oxides and complex
corrosion phenomena may illustrate [1]. The accumulated gases are
significantly present at hydrophobic interfaces [60] comparing to
the typical range of dissolved gas concentration in water at normal
pressure and temperature (∼5 ·10−3 M). Fine variations in the
experimental outcomes depending on the gas effects have been
noticed [53], and there were cases where certain effects, which
depended on many parameters, disappeared on removal of the
dissolved gas [60].
Interfacial self-association mechanisms can also be quite different
depending on the surface wettability. As a biological example, the
rate of blood coagulation tends to increase with an increase in
water-wettability of the tube surface [61]. Also, self- assembly
processes that occur during the drying steps of synthesis
procedures involve complex competition between the kinetics of
evaporation and the time scales with which solvated nanoparticles
diffuse on a substrate, and due to the specific role of hydrophobic
interactions and a variety of ways to nucleate evaporation may lead
to unexplored territories in the field of novel design [42].
Anyhow, treating water as a continuum medium in both theoretical
approaches (such as in the framework of DLVO theory) and
explanation of experiments, altogether with disregarding its fine
interactions with gases, salts and electromagnetic fields may in
future indeed cause ever increasing difficulties in attempts to
explain fine variations from the ranges of expected results.
3. Examples of chemical ‘butterfly effects’ in reverse
micelle-assisted syntheses
As far as the current state-of-the-art is concerned, it is
exceedingly difficult to predict the outcomes of experimental
settings aimed to produce novel fine structures and morphol- ogies
by means of reverse-micellar methodology, and the most attractive
results in this practical field come from trial-and-error
approaches. There are many evidences that slight changes in the
limiting conditions of particle synthesis experiments can produce
significant differences as the end results [1]. The following
examples may illustrate such a proposition and enrich one's belief
in crucial sensitivity and subtleness of the material design
procedures that involve wet environments and colloidal phenomena in
general.
Replacement of manganese ions with nickel ions in an ex- periment
of reverse-micellar precipitation synthesis of a mixed zinc–ferrite
resulted in the production of spherical particles in the former
case [62] and acicular ones in the latter [63,64]. When bromide
ions of cetyltrimethylammoniumbromide (CTAB) surfactant were in a
synthesis of barium-fluoride nanoparticles replaced by chloride
ions (CTAC), identity of the final product was no longer the same,
whereas a replacement of 2-octanol with 1-octanol significantly
modified crystallinity of the obtained powder [35]. Various choice
of precipitation agents can often result in distinctive
morphologies obtained [65,66].
The following examples may illustrate the idea that often routinely
neglected influences in the preparation procedures may leave
significant traces on the properties of the final products.
It has been evidenced that even the method of stirring in some of
the microemulsion-assisted procedures of preparation of
nanoparticles can have decisive influence on some of the final
particle properties. Thus, using a magnetically coupled stir bar
during an aging of a dispersion of particles influenced crystal
quality and in some cases resulted in a different crystal structure
as compared with non-magnetically agitated solutions [40]. In case
of a synthesis of organic nanoparticles in reverse micelles, the
use of magnetic stirrer led to the formation of nanoparticles
larger in size comparing to the particles obtained with using
ultrasound bath as a mixer, even though no changes in particle size
were detected on varying solvent type, microemulsion composition,
reactant concentrations and even geometry and volume of the vessel
[67].
Changes in the sequence of introduction of individual components
within a precursor colloid system could result in different
properties of the final reaction products [68]. Such a property is
directly related to the fact that microemulsions, like all colloid
systems, do not present thermodynamically equilib- rium phases that
spontaneously form, but are thermodynami- cally unstable and only
due to the existence of large interfacial energies that are
stronger than thermal energy, kT, their order is preserved.
Changes in size of a volume where the particle preparation
processes take place – as occurs when the transition from small-
scale research units to larger industrial vessels is attempted –
can lead to extensive variations in some of the properties of the
synthesized material [69]. For instance, absorptivity of
Fig. 2. X-ray diffraction patterns of the powders synthesized by
the same precipitation procedure, with (upper) and without (lower)
reverse-micellar microemulsion. The peaks denoted with an S are
ferrite-derived spinel reflections, whereas the peaks denoted with
a D are δ-FeOOH-derived.
27V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
cadmium-sulfide particles dramatically changed when the amounts of
the microemulsions used in the synthesis procedure were tripled
[70]. Also, when two same chemical procedures for a colloid
synthesis of nanoparticles were performed in closed and open,
otherwise identical vessels, perfectly uniform spherical particles
were yielded in the former case, whereby elongated particles of
similarly narrow size distribution were produced in the latter
[69].
Numerous examples of unexpected effects of reverse micelles on the
kinetics of encapsulated reactions may be provided as well. As
amatter of fact, whereas kinetic conditions in ordinary solutions
may reasonably be approximated as continuous, dynamics of solvation
effects and reaction kinetics can – depending on the structure of
the microheterogeneous colloid system – largely vary in different
local microenvironments, effectively producing sig- nificantly
complex outcomes. Slight changes in micellar dispersity towards
wider polydisperse distributions have, for instance, been shown as
capable of triggering the processes of Ostwald ripening of the
colloid particles that result in complete phase segregation [71].
The dynamics of solvation effects can drastically change with an
interfacial distance, which may prove to be a significant effect in
the cases of chemical reactions performed within micellar
aggregates.
The rate constants of chemical reactions performed within micellar
aggregates include the effects of Brownian diffusion of reverse
micelles, droplet collision, water channel opening, complete or
partial merging of micelles, diffusion of reactants and the
chemical reaction, as well as fragmentation of transient dimers or
multimers (wherein the slowest step determines the temporal aspect
of the overall process of synthesis) [72], rang- ing from the order
of magnitude of nanoseconds for diffusion- controlled intermicellar
reaction to an order of miliseconds for intermicellar exchange of
reactants [43]. However, despite the fact that dynamic response in
colloid systems is typically much slower compared to their bulk
counterparts [28], extremely fast responses may be favorable under
certain conditions, as can be illustrated by numerous examples of
catalytic effects produced by the influence of micellar
encapsulation [32,54,55,73] and
exchange [43,74,75] of reactants. For example, the rate con- stant
of the hydrolysis of acetylsalicylic acid in the presence of
imidazole catalyst increased by 55 times when the reaction was
performed in AOT/supercritical ethane microemulsion com- pared to
the aqueous buffer [74]. Numerous other AOT-based microemulsions
have been shown to possess catalytic effects upon particular
hydrolysis reactions [75]. It has also been reported that the rate
of oxidation of Fe2+ and a subsequent formation of needle-shaped
FeOOH particles by spontaneous air oxidation is from 100 to 1000
times faster in reverse micelles than in a bulk solution,
regardless of the differences in surfactant or other conditions
[73]. In the case of certain iron complexes, a two to tenfold
increase in the rate of dissociation was correspondingly measured
in comparison to pure aqueous solution [32].
4. The example of nickel–zinc ferrite
When the chemical procedure of preparation of δ-FeOOH is performed
in the presence of CTAB/1-hexanol/water reverse- micellar
microemulsion of particular composition, nickel–zinc ferrite is
obtained instead [53], as can be observed from Fig. 2. Faster rates
of oxidation and slower rates of precipitationwhen the synthesis is
performed in reverse micelles rather than in bulk conditions, are
suggested as the reason for the difference in chemical identities
of the final products. The reason for the faster rate of Fe2+
oxidation in reverse micelles compared to the bulk conditions might
lie in the atypical structure of water as a solvent in reverse
micelles. Oxidation of initial Fe2+ ions is generally regarded as
the first step in nucleation of precipitating, ferrite or
ferric-oxide phases [76]. It was suggested that the increase in
hydrogen bonding between water molecules in a thin layer
neighboring to surfactants may favor the transfer of electrons from
Fe2+ to Fe3+ by a tunnelling effect [54], whereby the excess
electrons will be consumed in aqueous solution to produce hydroxide
ions in the presence of dissolved oxygen. The oxidation of Fe2+
with the decomposition of H2O is, by considering thermodynamic
data, proven to be an energetically favorable
Fig. 3. XRD patterns of the as-dried powder synthesized by
hydroxide co-precipitation procedure in solution (a), of the same
powder calcined at 450 °C (b) and 600 °C (d) for 2 h in air, and of
the sample co-precipitated within hydroxide approach in reverse
micelles and calcined at 450 °C in air for 2 h (c).
28 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
process [77], and the solvent properties of reverse-micellar water
may significantly influence the process of oxidation and,
therefore, the crystallization of novel ferric-oxide phases.
We have previously shown that slight changes in the composition of
the parent reverse-micellar microemulsion may result in
significantly changed physical properties of the prepared powders
[78]. In case of the investigated synthesis of nickel–zinc ferrite
particles of specific composition, we unexpectedly arrived to areas
in the phase diagram of CTAB/ 1-hexanol/water microemulsion where
drastic increases in specific magnetization resulted from otherwise
identical preparation procedures [79]. A material with average
particle size of 10 nm and specific magnetization of 50 emu/g
(which is about two-thirds of the magnetization that sintered and
commercial nickel–zinc ferrites possess), was prepared by employing
such a technique at almost the room temperature with less than an
hour of aging time [79]. This effect was explained by referring to
the particular composition of the parent microemulsion employed,
wherein micellar percolation effects that led to efficient
redistribution of micellar contents were pronounced. Depending on
whether the encapsulated reactions were initiated by diffusion of
one of the reactants through the oil phase or by collision, merging
and micellar content exchange, the final product could end up with
having significantly different properties [64,78,79]. Similar as in
the field of evaluation of environmental and toxicological effects
of nanoparticles where small variations in chemical structure or
particle size may lead to drastic differences in the investigated
outcomes [80], the formulations of overgeneralized concepts in the
field of reverse-micellar synthesis of materials are proven as
exceedingly difficult in light of such sensitivity of the final
outcomes upon seemingly negligible variations in the initial
conditions of the synthesis experiments.
5. The example of lanthanum–strontium manganite
The following example related to reverse micelle-assisted
preparation of lanthanum–strontium manganites may offer
significant insights into how different mechanisms of formation of
identical compounds may proceed with and without the presence of
reverse micelles [81]. Similar as in the case of nickel– zinc
ferrite, performing identical chemical procedures in bulk
conditions and in the presence of reverse micelles resulted in
different chemical identities of the final powders. Whereas
precipitation of precursor cations in the form of oxalates from
aqueous solutions was limited by the formation of [Mn(C2O4
2−) NO3
−] coordination complexes (hence aqueous–alcoholic solu- tions had
to be employed), such an effect was absent when identical reaction
was performed within reverse micelles of CTAB/1-hexanol/water
microemulsion. Whereas strong bases, such as NaOH, could in aqueous
solution yield precipitate that would form the desired monophase
manganite upon annealing, and weak bases, such as (CH3)4NOH, could
not raise pH to sufficient level that would induce the subsequent
solid-state formation of manganite compound, completely different
situation was observed in the case of precipitation in reverse
micelles. Whereas strong bases led to disruption ofmicroemulsion
structure and phase segregation, the use of (CH3)4NOH as
precipitating agent resulted in sufficiently high pH levels that
favored the complete precipitation of cations and eventual
formation of pure manganite products.
The difference in the annealing mechanism of the formation of
bulk-prepared and microemulsion-assisted-prepared LaSr- manganite
powders – after the precursor cations were precipitated in forms of
hydroxides [82] – can be observed by comparing the X-ray
diffraction (XRD) patterns presented in Fig. 3. Whereas in case of
the bulk synthesis, the growth of SrCO3 crystallites comprising the
as-dried powder as well as the transformation of La(OH)2 into
La2O2CO3 is evident from comparing the XRD patterns (a) and (b),
the transformation of qualitatively identical as-dried powder as
prepared in micro- emulsion into an amorphous, more homogeneous
transient composition, is obvious by comparing XRD patterns (a) and
(c). Both powders after heating for 2 h in air at ≥600 °C yield
manganite perovskite samples. However, whereas the changes in
crystal structure, going from tetrahedral to orthorombic
Fig. 4. Dependencies of the average particle size (d) and crystal
lattice parameter (a) on the calcination temperature in the bulk
manganite-synthesis case (left), and of the average particle size
vs. calcination temperature for the sample synthesized by using
reverse micelles (right).
29V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
followed by the increase in La stoichiometric proportion (due to
gradual incorporation of La3+ from oxycarbonate transient compound
into the manganite phase) and the decrease in Mn proportion (due to
the compensation of charges), with XRD- determined average particle
size kept constant (Fig. 4a), are noticed with the further increase
in the temperature of calcination in case of the bulk-synthesized
sample, a linear increase in average particle size with calcination
temperature (the mean value of crystal lattice parameter being
constant at 0.5474 nm) is noticed in case of the
microemulsion-assisted- synthesized sample (Fig. 4b), obviously due
to more homoge- neous re-crystallization processes inherent in the
annealing transformation of the latter as-dried composition into
the manganite phase. Therefore, besides different mechanisms of
manganite formation up to 600 °C, the effect of the further linear
increase in magnetization with annealing temperature (observed in
both cases) is attributed to thoroughly different mechanisms:
rearrangement of crystal structure in the bulk–
Fig. 5. Normalized XRD patterns of the samples synthesized using
oxalate co-precip 500 °C (a, b), 700 °C (c) and 1000 °C (d) for 2 h
in air.
synthesis case, and grain growth in the microemulsion– synthesis
case.
In case of the synthesis of the same compound by precipitation of
precursor cations in form of oxalates, the comparison between
microemulsion-assisted and the bulk case yields thoroughly opposite
observations [83]. Namely, the process of the manganite formation
follows more homogeneous route when the approach in the bulk
solution is followed, comparing to the microemulsion-assisted
procedure. In case of the bulk synthesis, a mostly amorphous
transient structure is detected at 500 °C (Fig. 5a), whereby after
annealing at the same conditions, transient phases of La2O2CO3 and
cubic Mn2O3 are detected in case of the microemulsion synthesis
(Fig. 5b). The formation of the manganite is completed after the
heat treatment at ≥1000 °C in case of the latter approach (Fig.
5d), whereby 700 °C is proven to be sufficient temperature for the
desired manganite formation in case of the synthesis in
hydroalcoholic solution (Fig. 5c).
itation approach in bulk solution (a, c) and in reverse micelles
(b, d), annealed at
30 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
6. Correlations with the biological context
Reverse micelles have recently been proposed as candidates for the
most primitive membranes that hosted the first planetary
self-replicating chemical reactions that became the precursors of
living processes in the evolution of life [84]. Apart from the use
of reverse micelles in preparative organic chemistry for
compartmentalization and selective solubilization of reactants,
separation of products [55] and phase transfer [85], they have been
used in the field of biochemistry both for storing bioactive
chemical reagents [86] and as catalyzers [87,88] or inhibitors [89]
of biochemical enzyme-driven reactions. Because encap- sulating a
protein in a reverse micelle and dissolving it in a low- viscosity
solvent can lower the rotational correlation time of a protein and
thereby provide a strategy for studying proteins in versatile
environments [90], reverse micelles are used as a cell
membrane-mimetic medium for the study of membrane interactions of
bioactive peptides [91]. The observations that denaturation of
proteins can be prevented in reverse micelles [92] have spurred
even more interest in the application of these self-organized
multi-molecular assemblies as either drug- delivery carriers or
life-mimicking systems [93]. Such a biomimetic role of reverse
micelles has been further instigated by the discovery of
possibility of initiating self-replication of reverse micelles due
to reactions occurring within micellar structures [94,95]. As a
matter of fact, positioning reverse micelles right at the interface
between the domains of ‘living’ and ‘non-living’ may present a
crucial shift in improved understanding of their function and
bioimitative utilization of such knowledge for practical
purposes.
Such a widening shift in understanding of the roles of reverse
micelles in materials synthesis experiments goes together with the
current trend of thinking according to which neither lipid
membranes are seen anymore as passive matrices for hosting
biomolecular reactions [96], confirming that cellular activities
are in large extent controlled by lipids in addition to
conventional protein-governed mechanisms [97]. Although it is known
that chemical self-replication reactions need a sort of protection
membrane to selectively absorb the influences of the environment,
that is to say require “a sophisticated cradle to be lulled in”
[98], how these protective mesophases indeed ‘sing’ presents a
challenge for the future investigations.
Knowing that by actively regulating the flow of chemicals between
the cell and its surroundings and conducting electric impulses
between nerve cells, biological membranes play a key role in cell
metabolism and transmission of information within an organism
highlights the practical significance of investiga- tions oriented
towards reproducing or at least approaching a reproduction of such
an organizational complexity in artificial colloid systems. Also,
knowing that malignant cells have sig- nificantly different surface
properties comparing to normal cells [99], maybe the transition of
focus in apoptosis research away from the genetic code disruption
as the sole key influence towards information transmission
mechanisms that involve membrane mediation would herein
beneficially switch the major scope to the cellular epigenetic
network and finally to more holistic biological and biomedical
perspectives. Such an
integrative view at cellular structures may be further instigated
by the recent findings that a large percentage of body cells
(cardiac muscle cells, in particular) is, similar as the aforemen-
tioned reverse micelle model [28], in a ‘membrane-wounded’ state,
suggesting that continuous protective barrier is not essential for
cell functioning [100]. Also, if the cytoplasmatic medium is,
instead as an ordinary solution, considered as a colloid gel, rich
with interfaces between water and intracellular proteins,
polysaccharides, nucleic acids and lipid membranes, then an array
of interesting characteristics related to water- retaining
properties of cellular gel matrices may be reasonably arrived to,
similar as in the case of uninvestigated influence of unusual
structure and solvent properties of water confined in
reverse-micellar regions.
Both self-organization phenomena in living organisms and
self-assembly effects of amphiphilic mesophases are governed by
multiple weak interactions, such as hydrogen bonds, hydro- phobic
and hydrophilic interactions, van der Waals forces, salt bridges,
coordination complexes (forces involving ions and li- gands, i.e.
‘coordinate–covalent bonds’), interactions among π-electrons of
aromatic rings, chemisorption, surface tension, and gravity [101].
Whereas the traditional field of chemistry developed by
understanding the effects of covalent, ionic and metallic bonding
forces, an extension of the same approach to weak intermolecular
forces is nowadays suggested as a natural direction for achieving
future prosperity within the practical aspect of the field of
chemistry [102]. With attaching a more significant role to reverse
micelles in the prospect of advanced structural design, a general
shift towards approaching more complex supramolecular architectures
may be expected in this area of research and utilization of
self-assembly phenomena as well.
7. Future directions in the application of reverse micelles
As far as the future directions in the application of reverse
micelles in the field of materials synthesis are concerned, the
following approaches may be outlined. Because of the emphasized
uniqueness of particular designed structures and compositions
within specific parent microemulsions, the development and
application of highly specific and growth- directing surfactants
especially suitable for particular chemical compositions, crystal
structures and intended morphologies may be expected in future
[103]. In any case, the future prosperity in the use of reverse
micelles and microemulsions for inducing practical self-assembly
phenomena depends on the combined synergetic efforts of application
of basic principles of colloid chemistry (mostly based on the
simple framework of DLVO theory), trial-and-error approaches,
employment of diverse advanced microscopy techniques, and
theoretical prediction of specific molecular recognition
effects.
Unlike ordinary emulsions, microemulsions do not require high shear
rates for their formation and may due to potential existence of
fine and diverse metastable colloid states exhibit a wide range of
inherent multi-molecular configurations [104], including either
regular or reverse micelles of various oval shapes (spherical,
cylindrical, rod-shaped), vesicular structures,
31V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
bilayer (lamellar) and cubic liquid crystals, columnar, mesh and
bicontinuous mesophases, cubosomes (dispersed bicontinuous cubic
liquid crystalline phase), sponge phases, hexagonal rod- like
structures, spherulites (radially arranged rod-like micelles),
multiphase–substructured configurations (such as water in oil in
water droplets, for example), highly percolated pearl like
structures, supra aggregates that comprise various substructured
combinations of microemulsion aggregates, as well as numer- ous
transient configurations. The application of complex non
equilibrium phases and such transient configurations between
reverse-micellar and various other inherent multi-molecular
self-assembled entities could provide the basis for growth of
numerous attractive novel morphologies. As a matter of fact, each
particular point in a microemulsion phase diagram corresponds to
specific local conditions for physico-chemical transformations that
take place therein and result in unique material structures
‘templated’ in each of these cases [78].
Combinations of reverse-micellar or any other microemul- sion
method of synthesis with other processing methods, including
hydrothermal synthesis [40], ultrasonic and UV irra- diation
[105,106], pH-shock wave method [107] and flame- spraying [108],
have been investigated. Merging of two or more preparation
techniques into one can due to synergy ef- fects lead to multiple
advantages, such as improved control of the stoichiometry of the
final product (as adopted from sol–gel method) and extremely fine
and controllable grain size (as acquired from reverse-micellar
synthesis) in the examples [109,110] of the combination of sol–gel
and reverse-micellar approaches to materials synthesis. Low yields,
surfactant- contamination and difficulties arising out of the
attempts to separate precipitated powders in the form of
non-agglomerated particles – serious obstacles of the
microemulsion-assisted nanoparticle preparation procedures – were
overcome by feedstocking flame-spraying apparatus with
nanoparticles together with their parent microemulsion [108], at
the same time transcending poor control of particle size and shape,
which is a typical drawback of the conventional flame-spraying
methods of synthesis. Combining reverse-micellar synthesis of
cobalt particles with their subsequent evaporating deposition in
external magnetic field led to the formation of large-scale 3D
superlattices of cobalt nanocrystals [111]. Silver nanorods
encapsulated by polystyrene were prepared by combination of
reverse-micellar, gas antisolvent, and ultrasound techniques [112],
whereby specific carbon nanotubes were prepared by direct
introduction of in situ prepared catalytically active Co and Mo
particles by a reverse-micellar method [113,114].
Both weak soft-tech potentials for structuring self-assembled
products into functional systems of hierarchical organization and
inherent limitations in the resolution of lithographic techniques
in nanostructural design may be overcome by constructive coupling
of the soft-tech production of fine- structured materials with
hard-tech assembly methodologies. Excellent achievements have been
recently reported by relying on such an approach of combining
‘bottom–up’ and ‘top– down’ methodologies [46,115–118].
Langmuir–Blodgett films [119], obtained by coupling self-assembled
orientation of molecules at air–water interfaces with a technique
for their
deposition on solid substrates, present a classic example of such
complementary synthesis/processing methodology. In that sense,
layer-by-layer (LbL) techniques comprising adsorption of oppositely
charged polyelectrolytes on a solid surface of synthesized
particles in reverse micelles were used to overcome difficulties
arising out of the inabilities to carry out sequential reactions
inside the same reverse micelles in order to obtain multilayered
composites [120].
The assembling of particles formed in the processes of
reverse-micellar and, in general, microemulsion-assisted syn-
theses into precisely tailored, supra-nanocrystalline 3D struc-
tures, presents an important challenge, whereas in situ reactions
in well-organized amphiphilic matrices present only one step
towards this goal [39]. Self-assembly parallelism and the selective
patterning precision of lithographic and etching techniques can be
united in a multitude of hybrid techniques for the production of
fine structures [121]. External fields, such as electric and
magnetic fields, heat gradients or single layer shearing, can
induce unexpected orderings depending on the intensities and
directions of the field relative to the suspension cell [122], and
may be used to hierarchically organize particles into 3D matrices.
On the other hand, electrospraying, electro- coalescence and other
methods that involve various external fields, may be used for
ink-jet spraying, fluid atomization, phase and particle separation,
thus improving the functionalizational control of the
self-assembled fine structures [123].
To sum up, reverse-micellar and other microemulsional systems can
provide complex interfaces that can support parallel reactions
leading to surprisingly complex outcomes, and their relatively
stable existence in thermodynamically metastable states can support
significant modifications of the product structures by the pure
influence of aging treatment. However, small yields obtained due to
employing extremely small concentrations and expensively complex
environments used in most of the cases, altogether with the fact
that increasing the space of options for production of various end
results via extremely fine variations of certain experimental
parameters comes at the price of increased sensitivity of initial
experimental settings that lead to reproducible outcomes, provide
implicit difficulties within such an approach to advancedmaterials
synthesis. The future prosperity of application of reverse
micelles, microemulsions and other self-assembling amphiphilic
matrices in advanced structural design will in large extent depend
on the successful global balancing of these pros and cons.
8. Conclusions
The presented results may suggest that the role that re- verse
micelles play in ‘parenting’ materials formation pro- cesses is
more intricate than purely ‘templating’ one. Reverse micelles have
been shown as capable of significantly mod- ifying the reaction
pathways that take place in their presence. Instead of being
considered as chemically inert nano-reactors, reverse micelles may
be regarded as complex multi-molecular entities that could be under
specific conditions actively engaged in the chemical pathways of
the formation of given materials.
32 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
The application of reverse micelles for materials synthe- sis
purposes could be, therefore, acknowledged in part as a
biomimicking approach to advanced structural design. The presented
examples of pronounced sensitivity of processes that employ
reverse-micellar effects in materials processing may initiate an
apprehension of their active physico-chemical role, presumably
similar to the primitive biological membranes. Reverse micelles can
arise as an important step for the practical field of colloid
science oriented towards reaching highly organized and
ultra-sensitive functional structures and ‘templating’ environ-
ments. The consequence of such convergence between biological
features and self-assembly design is that with increasing the
complexity of advanced nanofunctional devices, an increased
sensitivity of the intended products, manifested either as
irreproducibility of synthesis procedures or exceptional functional
sensitivity towards slightest environmental effects, will start
appearing as a significant problem. However, knowing that every
advantageous challenge always has its risky side as well, such an
intricate situation could be,with a lot of effort involved, turned
into an optional range of convenient and potentially fruitful
outcomes.
References
[1] Uskokovi V, Drofenik M. Synthesis of materials within reverse
micelles. Surf Rev Lett 2005;12(2):239–77.
[2] Boutonnet M, Kizling J, Stenius P. The preparation of
monodisperse colloidal metal particles from micro-emulsions.
Colloids Surf 1982;5(3): 209–25.
[3] Yadav OP, Palmqvist A, Cruise N, Holmberg K. Synthesis of
platinum nanoparticles in micoremulsions and their catalytic
activity for the oxidation of carbon monoxide. Colloids Surf A
Physicochem Eng Asp 2003;221:131–4.
[4] Chen F, Xu GQ, Hor TSA. Preparation and assembly of colloidal
gold nanoparticles in CTAB-stabilized reverse microemulsion. Mater
Lett 2003;57:3282–6.
[5] Barnickel P, Wokaun A, Sager W, Eickel HF. Size tailoring of
silver colloids by reduction in w/o microemulsions. J Colloid
Interface Sci 1992;148(1):80–90.
[6] Wang CC, Chen DH, Huang TC. Synthesis of palladium
nanoparticles in water-in-oil microemulsions. Colloids Surf A
Physicochem Eng Asp 2001;189:145–54.
[7] Wu ML, Chen DH, Huang TC. Synthesis of Au/Pd bimetallic
nanoparticles in reverse micelles. Langmuir 2001;17:3877–83.
[8] ZhangX, ChanKY.Water-in-oil microemulsions synthesis of
platinum— ruthenium nanoparticles, their characterization and
electrocatalytic properties. Chem Mater 2003;15:451–9.
[9] Agostiano A, Catalano M, Curri ML, Monica MD, Manna L,
Vasanelli L. Synthesis and Structural characterisation of CdS
nanoparticles prepared in a four-components ‘water-in-oil’
microemulsion. Micron 2000;31: 253–8.
[10] Ward AJI, O'Sullivan EC, Rang JC, Nedeljkovi J, Patel RC. The
synthesis of quantum-size lead sulfide particles in
surfactant-based complex fluid media. J Colloid Interface Sci
1993;161(2):316–20.
[11] Murray CB, Norris DJ, Bawendi MG. Synthesis and
characterization of nearly monodisperse CdE (E = S, Se, Te)
semiconductor nanocrystallites. J Am Chem Soc
1993;115(19):8706–15.
[12] Fang X, Yang C. An experimental study on the relationship
between the physical properties of CTAB/hexanol/water reverse
micelles and ZrO2– Y2O3 nanoparticles prepared. J Colloid Interface
Sci 1999;212:242–51.
[13] Zarur AJ, Ying JY. Reverse microemulsion synthesis of
nanostructured complex oxides for catalytic combustion. Nature
2000;403:65–7.
[14] Giannakas AE, Vaimakis TC, Ladavos AK, Trikalitis PN, Pomonis
PJ. Variation of surface properties and textural features of spinel
ZnAl2O4
and perovskite LaMnO3 nanoparticles prepared via CTAB–butanol–
octane–nitrate salt microemulsion in the reverse and bicontinuous
states. J Colloid Interface Sci 2003;259:244–53.
[15] Sun L, Zhang Y, Zhang J, Yan C, Liao C, Lu Y. Fabrication of
size controllable YVO4 nanoparticles via microemulsion-mediated
synthetic process. Solid State Commun 2002;124:35–8.
[16] Gan LM, Zhang LH, Chan HSO, Chew CH, Loo BH. A novel method
for the synthesis of perovskite-type mixed metal oxides by the
inverse microemulsion technique. J Mater Sci 1996;31:1071–9.
[17] Bae DS, Han KS, Adair JH. Synthesis and microstructure of
Pd/SiO2 nanosized particles by reverse micelle and sol–gel
processing. J Mater Chem 2002;12(10):3117–20.
[18] Wu M, Zhang YD, Hui S, Xiao TD, Ge S, Hines WA, et al.
Structure and magnetic properties of SiO2-coated Co nanoparticles.
J Appl Phys 2002;92(1):491–5.
[19] Bae DS, Han KS, Adair JH. Synthesis of platinum/silica
nanocomposites by reverse micelle and sol–gel processing. J Am
Ceram Soc 2002;85(5): 1321–3.
[20] Lin J, Zhou W, Kumbhar A, Wiemann J, Fang J, Carpenter EE, et
al. Gold-coated iron nanoparticles: synthesis, characterization,
and mag- netic field-induced self-assembly. J Solid State Chem
2001;159: 26–31.
[21] O'Connor CJ, Sims JA, Kumbhar A, Kolesnichenko VL, Zhou WL,
Wiemann JA. Magnetic properties of FePtx/Au and CoPtx/Au core-shell
nanoparticles. J Magn Magn Mater 2001;226–230:1915–7.
[22] Hammouda A, Gulik T, Pileni MP. Synthesis of nanosize latexes
by reverse micelle polymerization. Langmuir 1995;11:3656–9.
[23] Fang J, Stokes L, Wiemann J, Zhou W. Nanocrystalline bismuth
synthesized via in situ polymerization — microemulsion process.
Mater Lett 2000;42:113–20.
[24] Summers M, Eastoe J, Davis S. Formation of BaSO4 nanoparticles
in microemulsions with polymerised surfactant shells. Langmuir
2002;18: 5023–6.
[25] Lopez-Quintela MA, Rivas J, Blanco MC, Tojo C. Synthesis of
nanoparticles in microemulsions. In: Marzan LML, Kamat PV, editors.
Nanoscale materials. Kluwer Academic Plenum Publ; 2003. p.
135–55.
[26] Ayyub P, Multani MS. Secondary recrystallization during
sintering of YBa2Cu3O7−δ derived from water-in-oil microemulsion.
Mater Lett 1991;10(9–10):431–6.
[27] Li F, Vipulanandan C. Production and characterization of YBCO
nanoparticles. IEEE Trans Appl Supercond 2003;13(2):3196–8.
[28] Levinger NE. Water in confinement. Science 2002;298:1722–3.
[29] Pileni MP, Zemb T, Petit C. Solubilization by reverse
micelles: solute
localization and structure perturbation.ChemPhysLett 1985;118(4):
414–20. [30] Pileni MP. Reverse micelles as microreactors. J Phys
Chem 1993;97:
6961–73. [31] Carpenter EE, Seip CT, O'Connor CJ. Magnetism of
nanophase metal and
metal alloy particles formed in ordered phases. J Appl Phys
1999;85(8): 5184–6.
[32] Sjöblom J, Lindberg R, Friberg SE. Microemulsions-phase
equilibria characterization, structures, applications and chemical
reactions. Adv Colloid Interface Sci 1996;95:125–287.
[33] Dresco PA, Zaitsev VS, Gambino RJ, Chu B. Preparation and
properties of magnetite and polymer magnetite nanoparticles.
Langmuir 1999;15: 1945–51.
[34] Arriagada FJ, Osseo-Asare K. Synthesis of nanosize silica in a
nonionic water-in-oil microemulsion: effects of the
water/surfactant molar ratio and ammonia concentration. J Colloid
Interface Sci 1999;211:210–20.
[35] Hua R, Zang C, Shao C, Xie D, Shi C. Synthesis of barium
fluoride nanoparticles from microemulsion. Nanotechnology
2003;14:588–91.
[36] Yener DO, Giesche H. Synthesis of pure and manganese-,
nickel-, and zinc-doped ferrite particles in water-in-oil
microemulsions. J Am Ceram Soc 2001;84(9):1987–95.
[37] Roux D, Bellocq AM, Bothorel P. Effect of the molecular
structure of components on micellar interactions in microemulsions.
Progr Colloid & Polym Sci 1984;69:1–11.
[38] Liu J, Kim AY, Wang LQ, Palmer BJ, Chen YL, Bruinsma P, et al.
Self- assembly in the synthesis of ceramic materials and
composites. Adv Colloid Interface Sci 1996;69:131–80.
33V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
[39] John VT, Simmons B, McPherson GL, Bose A. Recent developments
in materials synthesis in surfactant systems. Curr Opin Colloid
Interface Sci 2002;7:288–95.
[40] Lin JC, Dipre JT, Yates MZ. Microemulsion-directed synthesis
of molecular sieve fibers. Chem Mater 2003;15:2764–73.
[41] Hines JD. Theoretical aspects of micellisation in surfactant
mixtures. Curr Opin Colloid Interface Sci 2001;6:350–6.
[42] Chandler D. Interfaces and the driving force of hydrophobic
assembly. Nature 2005;437:640–7.
[43] Natarajan U, Handique K,Mehra A, Bellare JR, Khilar KC.
Ultrafine metal particle formation in reverse micellar systems:
effects of intermicellar exchange on the formation of particles.
Langmuir 1996;12: 2670–8.
[44] Lindman B, Wennerstrom H. Micelles. Amphiphile aggregation in
aqueous solution. Topics in current chemistry, vol. 87. Berlin:
Springer- Verlag; 1980.
[45] Gambi CMC, Carla M, Senatra D. Comments on a 'self-assembly in
fluorocarbon surfactant systems. Curr Opin Colloid Interface Sci
1999;4(1): 88–9.
[46] Soten I, Ozin GA. New directions in self-assembly: materials
synthesis over ‘all’ length scales. Curr Opin Colloid Interface Sci
1999;4:325–37.
[47] Antonietti M. Surfactants for novel templating applications.
Curr Opin Colloid Interface Sci 2001;6:244–8.
[48] Holmberg K. Surfactant-templated nanomaterials synthesis. J
Colloid Interface Sci 2004;274:355–64.
[49] Filankembo A, Giorgio S, Lisiecki I, Pileni MP. Is the anion
the major parameter in the shape control of nanocrystals. J Phys
Chem B 2003;107 (30):7492–500.
[50] Lisiecki I. Size control of spherical metallic nanocrystals.
Colloids Surf A Physicochemi Eng Asp 2004;250(1–3):499–507.
[51] Fowler CE, Li M, Mann S, Margolis HC. Influence of surfactant
assembly on the formation of calcium phosphate materials— a model
for dental enamel formation. J Mater Chem
2005;15(32):3317–25.
[52] Lopez-Quintela MA. Synthesis of nanomaterials in
microemulsions: formation mechanisms and growth control. Curr Opin
Colloid Interface Sci 2003;8:137–44.
[53] Uskokovi V, Drofenik M. A mechanism for the formation of
nanostructured NiZn ferrites via a microemulsion-assisted
precipitation method. Colloids Surf A Physicochem Eng Asp
2005;266:168–74.
[54] Inouye K, Endo R, Otsuka Y, Miyashiro K, Kaneko K, Ishikawa T.
Oxygenation of ferrous-ions in reversed micelle and reversed micro-
emulsion. J Phys Chem 1982;86(8):1465–9.
[55] Lopez-Quintela MA, Tojo C, Blanco MC, Rio LG, Leis JR.
Microemul- sion dynamics and reactions in microemulsions. Curr Opin
Colloid Interface Sci 2004;9:264–78.
[56] Kurihara K. Nanostructuring of liquids at solid–liquid
interfaces. Progr Colloid & Polym Sci 2002;121:49–56.
[57] Zhong Q, Steinhurst DA, Carpenter EE, Owrutsky JC. Fourier
transform infrared spectroscopy of azide ion in reverse micelles.
Langmuir 2002;18:7401–8.
[58] Linehan, JC, Fulton, JL, Bean, RM. Process of Forming
Compounds Using Reverse Micelle or Reverse Microemulsion Systems.
US Patent 5,770,172; 1998.
[59] Zana R, Lang J. Dynamics of microemulsions. In: Friberg SE,
Bothorel P, editors. Microemulsions: structure and dynamics. Boca
Raton: CRC Press; 1987. p. 153–72.
[60] Ninham BW. On progress in forces since the DLVO theory. Adv
Colloid Interface Sci 1999;83:1–17.
[61] Vogler EA. Structure and reactivity of water at biomaterial
surfaces. Adv Colloid Interface Sci 1998;74:69–117.
[62] Makovec D, Košak A, Drofenik M. The preparation of
MnZn–ferrite Nanoparticles In Water–CTAB–hexanol microemulsion.
Nanotechnolo- gy 2004;15:S160–6.
[63] Uskokovi V, Drofenik M, Ban I. The characterization of
nanosized nickel–zinc ferrites synthesized within reverse micelles
of CTAB/1- hexanol/water microemulsion. J Magn Magn Mater
2004;284:294–302.
[64] Uskokovi V, Drofenik M. Synthesis of nanocrystalline
nickel–zinc ferrites via a microemulsion route. Mater Sci Forum
2004;453–454: 225–30.
[65] Santra S, Tapec R, Theodoropoulou N, Dobson J, Hebard A, Tan
W. Synthesis and characterization of silica-coated iron oxide
nanoparticles in microemulsion: the effect of nonionic surfactants.
Langmuir 2001;17: 2900–6.
[66] Nooney RI, Thirunavukkarasu D, Chen Y, Josephs R, Ostafin AE.
Synthesis of nanoscale mesoporous silica spheres with controlled
particle size. Chem Mater 2002;14:4721–8.
[67] Debuigne F, Jeunieau L, Wiame M, Nagy JB. Synthesis of organic
nano- particles in different w/o microemulsions. Langmuir
2000;16:7605–11.
[68] Liu C, Zuo B, Rondinone AJ, Zhang ZJ. Reverse micelle
synthesis and characterization of superparamagnetic MnFe2O4 spinel
ferrite nanocrys- tallites. J Phys Chem 2000;104(6):1141–5.
[69] Matijevi E. Colloid science of ceramic powders. Pure Appl Chem
1988;60(10):1479–91.
[70] Bunker CE, Harruff BA, Pathak P, Payzant A, Allard LF, Sun YP.
Formation of cadmium sulfide nanoparticles in reverse micelles:
extreme sensitivity to preparation procedure. Langmuir
2004;20:5642–4.
[71] Hines JD. Theoretical aspects of micellisation in surfactant
mixtures. Curr Opin Colloid Interface Sci 2001;6:350–6.
[72] Bagwe RP, Khilar KC. Effects of intermicellar exchange rate on
the formation of silver nanoparticles in reverse microemulsions of
AOT. Langmuir 2000;16:905–10.
[73] Li F, Li GZ, Wang HQ, Xue QJ. Studies on
cetyltrimethylammonium bromide (CTAB) micellar solution and CTAB
reversed microemulsion by ESR and 2H NMR. Colloids Surf A
Physicochem Eng Asp 1997;127: 89–96.
[74] Ghosh KK, Tiwary LK. Microemulsions as reaction media for a
hydrolysis reaction. J Dispers Sci Technol 2001;22(4):343–8.
[75] Shervani Z, Ikushima Y. Compartmentalization of reactants in
supercrit- ical fluid micelles. J New Chem 2002;26:1257–60.
[76] Robbins H. The preparation of Mn–Zn ferrites by
co-precipitation. Proceedings of the 3rd International Conference
on Ferrites. Japan: Center for Academic Publications; 1981. p.
7–10.
[77] Todaka Y, Nakamura M, Hattori S, Tsuchiya K, Umemoto M.
Synthesis of ferrite nanoparticles by mechanochemical processing
using a ball mill. Mater Trans 2003;44(2):277–84.
[78] Uskokovi V, Drofenik M. Synthesis of nanocrystalline
nickel–zinc ferrites within reverse micelles. Mater Technol
2003;37(3–4):129–31.
[79] Uskokovi V, Drofenik M. Synthesis of relatively highly
magnetic nano- sized NiZn–ferrite in microemulsion at 45 °C. Surf
Rev Lett 2005;12 (1):97–100.
[80] Uskokovi V. Nanotechnologies: what we do not know. Technol Soc
2007;29(1):43–61.
[81] Uskokovi V, Drofenik M. Four novel co-precipitation procedures
for the synthesis of lanthanum–strontium manganites. Mater Design
2007;28 (2):667–72.
[82] Uskokovi V, Makovec D, Drofenik M. Synthesis of
lanthanum–strontium manganites by a hydroxide-precursor
co-precipitation method in solution and reverse micellar
microemulsion. Mater Sci Forum 2005;494:155–60.
[83] Uskokovi V, Drofenik M. Synthesis of lanthanum–strontium
manga- nites by oxalate-precursor co-precipitation methods in
solution and in reverse micellar microemulsion. J Magn Magn Mater
2006;303 (1):214–20.
[84] Bagwe RP, Khilar KC. Effects of intermicellar exchange rate on
the formation of silver nanoparticles in reverse microemulsions of
AOT. Langmuir 2000;16:905–10.
[85] Jung HM, Price KE, McQuade DT. Synthesis and characterization
of cross-linked reverse micelles. J Am Chem Soc
2003;125(18):5351–5.
[86] Carlile K, Rees GD, Robinson BH, Steer TD, Svensson M. Lipase-
catalysed interfacial reactions in reverse micellar systems— role
of water and microenvironment in determining enzyme activity or
dormancy. J Chem Soc Faraday Trans 1996;92(23):4701–8.
[87] Chen YX, Zhang XZ, Zheng K, Chen SM, Wang QC, Wu XX. Protease-
catalyzed synthesis of precursor dipeptides of RGD with reverse
micelles. Enzyme Microb Technol 1998;23(3–4):243–8.
[88] Sun PP, Zhang YM. The catalysis and asymmetric induction of
chiral reverse micelle: synthesis of optically active alpha-amino
acids. Synth Commun 1997;27(23):4173–9.
34 V. Uskokovi, M. Drofenik / Advances in Colloid and Interface
Science 133 (2007) 23–34
[89] Lee KKB, Poppenborg LH, Stuckey DC. Terpene ester production
in a solvent phase using a reverse micelle-encapsulated lipase.
Enzyme Microb Technol 1998;23(3–4):253–60.
[90] Babu CR, Flynn PF, Wand AJ. Preparation, characterization, and
NMR spectroscopy of encapsulated proteins dissolved in low
viscosity fluids. J Biomol NMR 2003;25(4):313–23.
[91] Melo EP, Aires-Barros MR, Cabral JM. Reverse micelles and
protein biotechnology. Biotechnol Annu Rev 2001;7:87–129.
[92] Melo EP, Costa SMB, Cabral JMS, Fojan P, Petersen SB. Cutinase
— SOT interactions in reverse micelles: the effect of 1-hexanol.
Chem Phys Lipids 2003;124:37–47.
[93] Chang GG, Huang TM, Hung HC. Reverse micelles as
life-mimicking systems. Proc Natl Sci Counc ROC(B)
1999;24(3):89–100.
[94] Bachmann PA, Walde P, Luisi PL, Lang J. Self-replicating
micelles — aqueous micelles and enzymatically driven reactions in
reverse micelles. J Am Chem Soc 1991;113(22):8204–9.
[95] Bachmann PA, Luisi PL, Lang J. Self-replicating reverse
micelles. Chimia 1991;45(9):266–8.
[96] Engelman DM. Membranes are more mosaic than fluid. Nature
2005;438:578–80.
[97] Hyde ST, Schröder GE. Novel surfactant mesostructural
topologies: between lamellae and columnar (hexagonal) forms. Curr
Opin Colloid Interface Sci 2003;8:5–14.
[98] Piazza R. Cradles for life: self-organization of biological
colloids. Curr Opin Colloid Interface Sci 2006;11(1):30–4.
[99] Levine IN. Physical chemistry. New York: McGraw-Hill; 1978.
[100] Pollack GH. The role of aqueous interfaces in the cell. Adv
Colloid
Interface Sci 2003;103:173–96. [101] Baglioni P, Berti D. Self
assembly in micelles combining stacking and
H-bonding. Curr Opin Colloid Interface Sci 2003;8:55–61. [102]
Tolles WM. Self-assembled materials. MRS Bull 2000;25(10):36–8.
[103] Berti D. Self assembly of biologically inspired amphiphiles.
Curr Opin
Colloid Interface Sci 2006;11:74–8. [104] Hellweg T. Phase
structures of microemulsions. Curr Opin Colloid
Interface Sci 2002;7:50–6. [105] Huang J, Xie Y, Li B, Liu Y, Lu J,
Qian Y. Ultrasound-induced formation
of CdS nanostructures in oil-in-water microemulsions. J Colloid
Interface Sci 2001;236:382–4.
[106] Mallick K, Wang ZL, Pal T. Seed-mediated successive growth of
gold particles accomplished by UV irradiation: a photochemical
approach for size-controlled synthesis. J Photochem Photobiol A
Chem 2001;140 (1):75–80.
[107] Koumoulidis GC, Katsouilidis AP, Ladavos AK, Pomonis PJ,
Trapalis CC, Sdoukos AT, et al. Preparation of hydroxyapatite via
microemulsion route. J Colloid Interface Sci
2003;259(2):254–60.
[108] Bonini M, Bardi U, Berti D, Neto C, Baglioni P. A new way to
prepare nanostructured materials: flame spraying of microemulsions.
J Phys Chem B 2002;106:6178–83.
[109] Zarur AJ, Hwu HH, Ying JY. Reverse microemulsion-mediated
synthesis and structural evolution of barium hexaaluminate
nanoparticles. Lang- muir 2000;16:3042–9.
[110] Althues H, Kaskel H. Sulfated zirconia nanoparticles
synthesized in reverse microemulsions: preparation and catalytic
properties. Langmuir 2002;18:7428–35.
[111] Legrand J, Ngo AT, Petit C, Pileni MP. Domain shapes and
superlattices made of cobalt nanocrystals. Adv Mater
2001;13(1):58–62.
[112] Zhang JL, Liu ZM, Han BX, Jiang T, Wu WZ, Chen J, et al.
Preparation of polystyrene-encapsulated silver nanorods and
nanofibers by combi- nation of reverse micelles, gas antisolvent,
and ultrasound techniques. J Phys Chem B 2004;108(7):2200–4.
[113] Ago H, Ohshima S, Uchida K, Komatsu T, Yumura M. Carbon
nanotube synthesis using colloidal solution of metal nanoparticles.
Physica B 2002;323:306–7.
[114] Ago H, Ohshima S, Tsukuagoshi K, Tsuji M, Yumura M. Formation
mechanism of carbon nanotubes in the gas-phase synthesis from
colloidal solutions of nanoparticles. Curr Appl Phys
2005;5(2):128–32.
[115] Kraus T, Malaquin L, Delamarche E, Schmid H, Spencer ND, Wolf
H. Closing the gap between self-assembly and microsystems using
self- assembly, transfer, and integration of particles. Adv Mater
2005;17(20): 2438–42.
[116] Frenot A, Chronakis IS. Polymer nanofibers assembled by
electrospin- ning. Curr Opin Colloid Interface Sci
2003;8:64–75.
[117] Tzeng SD, Lin KJ, Hu JC, Chen LJ, Gwo S. Templated
self-assembly of colloidal nanoparticles controlled by
electrostatic nanopatterning on Si3N4/SiO2/Si Electret. Adv Mater
2006;18(9):1147–51.
[118] Barth JV, Costantini G, Kern K. Engineering atomic and
molecular nanostructures at surfaces. Nature 2005;437:671–9.
[119] Bourgeat-Lami E. Organic–inorganic nanostructured colloids. J
Nanosci Nanotechnol 2002;2(1):1–24.
[120] Shchukin DG, Sukhorukov GB. Nanoparticle synthesis in
engineered organic nanoscale reactors. Adv Mater
2004;16(8):671–82.
[121] Mendes PM, Preece JA. Precision chemical engineering:
integrating nanolithography and nanoassembly. Curr Opin Colloid
Interface Sci 2004;9:236–48.
[122] Arora AK, Tata BVR. Interactions, structural ordering and
phase transitions in colloidal dispersions. Adv Colloid Interface
Sci 1998;78: 49–97.
[123] Shin WT, Yiacoumi S, Tsouris C. Electric-field effect on
interfaces: electrospray and electrocoalescence. Curr Opin Colloid
Interface Sci 2004;9:249–55.
Reverse micelles: Inert nano-reactors or physico-chemically active
guides of the capped reactio.....
Introduction
The need to reevaluate the functional representation of reverse
micelles
Examples of chemical ‘butterfly effects’ in reverse
micelle-assisted syntheses
The example of nickel–zinc ferrite
The example of lanthanum–strontium manganite
Correlations with the biological context
Future directions in the application of reverse micelles
Conclusions
References