synthesis of materials within reverse micelles
Transcription
synthesis of materials within reverse micelles
June 2, 2005 12:33 00700 Surface Review and Letters, Vol. 12, No. 2 (2005) 239–277 c World Scientific Publishing Company SYNTHESIS OF MATERIALS WITHIN REVERSE MICELLES VUK USKOKOVIĆ∗ and MIHA DROFENIK† Jožef Stefan Institute, Jamova 38, 1000 Ljubljana, Slovenia † Faculty of Chemistry and Chemical Engineering, Smetanova 17, 2000 Maribor, Slovenia ∗ [email protected] Received 1 March 2005 Reverse micelles as nanosized aqueous droplets existing at certain compositions of water-in-oil microemulsions are widely used today in the synthesis of many types of nanoparticles. However, without a rich conceptual network that would correlate the properties and compositions of reverse micellar microemulsions to the properties of to-be-obtained particles, the design procedures in these cases usually rely on a trial-and-error approach. As like every other science, what is presently known is merely the tip of the iceberg compared to the uninvestigated vastness still lying below. The aim of this article is to present readers with most of the major achievements from the field of materials synthesis within reverse micelles since the first such synthesis was performed in 1982 until today, to possibly open up new perspectives of viewing the typical problems that nowadays dominate the field, and to hopefully initiate the observation and generation of their actual solutions. We intend to show that by refining the oversimplified representations of the roles that reverse micelles play in the processes of nanoparticles synthesis, steps toward a more complex and realistic view of the concerned relationships can be made. The first two sections of the review are of introductory character, presenting the reader with the basic concepts and ideas that serve as the foundations of the field of reverse micellar synthesis of materials. Applications of reverse micelles, other than as media for materials synthesis, as well as their basic structures and origins, together with experimental methods for evaluating their structural and dynamic properties, basic chemicals used for their preparation and simplified explanations of the preparation of materials within, will be reviewed in these two introductory sections. In Secs. 3 and 4, we shall proceed with reviewing the structural and dynamic properties of reverse micelles, respectively, assuming that knowledge of both static and dynamic parameters of microemulsions and changes induced thereof, are a necessary step prior to putting forth any correlations between the parameters that define the properties of microemulsions and the parameters that define the properties of materials synthesized within. Typical pathways of synthesis will be presented in Sec. 5, whereas basic parameters used to describe correlations between the properties of microemulsion reaction media and materials prepared within, including reagent concentrations, ionic strength, temperature, aging time and some of the normally overlooked influences, will be mentioned in Sec. 6. The whole of Sec. 7 is devoted to reviewing water-to-surfactant molar ratio as the most often used parameter in materials design by performing reverse micellar synthesis routes. The mechanisms of particle formation within precipitation synthesis in reverse micelles is discussed in Sec. 8. Synthesis of composites, with special emphasis on silica composites, is described in Sec. 9. All types of materials, classified according to their chemical compositions, that were, to our knowledge, synthesized by using reverse micelles up-to-date, will be briefly mentioned and pointed to the corresponding references in Sec. 10. In Sec. 11, some of the possible future directions for the synthesis of nanostructured materials within reverse micelles, found in combining reverse micellar syntheses ∗ Corresponding author. 239 June 2, 2005 12:33 00700 240 V. Uskoković & M. Drofenik and various other synthesis procedures with the aim of reaching self-organizing nanoparticle systems, will be outlined. Keywords: Nanomaterials; reverse micelles; review; synthesis. 1. Introduction Reverse micelles as nanoscale hydrophilic cavities of microemulsions have been known since the 1960s, but these diverse multimolecular structures were for the first time used as nano-templates for materials synthesis (of monodispersed Pt, Pd, Rh and Ir particles) in 1982.1 After these pioneering researches, many different materials comprising de-agglomerated and monodispersed particles (Fig. 1) have been prepared2–6 by using reverse micelles. Reverse micellar synthesis of materials belongs to the class of wet materials synthesis procedures, and exhibits, in general, all the advantages that usually accompany other wet approaches to materials synthesis. Excellent control of the final powders’ stoichiometries with possibilities of obtaining homogeneity and mixing on the atomic scale, narrow particle sizes distributions, negligible contamination of the product during the homogenization of the starting compounds, low energy consumption, low aging times and simple equipment, are some of the ordinary qualities of wet syntheses, especially when compared to high-temperature traditional routes for the synthesis of common ceramic and metallic materials. Improved control of the particle sizes, shapes, uniformity and dispersity are additional general advantages of reverse micellar synthesis compared to other, bulk wet approaches. Materials resulting from wet powder preparation might have extremely small sizes, which implies a number of potential advantages, such as lower sintering temperatures in case of the preparation of ceramic materials. In cases where high-temperature treatment cannot be completely avoided by roomtemperature aging procedures in reverse micelles, the formation of nanoparticles with high specificsurface area can enable calcination temperatures to be set lower compared to traditional, solid-state approaches. Organized self-assembled surfactant phases have recently received a lot of attention as reaction and templating media, but have for a long time been used (a) (b) Fig. 1. (a) La–Ni oxalate nanoparticles5 and (b) silica particles6 as obtained by reverse micellar synthesis in microemulsion. for many other purposes: • As wash and cleaning systems — since they facilitate the solubilization of both hydrophobic and hydrophilic components at the same time and at reduced temperatures. • As separators — due to selective solubilization of certain molecules. • As lubrication compounds. • As pharmaceuticals. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles • As fuels — due to improved fuel atomization and evaporation of water which increases the heat and temperature of combustion. • As a medium in crude oil exploitation — due to their surface activity. • As catalyzers7,8 or inhibitors9 of biochemical enzyme-driven reactions — due to compartmentalization of reactants and products as well as changes in activity and substrate specificity of enzymes due to alteration in solubilized enzymes conformations when accommodating the micellar interior structure. • In preparative organic chemistry, in order to overcome reactant solubility problems due to the ability of microemulsions to solubilize both polar and nonpolar substances and to compartmentalize and concentrate reactants.10 • For storing bioactive chemical reagents.11 • As a cell membrane-mimetic medium for the study of membrane interactions of bioactive peptides. Encapsulating 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 novel strategy for studying proteins in a variety of contexts.12 Since it was noted that denaturation of proteins can be prevented in reverse micelles,13 these self-organized multimolecular assemblies have been used as lifemimicking systems,14 and as such, have received large interest after the proposition of hypotheses that self-replicating biochemical reactions of primordial planetary life were initiated in reverse micelles made of glycerine molecules, palmitic, stearic and oleic acid, which existed at air–water interfaces close to the shores of some ancient seas.15 Treating reverse micelles as active and even as the most basic structures of life was especially emphasized after the discovery of the possibility of initiating self-replication of reverse micelles due to a reaction ocurring within micellar structures.16,17 2. Basic View of Reverse Micelles, Microemulsions and Their Potentialities Within Materials Synthesis Reverse micelles exist at certain compositional range of water-in-oil microemulsions. Microemulsions — a term coined by J. H. Schulman in 195918 — are transparent thermodynamically stable dispersions 241 of two immiscible liquids containing appropriate amounts of surfactant. Amphiphilic substances, to which surfactants belong, possess significantly distanced hydrophilic and hydrophobic parts within their molecules, and, therefore, each of the two parts of the surfactant molecule has its preferent solvent. Two mutually immiscible components are usually water and an alcohol, hydrocarbon or, lately, environmentally-viable supercritical carbon dioxide,19,20 which proves to be of extreme convenience due to solving the problems of difficult separation and removal of solvent from products in conventional reverse micellar syntheses. Surfactant monolayers separate water and oil domains and hence reduce the unfavorable oil–water contact. In contrast to macroscopic emulsions which are thermodynamically unstable, nanosized microemulsion droplets are formed spontaneously and, although the reverse micellar systems are heterogenous on a molecular scale, they are equilibrium phases and are thus thermodynamically stable. Since the interactions between polar head-groups of the surfactant molecule and the interactions between the nonpolar tails favor only aggregates of a very specific size and molecular configuration, microemulsions typically have narrow droplet size distributions. The droplet uniformity of microemulsions is especially important, having a direct effect on the distribution of resulting particle sizes during precipitation reactions. Colloidal particle formation is a complex process, which involves interplay between nucleation, nanocrystal formation, intermediate growth as well as eventual coagulation and flocculation, that all depend on the specific interactions between ionic and molecular species within microemulsion. The success of the reverse micellar procedure for materials synthesis is closely related to the fact that particle nucleation can be initiated simultaneously at a large number of locations within reverse micelles, with the nucleation sites well isolated from each other due to the presence of surfactant films that may act as stabilizers of the formed particles. Normally, monodisperse particles are formed only when the nucleation and growth stages are strictly separated, which is a property of reverse micelle material synthesis due to the uniform nanodroplet structure and specific intermicellar interactions. Reverse micelle phases are tiny droplets of water, encapsulated by surfactant molecules and thus June 2, 2005 12:33 00700 242 V. Uskoković & M. Drofenik physically separated from the oil phase (Fig. 2). Simplified representation of the reverse micellar preparation of particles takes that aqueous “pools” of the reverse micelles act as nanoreactors for performing simple reactions of synthesis, and that the sizes of the microcrystals of the product are directly determined by the sizes of these pools.21–23 It is possible to control the sizes of reverse micelles by controlling the parameter w, defined as molar ratio of water-tosurfactant. The higher the w, the larger the water pools of the micelles and the nanoparticles formed within, and vice versa. Although such a correspondence between the size of the synthesized nanoparticle and the parameter w was approved in many experiments, it has been put into question lately, since within a number of performed syntheses a (a) (b) Fig. 2. A drawing of (a) a reverse micelle and (b) a more realistic model of reverse micelle.35 Blue spheres represent surfactant’s head-groups, whereby smaller yellow spheres denote counterions. Note that the surfactant head groups do not completely shield the aqueous interior of the modelled reverse micelle (b). similar correspondence could not be established. Therefore, the dynamic interaction among micelles has since lately been generally considered as the most important factor that influences the morphologies and properties of the final products.24 Particle size control results from regulation of the nucleation site size and the particle growth rate. Within reverse micellar materials synthesis, these two factors are often implicitly conveyed to reverse micelle sizes and material exchange between reverse micelles, respectively. It has been observed that many properties of the synthesized powders can be controlled, and thus designed by using proper conditions, regarding primarily the composition of a parent microemulsion. Control over quantum states of the particles or interparticle spacing can lead to novel mesoscopic properties of materials, which are sometimes very different from those of their atomic and bulk counterparts.25,26 Not only it is possible to synthesize nanosized uniform particles, but the obtained particles might also be extremely well dispersed,2 which is a key property for some applications. Even though reaction kinetics are neglected in many models since the intermicellar exchange is slow and governs the growth of particles, enhancement of reaction rates is also one of the known possible advantages of micellar synthesis routes.24,27 Since nanosized reverse micelles cannot yet be directly observed in their dynamic interactions within a nonpolar medium, indirect techniques are usually applied in order to evaluate certain, both static and dynamic properties of the micelles. In the approximations and different implicit pre-suppositions of various such techniques lie the answer to the sometimes pronounced impossibility of matching28 the concluded properties attributed to the same systems by using different experimental methods. Various indirect experimental methods may be used for the evaluation of the microemulsions’ phase diagrams, but generally, transparency measurements, conductivity measurements29 and various dynamic spectroscopy methods are the most often used. Scattering methods have in general become standard methods for probing microemulsions due to nondestructivity and versatility of such approaches. Since reverse microemulsion systems typically have June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles a high degree of optical clarity, the transition from the translucent, bright color of microemulsion to a turbid, opaque or viscous white solution might visually (with the naked eye) be interpreted as the breakdown of the microemulsion.30,31 However, it was shown that the solubilization limits obtained by titration are different from those obtained by contacting the organic phase with an excess aqueous phase,32 and such a difference was ascribed to the fact that in the titration method, the composition of water pools is fixed by the composition of the titrant, while in the contacting method, the composition of the water pools depends on the exchange of ions between the reverse micelles and the excess aqueous phase.32 Conductivity measurements are often used to routinely characterize microemulsion systems and determine the maximum amount of water that can be introduced into the system with maintaining the given water-in-oil system33 in stable condition. Upon addition of water into the system, conductivity increases due to percolation of charges through the droplet clusters, all until continuous introduction of water into the system makes the microemulsion unstable and eventually induces phase separation resulting in sharp decrease in conductivity. Since reverse micelles are most often regarded as “nano-templates” or “nano-reactors” for the synthesis of nanosized particles, special emphasis in the field of their investigation is, from the point of view of materials synthesis, placed on the attempts to determine their structural parameters, that is primarily micellar sizes and shapes as well as spatial distributions of these parameters. Beside many attempts to model reverse micellar structures,34–36 initiated by the pioneering attempts by Brown and Clarke,37 within indirect experimental methods that are regularly used in order to determine or evaluate different properties of reverse micelles, are included: light scattering (LS)38 and small angle neutron scattering (SANS) studies,39–42 small angle X-ray scattering (SAXS),41 quasi-elastic light scattering (QELS),43 infrared spectroscopy, including FTIR,44,45 picosecond IR pump-probe spectroscopy46 and near-IR spectroscopy,11 dielectric spectroscopy,47,48 fluorescence light and visible light scattering measurements,49 X-ray scattering techniques,50,51 ESR spectroscopy,52 freeze-fracture 243 etching transmission electron microscopy,53 53,54 NMR, measurements of diffusion coefficients,55 conductometric measurements29,33,56,57 spectrophotometric measurements,57,58 shear and extensional viscosity measurements,59–61 photon correlation spectroscopy,62 fluorescence quenching measurements,63 dynamic light scattering,46 Raman spectroscopy,64 and DSC measurements.45 In general, spectroscopy has been a primary tool for investigating the structures of reverse micelles. Combined with time-resolved methods, the spectroscopic methods can also yield dynamic information.35 The principle factors for explaining structural changes in microemulsions are surfactant shape, entropy, energy terms, as well as solvent properties such as ionic force and pH. Some of the most important quantities used for defining surfactants activities within microemulsion systems are: critical micelle concentration (CMC) — defined as the minimal concentration of surfactant molecules above which micelles are formed; and aggregation number — the number of surfactant molecules per micelle. The phase diagram of CTAB/1-hexanol/water microemulsion is shown in Fig. 3.65 The balance between electrostatic or polar interactions, geometry packing factor and topology mostly determine the geometry of the colloidal structure formed. The region denoted as L2 in Fig. 3, typical of its low water-to-oil phase ratio, belongs to the phase compositions at which reverse micelles exist as multimolecular structures that the microemulsion comprises. Beside reverse micellar structures, membraneous structures, bilayer (lamellar) and cubic liquid crystals, sponge phases, hexagonal rod-like structures, and uninverted (regular) micellar structures can be formed by varying the composition of the given, most often three- or four-component microemulsion system. The general remark is that microemulsion systems are very difficult to treat when the number of their components exceeds three. The use of normal micelles,66–68 bicontinuous structures,69,70 water-inoil-in-water pseudovesicular structures71 or highly percolated pearl-like structures72 for the materials synthesis are also receiving more and more attention. AOT [sodium bis(2-ethyl hexyl) sulfosuccinate] and CTAB (cetyltrimethylammonium bromide) have been two mostly used surfactants for materials June 2, 2005 12:33 00700 244 V. Uskoković & M. Drofenik Fig. 3. Phase diagram of CTAB/1-hexanol/water microemulsion.65 Fig. 4. Molecular structures of some of the most commonly used surfactants as components of microemulsions for materials synthesis. synthesis (Fig. 4), but surfactants such as dodecyl penta(oxyethylene) ether (C12E5), ndodecyloctaoxyethylene glycol monoether (C12E8), cetylbenzyldimethylammonium chloride (CBAC), didodecyl-dimethylammonium bromide (DDAB),14,73 sorbitan monooleate,74 and sodium dodecylbenzenesulfonate (NaDBS)75 have all been used often. Surfactants can be ionic in nature — like anionic AOT or cationic CTAB, or non-ionic (in which case they can be zwitterionic or dipolar) — like Triton X-100 [polyoxyethylene(10)isooctylphenyl ether],76 polyoxyethylene(4) lauryl ether (known as Brij30),77 pentaoxyethylene-glycol-nonyl-phenyl ether (known as Igepal-CO520)78 or the various mixtures of poly(oxyethylene)5 nonylphenol ether (NP-5),79 poly(oxyethylene)9 nonylphenol ether (NP-9)80 and poly(oxyethylene)12 nonylphenol ether (NP-12). Cationic surfactants such as CTAB have been shown not only to be effective in retaining the powder’s properties that are the direct consequence of deagglomeration of particles — such as superparamagnetism is, for instance81 — but also to be effective at condensing and thereby accelerating the transport of genetic material (DNA) across biological membranes.82 Molecular structures of some of the commonly used surfactants for the reverse micellar materials synthesis, are shown in Fig. 4. Whereas AOT/isooctane/water83 has been the major AOT-based microemulsion for the materials synthesis, AOT/cyclohexane/water, AOT/ n-heptane/water42,84 and AOT/toluene/water85 have been used as well. Many different microemulsions based on CTAB have been used: CTAB/ CTAB/2-octanol/water,1,87 hexanol/water,53,86 June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles CTAB/1-butanol/n-octane/water,29,69,88,89 CTAB/ and CTAB/nn-pentanol/hexane/water,90,91 butanol/isooctane/water.92 It was suggested that n-pentanol is, in combination with CTAB, a better co-surfactant compared to n-butanol since it has a stronger van der Waals interaction with CTAB, which ensures the formation of more compact and stable interfacial film protecting nanoparticles from aggregation and nonuniform growth.90 Hexane is denoted as a desirable organic solvent since, due to its high vapor pressure, it may induce the selforganization of monodisperse monocrystals upon dewetting and/or rapid solvent evaporation and even the formation of energetically unfavorable 1D structures.90 Complex processes of particles formation in solution, which involve nucleation, growth, coagulation and flocculation are largely influenced by the presence of surfactant and its ability to compartmentalize reactants as is the case in reverse micellar microemulsions. This role may largely alter not only properties of the final product such as atomic arrangement, morphology, ground transition and product states by affecting the step velocity and rate of crystal growth in certain preferred directions,93 but by changing the pathway of the reaction it may promote the formation of a product of different identity compared to diluted aqueous solution94 or to the reverse micellar medium comprising different surfactants.93 Only surfactant as an additional component in the precipitation reaction is often used in order to increase the specific surface area of the products and stabilize the colloidal solution,95–98 but using ligand or complexing stabilizing agents to inhibit particle growth is a much older procedure than the microemulsion technique.99 Surfactant additives have been shown to act like mobile impurities and thus affecting the step velocity and rate of crystal growth in certain preferred directions.100 Compared to this approach, reverse micelles present more complex structures that medium for the wet synthesis comprises, since in these systems nanosized droplets of water possess a significant individuality and specific organization, which is supposed to reflect on potentially unique and monodispersed properties of the materials synthesized within them. Mixed surfactants have also been used for the reverse micelle preparation of nanoparticles, and are shown to be especially promising for the formation of silica-coated 245 nanoparticles101,102 and 1D nanostructures,103,104 such as nanotubes, nanowires, nanorods, nanobelts and tree-like superstructures, but the effect of any used mixture is generally unpredictable. The components in a surfactant mixture could enhance the efficiency of processes due to synergy or even antagonism among them under different conditions.105 In general, due to the enormous complexity of physicochemical influences, the synthesis of materials within microemulsions today relies more on trial-and-error approaches and the tendency to reproduce experiments than on the formal predictions of morphology, dispersity and various other properties, prior to the performance of the corresponding synthesis procedures. 3. Structural Properties of Reverse Micelles Micellar structures are often presented as being raspberry-like, with hydrophilic charged head-groups closely packed to each other and hydrocarbon chains stretched towards the center of the micelle. This picture is thought to be wrong for two reasons. First, owing to electrostatic repulsion it is not possible to spontaneously pack up to a hundred charged entities close to each other, even if the counterion binding is taken into account. Second, the conformation of all tails being stretched and almost lined up would lead to an enormous local pressure. As an example, an NMR study showed that the uninverted micelles formed by SDS have about one-third of their surface covered by hydrocarbon tails.49 In an attempt to model reverse micelle by computer simulation, nearly all counterions of a reverse micelle comprising 70 AOT molecules, 70 Na+ counterions and 525 water molecules, resided at the interface. The surfactant head-groups did not completely shield the aqueous core from the nonpolar exterior, and some water molecules were trapped between head-groups in contact with the nonpolar phase. This model suggests that it should be possible to catalyze a reaction of an insoluble probe at the micellar interface.35 FTIR studies suggest that the water interior of reverse micelles has a multilayered structure, consisting of the so-called interfacial, intermediate and core water. The interfacial layer is composed of water molecules that are bounded directly to June 2, 2005 12:33 00700 246 V. Uskoković & M. Drofenik the surfactant’s polar head-groups; 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.44 It was reported that five water molecules are tightly bound per one CTAB molecule,52 whereby by using NMR measurements it was realized that only two molecules of water are tightly bound to one molecule of AOT.106 Thus, if parameter w < 6 − 10, the water which occupies the interior of a reverse micelle is highly structurized due to association with the polar head-groups of the surfactant molecules, and the conditions for a typical design of the particles in reverse micelles are said to be not available.73,107 On the other hand, if w > 10, micelles have a free water core with bulk water solvent characteristics. Reverse micelles of various surfactants can solubilize different amounts of water. AOT reverse micelles are known for their ability to enclose very large amounts of water; parameter w ranges from 0 to 70 for many systems.35 Exactly in its ability to solubilize large amounts of water while retaining spherical micellar shapes in a variety of hydrophobic organic solvents lies the reason for the often use of AOT as a surfactant for the application of materials synthesis.108 However, compared to AOT-based systems, CTAB reverse micellar systems, due to higher flexibility of surfactant film that gives rise to a higher exchange dynamics of the micelles, enable significantly higher solubilization capacities of high concentration aqueous salts.28 In contrast to CMC, micellar size varies with various factors in a manner which is complex and at present still difficult to predict.28 A large variety of different techniques, including classic light scattering, viscosity measurements, tracer diffusion, NMR and quasi-elastic light scattering have been performed in order to obtain information about the size distribution of micelles for given parameters such as w or weight percentage of water or surfactant. Though some of these methods gained very precise results according to their performers, a clear criticism was raised regarding the reliability of these methods.28 For a certain system, NMR measurements have clearly indicated that the distribution of micelle size is narrow and somewhat asymmetrical around the average value,28 whereby the SANS measurement results have shown that reverse micelles are spherical and monodisperse at low water content, whereas with increase in water content, the micelles increase in size and polydispersity.109 With an increase in water content of a microemulsion system, the size of the reverse micelles increases as well. In general, aggregation number together with micelle radius increases with w. Other influences on the aggregation number of micellar configurations are dependent on the whole system used. Thus, it was reported that the aggregation number showed little dependence on the concentration of the surfactant or the addition of salt (up to 2.3 M of sodium azide, the salt consisting of azide ion — N=N=N− , which, due to its small size and high charge is likely to dive into the core of the investigated reverse micelles),44 whereby it was observed that size of the reverse micelle is largely influenced by the ionic strength of its aqueous pools. The aggregation number of CTAB micelles increased from 81 to 121 with the addition of NaOH in the concentration of 0.01 M, and micelle shapes were found to change from spherical to elliptical with the addition of NaOH.110 As expected, micelle size increases with increasing alkyl chain length of the amphiphile. In several cases, the growth in size is small and corresponds to a retention of the same micellar shape, but in others, shape changes must be invoked to explain the data. Surfactants with longer tails will according to theory have a lower CMC and a larger aggregation number than analogues with shorter tails; also, counterions that are more strongly bound to the surfactant will induce a lower CMC and a higher aggregation number.49 Whereas the CMC depends little on head-group structure and counterion (ionic surfactant in a solution dissociates on surfactant ion and counterion), micelle size can vary by orders of magnitude. The aggregation number calculated for linear surfactants with C12 , C14 and C16 atoms by using a theoretical model49 were 55, 75 and 95, respectively. These results are in agreement with experimental findings for the corresponding alkylsulphate surfactants, whereby this model predicts aggregation numbers that are too small for the longest alkyltrimethylammonium surfactant. This is due to a combination of the bulkier head-group of this surfactant as well as a change in shape of the aggregate from a spherical to a more prolate reverse micelle. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles However, the main point of the calculation is still present — within a series of surfactants differing in hydrocarbon chains length only, the aggregation number should increase with increasing tail length. Closely related to the fact that a single entity or variable cannot be invoked for any scientific explanation, the size of the reverse micelles is not dependent upon any single variable, but on the complex interactions which are conditional for their existence. Thus, the microemulsion droplet sizes in CTABbased reverse micellar microemulsion were in the range of 30–70 nm111 at 55◦ C, whereby spectroscopic studies have shown that size of the CTAB reverse micelles is in the order of 1–4 nm for different water percentages.63 Reverse micelles based on CTAB ranging in length from 10 to 1000 nm were also found to coexist together.53 Albeit expectations to establishing direct links between thermodynamic stability of reverse micelles in microemulsions and their low polydispersity in size, predictions based on theoretical calculations have shown that micelles are broadly distributed in size with polydispersity index equal to 2.106 Increase in water content results not only in a nonlinear increase in the average diameter of the nanodroplets and the synthesized particles within, but in a more enhanced polydispersity as well.3 For the cases of growth to very long, rod-like micelles, the polydispersity becomes large. For many systems, the low degree of polydispersity is explained by the fact that after the increase in association constant up to a certain aggregation number there is a region of marked anticooperativity with the equilibrium constant decreasing with aggregation number. Until recently it was thought that only reverse micelles spherical in shape exist. However, it has been observed that the probe motion within reverse micelle becomes more anisotropic with increase in water content.52 It was proposed that a slight anisotropic rotational motion of the probe in micelles, as detected, for instance, by electron spin resonance measurements, might be explained by the formation of cylindrical aggregates. Generally, spherical micelles at low surfactant concentration (close to CMC) might, due to aggregate–aggregate interactions, alter their geometries at higher monomer concentrations. For micellar systems showing the sphere-to-rod transition, the anti-cooperativity, which lies in head-group 247 repulsions, is partly eliminated this way. As shown by the case of CTAB, the elimination of headgroup repulsions can be brought about by counterions which may approach the charged groups closely or intercalate between them, or by certain solubilizates which are located in the head-group region.28 Linear surfactants tend to form ellipsoidal micelles more often than branched surfactants, which almost always form spherical micelles. Thus, for instance, a linear surfactant cetyltrimethylammonium 4-vinylbenzoate forms viscoelastic solutions in water containing cylindrical micelles of 4 nm in diameter and thousands of nanometers long,40 whereby AOT-based micelles are known to exist only in spherical shapes. In general, the micelles formed by surfactants with tails of moderate length (approximately C10 − C16 ; note that both CTAB and AOT are C16 surfactants) are thought to be spherical or nearly spherical — at least close to CMC. One of the major challenges within modern surfactant science is drawing a precise link between the geometry of surfactant self-assemblies and the final structures of the synthesized materials.112 Many cases in which acicular particles were unexpectedly produced by reverse micellar synthesis were ascribed to the templating effect of worm-shaped reverse micelles.4,66,113 Reverse micelles based on CTAB as the surfactant are known to produce such effects, so another CMC at which spherical-to-worm-like transition occurs is ascribed to such microemulsion systems. It is also known that an alkali substance,110 salts66 or co-surfactant,53,114 such as 1-hexanol28 or an alkane,66 might induce a spherical-to-worm-like transition. Elliptical and spherical CTAB micelles are known to be able to coexist.53,59 However, adding water to the CTAB/water/n-pentanol/nhexane microemulsion always results in an increment of the reverse micellar radii, whereby the micelle shapes remains spherical.91 While the formation of rod-like micelles is well demonstrated for several cases, especially when CTAB is used as a surfactant, other geometries of large non-spherical micelles have not been observed yet, which is in agreement with theoretical predictions.28 Meanwhile, the computer simulation of a reverse micelle (w = 10) formed from a novel double-chained phosphate surfactant in CO2 , including 1616 water molecules, 160 surfactant molecules, June 2, 2005 12:33 00700 248 V. Uskoković & M. Drofenik 160 counterions and 6991 CO2 molecules, showed that the modeled reverse micelle is not restricted to a spherical shape.35 Molecular interactions drive the formation of the reverse micelles, leading to a range of shapes. Sphere-to-rod transitions in CTAB micelles at higher concentrations have been reported in both the presence and the absence of added salts for surfactant concentrations exceeding 0.10 M.51,56 The concentration at which these transitions occur depends on the property measured. By using absolute SAXS technique it was shown that at c ∼ 0.05 M, CTAB micelles are spherical, and that at room temperature the sphere-to-rod transition occurs right above 0.05 M, at 50◦ C at 0.17 M and at 70◦ C at 0.25 M.51 There is a number of evidence indicating a transition from closely spherical to very long rod-like aggregates for CTAB micelles at a concentration of 0.2–0.3 M.28 Increasing the temperature from 30 to 50◦ C, as well as substituting Cl− for Br− as counterion, eliminates the transition. Addition of small amounts of simple solubilizates, such as benzene or a long-chain alcohol (hexanol, octanol, etc.), may markedly facilitate the transition to rod-like micelles, whereby alkanes have no effect. Decreasing the alkyl chain length considerably increases the transition concentration or eliminates the transition completely.28 By using SANS studies it was shown that micelles of CTAB molecules in 0.2 M solution were ellipsoidal with semi-minor axis of 2.12 nm and semi-major axis of 5.62 nm, whereby aggregation number was 186 and fractional charge 0.09.39 Other SANS studies have concluded that CTAB micelles are ellipsoidal in shape, have an aggregation number of 177, and that in the presence of hydrotropes the aggregation number increases dramatically with decrease in the fractional surface charge of the micelles.115 Certain calculations related with QELS measurements of CTAB micelles in aqueous solutions, and the diffusion of mesoscopic optical probes through the same solution have implied extensive micellar growth and failure of the spherical micelle assumption.43 It has also been reported that CTAB forms rod-shaped micelles in aqueous systems above the second CMC of 0.3 M and assembles into hexagonal liquid crystals above 1.1 M.116 Sizes and shapes of the reverse micelles are dependent on temperature. When the temperature increases, the aggregation number was found to increase as well.44 Although the CMC of ionic surfactants is insensitive to temperature changes, the tendency to form micelles different from the spherical increases with decreasing temperature. The size of the reverse micelles comprising nonionic surfactants increases with increasing temperature,28 such that sometimes at temperatures just above the room temperature, radii of the micelle increases together with broadening of the micelle size distribution and increased excluded volume effects in magnitude, which can be explained by considering the sphere-to-rod or sphere-to-disk transitions of the micelles.55 It is thought that two processes which disrupt microemulsions, that is colescence and Ostwald ripening, accelerate at higher temperatures. Coalescence of reverse micelles is Brownian motiondriven and hence is more present at higher temperatures. On the other side, Ostwald ripening, the disruption of emulsions by the growth of larger droplets at the expense of smaller ones, is driven by Kelvin effect (a high curvature of small droplets creates a high internal Laplace pressure, which consequently increases the vapor pressure of the emulsified monomers). The smaller the droplet is, the greater the tendency for the droplet to shrink and disappear, since the Laplace pressure increases as droplet diameter decreases. It has been shown that L2 range is widening with increase in temperature.117 One common feature of reverse micellar systems is the relative lack of stability even without encapsulated salts. One way of minimizing this phenomenon is including a co-surfactant in the reverse micellar system. It was even noted that for CTAB-based reverse micelles, a co-surfactant is necessarily required for the formation of the stable microemulsion.118 The incorporation of short chain alkanols makes reverse micelles more stable and in addition, the presence of alkanols decreases the aggregation number of the surfactant molecules and the diameter of the reverse micelles.113 N-butanol or SDS (sodium dodecyl sulfate) are often used as co-surfactants, increasing the polarity of the surfactant and helping to stabilize the reverse micelle solutions.73,119 N-butanol as a co-surfactant is used together with CTAB to help decrease the fraction of the micellar head-group that is neutralized and thereby increase the stability of the micelles. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles Without the addition of a co-surfactant, the amount of free water available to carry on the reactions is greatly reduced, as most of the water is locked in the head-groups of surfactant molecules. When using 1-hexanol as a co-surfactant in the system CTAB/isooctane/water, the encapsulated enzyme’s half life was increased 45-fold.118 As an oil-phase, a non-branched alcohol has been considered as an optimal choice, but the use of diesel oil (together with CTAB as a surfactant, 1-butanol, 1-propanol, 1-octanol or 1-pentanol as a co-surfactant) — a complex mix of aliphatic alkanes with chain lengths of up to C32 — was also noticed.30 In relation to the application of microemulsions for materials synthesis, it is generally recommended that surfactant and aqueous fluid be each from about 1–30% by weight of the total system and the maximum total amount of surfactant and water be up to 50% (preferably to 30%). The amount of co-surfactant is preferably between 25 and 75% by weight of composition.73 Penetrating into the micellar interface, cosurfactant molecules change the mean distance between the polar head-groups of surfactant molecules and thus reduce the electrostatic repulsion between the head-groups, promoting the sphericalto-worm-like micelle transition.59,114 This penetration may increase the volume of the micelle core, which is equivalent to increasing effective head-group (hydrophilic) cross-sectional area. It is well-known that the curvature of a micellar aggregate (whose aqueous side is concave rather than convex, being an essential property of reversed micelles), which is dependent on the interfacial tension between micelle and oil phase,34 is strongly influenced by the ratio of the effective head-group cross-sectional area to the effective cross-sectional area of the aliphatic chain. By decreasing this ratio, the surfactant aggregate shapes should follow the trend: spheroidal micelle → worm-like (rod-shaped) micelles → bilayer structures → reverse structures.59 When the alcoholic co-surfactant chain length decreases for a given micellar radius, the thickness of the penetrated layer is increased and the interaction between micelles is therefore stronger, because attractive intermicellar potential depends on the volume of the interlapping region.62 The volume of this region increases with an increase of micellar radius as well. 249 4. Dynamic Structure of Reverse Micelles In order to understand processes in which the synthesis of materials within reverse micelles takes place, it is useful to outline the basic dynamic interaction among reverse micelles. In Fig. 5, a schematic representation of certain processes ocurring within reverse micellar microemulsion as a medium for the synthesis reaction, is shown. On one side, the reverse micellar systems are extremely dynamic in nature and such dynamic character lies at the basis of their thermodynamical stability. On the other hand, the appearances of reverse micelles which emerge from numerous experimental results are one of a labile and sensitive multimolecular configurations.107 Surfactants associate into liquid-like aggregates that possess rapid molecular motions often on different time scales. The complexity of the system is due to the highly dispersed and heterogeneous nature of the overall phase. A microemulsion system during the synthesis procedure comprises a variety of species such as surfactant molecules, solutes in the aqueous phase, species which may be preferentially distributed in one of the phases, and the solid particles formed in reaction followed by nucleation. Some of the basic interactions in such a system include the exchange of surfactant molecules between reverse micelles and the bulk phase, the formation–breakup of dimers, trimers, and lower-level aggregates which may be present in relatively small concentrations, intermicellar fusion–fission of reverse micelles, and effects involving solute species such as intramicellar kinetic reactions, particle nucleation, and growth via intramicellar attachment and intermicellar exchange of aqueous phase contents.24 For reactions in microemulsions involving reactant species completely confined within the dispersed water droplets, a necessary step prior to the chemical reaction is the exchange of reactants by the coalescence of two droplets.120 Microemulsions have a dynamic structure wherein the droplets of the dispersed phase diffuse through the continuous phase and collide with each other. These collisions are inelastic because droplets coalesce and temporarily merge with each other, but subsequently break to form separate droplets again, so that the average size and number of the micelles remain the same as a function of time.120 Surfactant film flexibility is one June 2, 2005 12:33 00700 250 V. Uskoković & M. Drofenik Fig. 5. Schematic representation of entities and dynamic physicochemical effects existing in a reverse micellar system.24 of the key parameters in determining the intermicellar exchange and particle growth processes, since the interdroplet exchange of the particles growing inside the reverse micelles is inhibited by the inversion of the film curvature in the fused dimer which, in turn, depends on the film flexibility.99 The intermicellar exchange of content leads to a distribution of material throughout the system, and a mathematical description of this phenomenon requires the use of population balances. Partial opening of the surfactant layer or interfacial transfer of aqueous pool solubilizates present close to the micellar surface can result in a partial exchange of material. Under equilibrium conditions, the size and the aggregation number of the reverse micelles are dictated by thermodynamics and equilibrium transport between micelles and bulk phase. As is evident from experimental reports, even the molecules of different coexisting species follow distributions of their occupancy numbers, defined as the number of particle species solubilized in the micelle core. For noninteracting and random kinetics, the distribution is known to be Poissonian, whereas geometric and binomial distributions can result under other conditions. This feature clearly leads to a situation different from continuous conditions. The typical time scales that have been reported24 for different mechanistic effects in polar reverse micellar systems are: 1–10 ns for diffusion-controlled intermicellar reaction, 50–100 ns for the lifetime of a pair of reactants confined to the micelle core, 300–500 ns for the formation of an encounter complex via diffusion on the micellar surface, 1–100 ns for a diffusion-controlled entry of a surfactant molecule June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles onto the micellar surface from the bulk phase, 0.1–1 µs for diffusion-controlled collision of micellar aggregates, and 0.1–1 ms for intermicellar exchange to occur. While micellar re-orientation and the molecular diffusion around the micelle occur on the nanosecond scale, the local motion of spherical micelles is very fast (1–40 ps), compared to reorientation of rod-shaped aggregates, which is seen as the third contribution to the relaxation times.121 The retention time of water molecules on the surface of reverse micelles lies within a nanosecond in order of magnitude.107 Although reverse micelles have been studied in detail and are well characterized, especially on AOT systems, the mechanistic details of reactions of finite rates, ultrafine particle formation, and growth seem to have not been studied through a single consistent mathematical model.24 There are many unanswered questions and challenges concerning role and behavior of surfactant within microemulsion during the materials synthesis. Even if a universal relationship or a set of relationships are developed between the structure of surfactants and their behavior at interfaces, the question regarding their dynamics when adsorbed layers are perturbed, as in the case of nanoparticles formation, will remain open due to specific interaction with reactants and products of the synthesis procedure.105 It was shown that two populations of reverse micelles exist depending on whether they encapsulate probe molecules or not,13 which is an apparent sign of the feedback effects of product properties on the structure of reverse micelles, neglected in models that are built on implicit adoption of chemically inert, nano-templating role of reverse micelles, and normally based on the introduction of parameter w as the major influence on product’s properties. Reference 99, which is often cited, concerns the independence of the particle size of the synthesized Pt, Rh, Pd and Ir, always in the range of 2–5 nm,1 on the surfactant, water amount and reactant concentration used in the experiments. Therefore, particle sizes can be seen as controllable by solvent stabilization of the particles, whereas the surfactant acts as a stabilizing agent.10 There is a large attention paid to developing a convenient theoretical framework for correlating complex combinations of compositional, structural and interactional parameters with morphological properties of synthesized particles. Due to their 251 nonlinear nature, colloidal spheres are in general difficult to model in their dynamic interactions. Contrary to expectations based on DLVO theory122 (according to which, only few static charges on colloidal particles’ surfaces can cause repulsions strong enough to keep them stably separated), it has been observed that like-charged colloidal particles sometimes attract each other.123 The attraction between aggregates is found to increase when the micellar size increases, the alcohol chain length decreases, the polar head-area increases and the molecular volume of oil increases.38,124 Attractions seem to be favored by highly charged spheres in very low salt concentrations.123 Basically, the balance of forces acting on three energy scales — Van der Waals attraction, randomizing influence of thermal energy, and hierarchy of electrostatic interactions among highly-charged colloidal particles and the single-charged simple ions around them determines all static and dynamic properties of microemulsions. It is generally thought that since being confined to closed micellar structures, water molecules have restricted motions and their diffusion slow and approximately equal to the diffusion of the droplets. If all the surfactant molecules are located at the water — oil interface, they will obviously diffuse with equal rates as the droplets too.121 However, from comparing experimental results on the diffusion behavior of various components in CTAB and watermediated reverse micellar microemulsion, it can be realized that water has the highest self-diffusion coefficient (D = 2×10−9 m2 · s−1 ) — higher than bromide ion (around 7 × 10−10 m2 · s−1 ), surfactant (around 8 × 10−11 m2 · s−1 ) and the least diffusive solubilizates with D ranging from 7 × 10−11 m2 · s−1 to 7 × 10−12 m2 · s−1 depending on the concentration. Water, bromide ions and surfactant molecules had diffusion coefficients almost independent on the concentration above CMC value.125 NMR measurements126,127 have shown that mobility of the ionic surfactant’s molecular chain progressively increases from its head-group to the terminal methyl group. However, this is not always the case since mobility of 2-ethyl side chain in AOT molecule is very restricted. Anyhow, the general opinion is that mobility of different parts of long surfactant molecules is independent on the surfactant’s nature and presence of a co-surfactant. As water June 2, 2005 12:33 00700 252 V. Uskoković & M. Drofenik content in a four-component water-in-oil AOT-based microemulsion is increased, mobility of surfactant’s polar head-group progressively decreases. This effect was ascribed to an increased order in interfacial layer as well as to the decreased freedom for inner movement in the case of a larger number of molecules, including the polar head when water content is decreased. The largest increase in mobility was observed at low water content when water molecules were engaged in hydratation of sodium ions and sulphonate head, and when the water core with properties of bulk water had not yet been formed. Beside exchanging molecules of reactants by direct coalescence of micelles, it is necessary to take into account the exchange of molecules between phases other than only interiors of aqueous nanodroplets. This exchange might in the general case be considered so as to occur between bounded, interfacial states and free, bulk states. NMR investigations have revealed that the exchange time of Na+ ion between the bounded and the free state in AOT-based water-in-oil microemulsion is less than 10−4 s.128 However, the same time was reported to lie in nanosecond range.129 On the other hand, investigations based on electric field jump130 concluded that the retention time of I− ion in Stern’s layer is about 10−7 s. NMR investigations have concluded that the typical time for the exchange of alcohol molecules between the interfacial layer, the continuum phase and/or the dispersed phase (depending on its affinity) is a lot less than 10−4 s.107 The addition of NaCl Fig. 6. in dispersed aqueous phase leads to the decrease in an exchange rate of alcohol molecules,131 probably due to an increase in compactness of surfactant monolayers induced by an increase in ionic strength of aqueous phase. The exchange rate of neutral probe molecules between the interfacial monolayer and the aqueous core of reverse micelle was estimated to ∼ 107 /sec in AOT/heptane/water microemulsion system at w = 31.132 The relaxation time of the exchange of surfactant molecules between the interfacial monolayer and aqueous phase in oil-in-water microemulsions was experimentally determined to be 3 × 10−8 s,133 whereby proof for the existence of the surfactant’s exchange between two separate phases within water-in-oil microemulsions do not yet exist. NMR researches have shown that movement of surfactant is largely limited to the interfacial space between the oil and water phases.107 It is also known that the fluidity of the surfactant’s monolayer might be increased by decreasing the length of alcohol co-surfactant’s hydrocarbon chain. Typical times for the rearrangement of surfactant monolayers in birefringent microemulsions are estimated to > 0.1 µs.134 Two major processes during which exchange of reactants between reverse micelles takes place are shown in Fig. 6. Solutes can be exchanged between reverse micelles either by processes of coalescence of reverse micelles — during which temporary merging of aqueous droplets occur (fusion) — that subsequently fragment (fission), or by partial fragmentation of droplets with consequent loss of Processes of exchange of reactants (others than via continuous phase) in reverse micellar microemulsion.107 June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles aqueous solubilizates, that might later merge with other droplets (coagulation). There are many indications that the merging of reverse micelles as well as fast intermicellar exchange of their aqueous contents are diffusion-controlled and are independent on the surfactant’s nature.135 The time needed for single fusion of reverse micelles to occur is estimated at ∼ 10−6 s, whereby the time constant for the collision of reverse micelles with subsequent merging is between 108 and 109 /Msec in respect to the droplet concentration. A model according to which intermicellar material exchange occurs through a water channel formed at the point of temporary merging of reverse micelles (temporary dimer) has been considered.107 It has been discovered that compartmentalization of reactants in reverse micelles decreases reaction rates due to the coupling of chemical reaction and the rate of micellar mergings. However, there are many cases of catalytic effects derived from the compartmentalization of reagents, including the increase of the rate constant of the hydrolysis of acetylsalicylic acid in the presence of imidazole catalyst by 55 times when the reaction was performed in AOT/supercritical ethane microemulsion compared to the aqueous buffer,136 and the reaction of acid hydrolysis of phthalomohydroxamic acid in AOT/isooctane/water microemulsions.27 Kinetic analyses of the bimolecular rate constant of exchange ke of reaction have concluded that rate constant ke is complex since the reaction of exchange comprises Brownian diffusion of reverse micelles leading to collisions, collision of two droplets, water-channel opening (merging), diffusion and chemical reaction between reactants, as well as fragmentation of a transient dimer. The slowest step of these basic rate processes will scale the temporal aspects of the overall process of synthesis,137 and the second stage of this complex reaction is decisive for limiting reaction rate. The rate constant ke takes an approximate value of 107 /sec at room temperature.138 Therefore, approximately one in a thousand collisions result in a temporary merging of dispersed aqueous nanodroplets with exchange of reactants. However, such 253 effectiveness of reactants’ exchange is said to be valid only for rigid surfactant films, such as AOT-formed reverse micelles.10 For flexible films, such as the ones that CTAB forms, one in ten collisions can give rise to micellar content exchange.10 Large activation entropy of this process is ascribed to the fact that after the temporary merging of reverse micelles (that is, the formation of activated state) some of surfactant molecules from the interfacial layer are released to one of the separate phases. The lifetime of the transient dimer cannot be longer than a few µs since a longer lifetime would lead to the separation of phases due to the uncontrolled droplets ripening. Uneven fragmentation of the previously merged droplets leads to an increased polydispersity of reverse micelles. However, modification of the system ought to result in a decrease in polydispersity. In the case of microemulsion systems based on CTAB and pentanol, large variations in reverse micelles sizes were noticed, but an addition of alkane led to a decrease in polydispersity, which was most significant for the alkanes with long hydrocarbon chains.107 Since larger ke implies that microemulsion acts more like a critical system, it is thought that there must exist a correlation between the increase in ke , the polydispersity of reverse micelle sizes, their mutual interaction and the critical behavior of microemulsion. It has been observed that a low reaction rate may lead to very large particles irrespective of the exchange protocol of micellar content.139 The limiting situations concerning chemical reactions occurring between reactants enclosed in reverse micelles — merging-controlled and reactioncontrolled cases are shown in Fig. 7. Co-surfactant leads to improved chemical reaction rates within reverse micellar microemulsions. It has been noted that the addition of benzylalcohol as a co-surfactant increased ke by almost 20 times.140 The rate constant ke is dependent on the length of alcoholic co-surfactant’s hydrocarbon chain — it increased when 1-hexanol as a co-surfactant was replaced by 1-pentanol.140 The value of ke is largely dependent on the oil hydrocarbon chain’s length — decrease in ke by a factor of 10 was noticed when moving from dodecane to hexane.107 The micellar exchange rates of cyclohexane, heptane, and decane in AOT-based microemulsions are approximately 106 , 107 and 108 M−1 · s−1 , respectively.137 In some cases, such as for AOT/n-heptane/water June 2, 2005 12:33 00700 254 V. Uskoković & M. Drofenik of the enhancement of reaction rates is not definitely known, but it is widely accepted that an electrostatic effect in the aqueous phase of the reverse micelle is one of the reasons for the acceleration of reactions. The properties of local reaction media are quite different from those of the bulk solutions as a consequence of the intense local electric fields, affecting all the relevant parameters that modulate the reaction rates.10 Specific intermolecular interactions at the hydrophilic sides of surfactants surrounding the aqueous cores and specific water structure in this region are proposed to have catalytic effects on the rates of chemical changes.141 However, reaction kinetics are neglected in many models because intermicellar material exchange is relatively slow, and it plays a major role in particles growth.24 5. General Synthesis Procedures Fig. 7. Limiting situations for the kinetics of chemical reactions taking place within reverse micelles.107 system at w = 33, ke was only slightly smaller (∼ 1.5 × 109/Msec) than the rate of droplet collisions (∼ 109/Msec), which suggests that in that case reactants exchange between reverse micelles is no longer limited by the rate of interfacial layer opening, but primarily by the diffusion of droplets.107 Even though kinetic equations describing reactions of synthesis include factors such as merging of micelles which carry reactants, that must happen prior to the reaction, increases in rates of chemical reactions taking place within reverse micellar systems are known phenomena, and have recently been almost taken as the general advantage of this method.24 It has been reported that the rate of oxidation of Fe2+ and subsequent formation of needlelike 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.52 In the case of certain iron complexes, a two to tenfold increase in the rate of dissociation was measured in comparison to the pure aqueous solution.121 The cause One way to perform reverse micellar synthesis of materials is to produce one parent microemulsion and then to successively let reactants to diffuse into the interior of reverse micelles and react. The major problem with this so-called single-microemulsion approach is that all reactants do not react at approximately same conditions defined by their physical surroundings, but significant concentration gradients are involved. The second problem is that the composition of the microemulsion is gradually changed as solutions of different reactants are successively introduced, which might induce significant transitions in micellar or some other multimolecular structural properties. Therefore, a multi-microemulsion approach, within which separate microemulsions of the same compositions are prepared for every reactant involved in the synthesis, is used most often in order to overcome the problems of the single-microemulsion approach and achieve a better control over synthesis parameters. Schematic illustrations of single- and multimicroemulsion approaches to the materials synthesis, are presented in Fig. 8. It has been observed that the latter, so-called multi-microemulsion approach yields finer particles when compared to the single microemulsion approach.80,142 The particle size distribution was also found to be different when synthesis is performed by single- and double-microemulsion approaches. The particle size distribution had a June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles 255 Fig. 8. Schematic illustration of various stages in the growth of nanosized particles in reverse micelles in multimicroemulsional approach (up) and single-microemulsional approach (down).121 Gaussian symmetrical bell shape for the singlemicroemulsion approach, whereby bimodal distribution was attributed to the powder synthesized by the double-microemulsion approach.142 Whereas the procedures based on separately dissolving the reagents within dispersed reverse micelles are used most frequently in the context of materials synthesis, the procedures according to which a type of ion that is to become incorporated in the final product, is initially introduced as the constituent of the surfactant molecules, are also often followed. CdS nanoparticles were synthesized by using Cdbased surfactant, such as cadmium bis(ethyl-2-hexyl) sulfosuccinate.143,144 Ultrafine ZnS and CdS particles were, as well as CaCO3 particles,145 prepared in reverse micelles by using extractant-metal ion surfactant complex as a metal ion source.146 In the case of the preparation of metallic particles, the reduction reactions are initiated, whereby in the case of the preparation of ceramic particles, usually a precipitation agent in the form of an acidic or alkali agent is used and further thermal treatment is performed; otherwise, an oxidation reaction is initiated as well.147,148 Disruption of micelles in the reaction mixture might be carried out by introducing alcohols in excess, causing the nanoparticles to precipitate. The major problem of the recovery procedure is caused by high surface energy of ultrafine particles, which makes them coagulate irreversibly when reverse micelles are destroyed without any protection treatment,149 such as surface modification of the particles. Removing the surfactant caps surrounding the individual, well-dispersed particles most often yields highly agglomerated samples even though the particles were well dispersed in their parent microemulsion. Centrifugation is mostly used for the separation of precipitate from the liquid phase, but magnetic separation by a permanent magnet has been used as well,150,151 just like filtration, freezedrying or phase separation by cooling the reverse microemulsion.152 To separate CTAB from sensitive polypeptide nanoparticles, gel permeation chromatography (GPC) was used.111 Washing of the precipitate in order to remove residual surfactant and oil-phase molecules might be done with water, ethanol, chloroform or the mixture of two. Since most surfactants including CTAB and AOT are soluble in these solvents, their molecules are together with byproducts to be separated from the particles of the desired product. Even though the calcination is sometimes a necessary step to obtain fine crystalline samples, in situ syntheses of nanocrystalline as-dried powders have been frequently reported. Nanocrystalline MnZn-ferrites117,148,153–156 [Fig. 9(a)], NiZn-ferrites147,157,158 [Fig. 9(b)], TiO2 159 were synthesized in situ at room temperature as well as a number of various metal particles such as Fe 29 or Bi.160 Metallic particles can be obtained much June 2, 2005 12:33 00700 256 V. Uskoković & M. Drofenik (a) of surfactant molecules acting as cages for growing crystallites which thereby reduce the average size of the particles during the collision and aggregation process76 is seen as the major reason for such a difference in particle sizes and crystallinity. Since many chemicals are included in the ordinary microemulsion synthesis (where surfactant, oil phase and co-surfactant are normally environmentally degradable162 ), one of the most important challenges of future researches is to develop a way to recycle163 the components which are included in the synthesizing procedure, so that literally the same reverse micelles might be used in the repeating cycles of materials production. Many ‘green’ options are nowadays available, with lecithins, bile salts, dipotassium glycerrhizinate, biocompatible solvents like Transcutol, glycofurol, ethanol and isopropanol, and surfactants like poloxamers and polysorbates being routinely used in microemulsion-assisted syntheses of drug nanoparticles.164 Since the synthesis reactions proceed in stable reverse micellar microemulsions only in the dilute solution of precursors, the result is a low yield of nanomaterial, which places the problem on the economic feasibility of the method. 6. General Overview of the Influences (b) Fig. 9. (a) In-situ-obtained nanocrystalline MnZnferrite117 and (b) NiZn-ferrite161 in CTAB/1-hexanol/ water reverse micellar microemulsion, with the latter having saturation magnetization of 50 emu/g, which is approximately two-thirds of a value that the same stoichiometric samples obtained by the traditional solid-state route exhibit. easier in situ by using reverse micelles compared to ceramic materials, since they are usually prepared in such a way that they are the direct products of the precipitation reaction between the dissociated cationic species and a reduction agent. The samples synthesized in the reverse micelles are often of better crystallinity when compared to the samples synthesized in the bulk aqueous phase. The existence There are many factors that might be discerned as having an effect on the properties of the produced powders within the synthesis procedure used. Within this paragraph, we shall number and discuss some of them. Besides parameter w (that will be discussed in detail in the following chapter), mentioned earlier as a major variable in the design of materials that are to be prepared in reverse micelles, particle sizes might in certain cases be controlled by varying the amount of metal ion in each reverse micelle — more metal ions will cause larger particles to grow because of diffusion.73 It was generally proposed that an increase of particle size might be induced by increasing the reactant concentration, whereas a decrease of particle size can be induced when one of the reactants is increased in excess until a plateau is reached at high excesses.99 However, one of the major problems with microemulsion-assisted material processing is the effect of the reactants and products on the stability of the microemulsion, particularly June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles the metals concentration in the aqueous phase. It is mentioned that the ion concentration is, after water-to-surfactant ratio, the second most important parameter to control the particle’s size.21 As has been long known in the literature, the introduction of a small amount of a species highly insoluble in the continuous phase into the emulsion droplets substantially increases the stability of the droplets with respect to the Ostwald ripening process.165 As droplets shrink and lose dispersed monomer, the concentration of the insoluble species within the droplet grows. The desire of the droplets to maintain osmotic equilibrium by maintaining the same osmotic pressure (same concentration of insoluble species within each of the droplets) competes against Ostwald ripening. Droplets cannot shrink and disappear because of the insoluble species, and so the number concentration of droplets, along with the number concentration of insoluble species within the droplets, remains constant (in the absence of coalescence processes). Electrolyte addition in ionic systems in general gives an increased micelle size; in most cases the effect is rather small, while in others dramatic changes may occur. Adding [Ru(bpy)3 ]2+ to small CTAB reverse micelles was found to change the reverse micelle water pool size distribution from monodisperse to bimodal.44 The influence of electrolyte addition on the reverse micellar solution phase L2 obviously depends on the interaction of individual components, and to date defies any general conclusions. Some researches have demonstrated narrowing the L2 range in DDAB/dodecane/water microemulsion with an increase in salinity,166 which is consistent with observed narrowing of the L2 range in CTAB/1-hexanol/ water microemulsion.117 There were also results according to which extensions of one-phase regions in C9 H19 C6 H4 O(CH2 CH2 O)9.7 H/cyclohexane/ water microemulsion were not affected by the addition of NaCl.167 Salts have been found to affect both aggregation and micelle formation. When the ionic strength of the aqueous phase is increased, the water uptake will decrease dramatically, and the type of ions has a great effect on the water uptake.53 Aggregation number, CMC and critical micelle temperature all depend on ionic strength. Studies of AOT reverse micelles show that introducing electrolytes such as NaCl reduces aggregation number 257 and micelle radius. Others report that at low concentrations (10−3 –10−2 M) ions do not alter reverse micelles appreciably.44 It was reported that adding salt to an ionic micellar solution will decrease the CMC and increase the aggregation number owing to the screened electrostatic repulsion.49 With a decreased electrostatic repulsion between the charged head-groups of surfactant, it is possible to pack the surfactant head-groups closer to each other, with a subsequent increase in aggregation number. Besides expectations that the higher concentration of salt precursors within reverse micelles tends to stabilize the micelle structure, because of the suppression of surfactant head-groups, which causes the head-group area to decrease and the packing ratio to increase, inducing the formation of more stabilized microemulsions, some authors have found out that excessively high salt concentrations tend to drive the alcohol (co-surfactant) into the oil-phase and some studies even suggested that this would limit the salt concentration at which microemulsions can form at around 0.4 M in the aqueous phase.168 Nevertheless, microemulsions, albeit with extremely soluble sodium nitrite and alkanes with a variety of chain lengths comprising the oil phase, were formed at concentrations in the order of ∼ 3.5 M.168 The solubilization capacity is dependent on the nature of the surfactant as well, since, as has been already mentioned, it was shown that CTAB-based reverse micelles have higher solubilization capacities of high concentration aqueous salt compared to AOT-based systems.53 Beside w and pH, the particle size and morphology is greatly influenced by the ratio of surfactant to co-surfactant. In the CTAB/n-butanol(cosurfactant)/isooctane/water microemulsion system, the higher the ratio of CTAB to n-butanol was, the smaller were the particles obtained. Therefore, it was proposed that the key factor affecting the particle size is the interfacial property rather that the size of the microemulsion droplets.169 An increase in particle size could be obtained by directing an increase of the surfactant film flexibility, which might be achieved not only by approaching the microemulsion instability phase boundaries or by changing the droplet size, but by increasing the amount of cosurfactant (alcohols) or changing the chain length of the oil or co-surfactant as well.99 An increase in the chain length of oil results in an increase June 2, 2005 12:33 00700 258 V. Uskoković & M. Drofenik in the value of intermicellar exchange rate coefficient due to the fact that as the chain length of the oil increases it becomes increasingly coiled and therefore its penetration in the surfactant layer becomes more difficult, resulting in a stronger mutual interaction between surfactant tails compared to the intensity of interaction between surfactant tails and oil molecules.137 The co-surfactant in a quaternary microemulsion system may have a more significant effect on the interfacial properties than the oil phase in a ternary one.169 It was shown that in CTAB/n-hexanol/water microemulsion, n-hexanol acts mainly as the continuous oil phase, although it also affects the interfacial properties. It was observed that the average size of the nanoparticles increased with the increase in the n-butanol/CTAB (cosurfactant/surfactant) weight ratios, except when the weight ratio of n-butanol/CTAB was below 0.5 (gelation occurred). It was also revealed that the variation of the weight ratio of n-butanol/CTAB affected only the interfacial properties and not the size of reverse micelles.169 The presence of alcohol in the CTAB/hexane/pentanol/water microemulsion is shown to be an important factor in regulating the size distribution of the synthesized nanoparticles, acting on the particles growth by influencing the flexibility of the interfacial film.170 Different particle morphologies can in some cases be achieved simply by choosing NaOH over NH4 OH.76,116 However, in the presence of a large amount of strong base, the total ionic strength of the water pool increases, which causes instability in the microemulsion system.76 The ability of a strong base to promote the hydrolytic decomposition of an ionic surfactant will contribute to the destabilization of the surfactant aggregates77 and possible subsequent phase separation of the microemulsion system.3 In order to keep the pH value of the precipitation at the same level during the process, a computercontrolled constant-pH apparatus was used in some experiments.171 However, it is important to note that water pool properties of reverse micelles — local viscosity, local polarity, local acidity — can be substantially different compared to the effective macroscopic properties of an overall measured system.10 Investigations of local pH in reverse micelles by using pHsensitive probes concluded that when NaOH and HCl were used for pH adjustments, an almost constant intensity ratio over a wide pH range was obtained, suggesting that the water pools of microemulsions, due to a large number of polar surfactant headgroups localized at the oil–water interface, may have buffer-like action.172 The temperature of the synthesis procedure influences the properties of the synthesized powders. It was shown that higher temperatures induce better crystallinities of the synthesized powders.154,156 The reason for this effect might be found not only in an increased thermal motion of the active species, but by the fact that the micellar structures change as well. Average radii of AOT-formed micelles increase by approximately 50% when the temperature is increased from 20◦ C to 38◦ C.46 Similar findings were applied to CTAB-117 and PEGDE-based173 systems. On the other hand, SANS studies have concluded that both AOT shell thickness and the inner water pool in the AOT/n-heptane/water system reversibly decreased with increasing temperature.42 Contrary to expectations, an increased temperature did not increase the reduction rate of Pt4+ with H2 in the CTAB/octanol/water system.1 Albeit the existence of many cases where in a matter of moments after the key reaction was initiated, particles stable for months were obtained,174 the average size of the particles can sometimes be tuned by setting the desired time between the mixing of the microemulsions containing the reacting species, and their disruption.21 By prolonging this time from one minute to two hours, the average diameter of the particles, as observed by TEM, changed from 13 to 35 nanometers. On reducing the reaction time, the morphologies of the same particles evolved from perfect cubic shapes to more spherical ones with rounded edges.21 On the other hand, prolonging the reaction time can sometimes lead to a change from cubic to spherical-shaped particles.175 Within an in situ synthesis of MnZn-ferrite particles, it was observed that the average particle size did not significantly change in the aging time range between 1 min and 3 h, whereas ∼ 25% increase in size was detected after aging of the powder in microemulsion up to 26 h.156 Aging time variations can be applied in order to induce self-organizing effects on particles assembly scale. During an aging time of 60 days, the particles of PANI/TiO2 composite were shown to self-organize in reverse micelles of AOT/iso-octane/water and CTAB/hexanol/water microemulsions into either June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles sea urchin-like or needle-like shapes.176 Four-day aging time of prepared Prussian blue nanoparticles in microemulsion led to the formation of highly ordered cubic superlattices.177 By letting the reaction microemulsion mixtures age for more than 10 min in the process of preparation of CaCO3 and BaCO3 , aggregation of the already-formed spherical nanoparticles is initiated, leading to the slow formation of networks (after aging of up to 1 h) and solid wires (after aging of up to 48 h).178 Introducing an external magnetic field to the process of evaporating deposition of cobalt particles prepared in reverse micelles, led to the availability of designing the arrangements of the nanocrystals into either hexagonal patterns of micrometer-sized dots or labyrinthine 3D superlattices, depending on the intensity of the applied magnetic field and the evaporation rate of the previously prepared particles.144 A similar procedure, but without the use of magnetic field, was successfully applied in obtaining “supracrystals” made of silver nanoparticles self-organized in 3D superlattices.179 Some of the often skipped influences might present decisive parameters in the synthesis procedures. Bunker et al. have found that tripling the amount of microemulsions used in the synthesis procedure leads to significant changes in the obtained CdS particles absorptivity.180 It was found that the procedure of passing N2 through the reaction mixture not only protects critical oxidation but also reduces the particle size by 25% when compared with methods without removing the oxygen.181 Purging of inert gases through the reaction mixture is especially important when easily oxidized elements are present, as is the case for the synthesis of ferrites147,157 and other iron-comprising compounds.182 The matter of purity is significant, since any introduction of foreign substances might induce heterogenous nucleation with avoiding the desired monodispersity and uniform morphology of the synthesized particles. Sequence according to which adding of microemulsion components into the mixture is done is also shown to be important.75 It was found that even stirring microemulsion with a magnetically coupled stir bar during the powder’s aging time can influence crystal quality and in some cases result in a different crystal structure as compared with nonmagnetically agitated microemulsions.183 In case of the synthesis of organic nanoparticles in reverse 259 micelles,174 the use of a magnetic stirrer led to the formation of nanoparticles larger in size compared to the particles obtained with using ultrasound bath as a mixer, even though no changes in particle size were detected on varying solvent type, parameter w, reactant concentrations and even geometry and volume of the vessel. Particle size can be controlled by decreasing the exchange of materials between different micelles, which might be accomplished by lowering the temperature or slowing the agitation of the solution, which will in return slow the intermicellar exchange of materials, giving either smaller particles or equally-sized particles over a longer time scale.73 All of the processing conditions which affect the formation of a nanocompound, i.e. its microstructure and stoichiometry and therefore its potential application, may be summarized as follows. Besides the identities of surfactant (the size of hydrophilic headgroup, ionic or neutral, branched or narrow, etc.), oil phase (oil chain length, primarily), co-surfactant if included (its structure, branching, chain length, etc.), metal salts, precipitating agent and/or other reactants, there are: curvature free energy, determined by the elastic constant and the curvatures of the surfactant film;99 concentration of reactants within reverse micelles; pH; the sequence of introducing the components in the micellar mixture; single- or multiple-microemulsion synthesis approach; overall composition of microemulsion, which is usually decoupled to molar ratio of water-to-surfactant described by parameter w and/or surfactant-to-cosurfactant ratio; aging time of the precipitate in microemulsion; intensity and mechanism of stirring; a method for separation of precipitated particles from the liquid phase; purity of used substances and incorporation of impurities during the synthesis; and temperature under which various processes within micelles proceed. However, it is necessary to mention that tendencies to outline universal relationships which govern the processes of materials synthesis are disobeyed by many cases in which changing only one ion in the complex microemulsion had dramatic qualitative effects on the product’s properties. When only Mn2+ ions were replaced by Ni2+ ions,particles of mixed zinc-ferrites had completely different morphologies. Thus, the synthesized MnZnferrite117 comprised spherical nanosized particles, whereas particles of NiZn-ferrite were of acicular June 2, 2005 12:33 00700 260 V. Uskoković & M. Drofenik shapes.157,94 When bromide ions of CTAB surfactant were replaced by chloride ions, the product’s identity was not the same.87 When 2-octanol was replaced by octanol, much higher background noise was observed within XRD measurements of the synthesized powders.87 The inability to discern and predict the influences of all the species involved in the synthesis procedure and provide a set of relationships between properties of the microemulsion used and the potentially designed properties of the synthesized materials, is the reason why the method of synthesis discussed herein had relied mostly on the trial-and-error approach in the past. 7. Influence of Water Content Over the past two decades, a great deal of research within the reverse micellar method for the materials synthesis was undertaken to discern the various parameters which significantly influence the properties of the synthesized materials. By precise evaluation of the effects of all the major influences within the method, through control over processing parameters the desired material properties might be designed. The parameter, which is in this light regarded as the most important, is water-tosurfactant molar ratio, described by parameter w. The higher the parameter w, the larger the water pools of the reverse micelles, and if the micelles are imagined as nanosized templating cages, then the particle sizes might be seen to increase with the increase of w value. However, as we shall see, there are many cases of synthesis for which reversed correspondence or no linear correspondence at all has been concluded. It was found that reverse micelle size generally increases linearly with parameter w as well as with water content of the microemulsion. The relation between water core radius r of reverse micelle in nanometers and parameter w is in the case of AOTformed reverse micelles found to be very simple: r = 1.5w.22,50 Even though this relation between the size of the water pool radius of the reverse micelles and w was confirmed by SANS and SAXS measurements,50 it is sometimes mistakenly identified121 with all reverse micelles, whereas it was concluded only for certain types of AOT-based microemulsions. Beside increasing the volume of the reversed micelles with the water content when surfactant concentration is kept constant and decreasing the volume of the reversed micelles with the surfactant content when the water concentration is kept constant, it is known that if both the surfactant and water are increased in a constant ratio, the volume of the reversed micelle will not change, but its numbers will increase.44 In the case of reverse micelles formed by CTAB molecules, it was reported that the reverse micelle size is equal to parameter w.23 There were also indications that parameter w might be taken to influence the particle size distribution.4 It was reported that at constant surfactant concentration S [mol/dm3 ], the radius r [Å] of the droplets encapsulated by surfactant monolayers increased linearly with the volume fraction of the dispersed phase φ and could therefore be readily controlled through this parameter184 : r = (4.98 × 103 )φ/As S where As [Å2 ] is the area occupied by the surfactant at the droplet surface. However, this relation would provide a templating relation for control of particle size only in cases where the microemulsion structure is not perturbed during the reaction. In cases where transfer of the components through intermicellar diffusion and exchange processes between reverse micelles permits fast exchange of both solutes and surfactant between droplets, the result will be an uneven growth from which the structure of the droplet template may be altered during the reaction or be totally destroyed as a result of phase separation. Today, it is known that it is not only synthesized particle sizes that do not show in all cases linear dependencies on parameter w, but the very size of the reverse micelles also do not follow the same dependency. Experimental results indicate that the size of reverse micelle depends not only on parameter w, but also on temperature, organic solvent (oil phase), type of surfactant, its concentration and present electrolytes. The same SAXS researches which have come to the conclusion that dependency of reverse micelle radii on w is parametric,50 have shown that reverse micelle radii in AOT/isooctane/water system vary upon addition of small amounts of compounds solubilized in the microemulsion.22 It was noted that the water core radius of microdroplets of water/AOT/toluene microemulsion changes from June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles 1 to 1.6 nm when parameter w changes from 5 to 10.185 As the concentration of the surfactant decreases, the amount of water that can be solubilized decreases too.44 Some studies have shown that the smallest Stokes radii, which correspond to the droplet size, were obtained between 15◦ C and 18◦ C in case of AOT-formed microemulsions, and that there was no relation between the Stokes radii of the microemulsion and AOT content of the system.2 It was shown that the structure and composition of an elementary water droplet enclosed by surfactant molecules remained unchanged up to a water volume fraction per composition equal to 0.3, but this observation generally refers only to microemulsions where attractive forces are not very strong.62 It was very often found that the synthesized particles are of greater average particle size than the size of their parent reverse micelles as derived from parameter w.185 Within a synthesis of silica particles, as the parameter w increased from 0.7 to 2.3, the particle size decreased (and the size distribution became narrower) not in monotonous manner.6 Within the reverse micelle synthesis of BF2 nanoparticles, the particle size was larger than the swollen micelle size,87 which demonstrates that the microemulsion does not consist of hard spheres that strictly limit particle growth. In case of microemulsion synthesis of AlPO4 -5 fibers, the dimensions of the crystals produced were many times larger than the typical dimension of microemulsion droplets.183 Directly proportional correspondence of reverse micelle sizes and average particle size turned to inversely proportional at higher water-to-surfactant ratios.186 Within the synthesis of Pd particles in CTAB/ n-butanol/isooctane/water microemulsion, the average diameters of the produced particle also decreased slightly with the increase in water content when the weight ratio of surfactant-to-isooctane was fixed.169 The following explanation was assigned to this phenomenon: increase in w pushes the composition of the microemulsion toward the solubilization curve, and therefore to increased intermicellar interaction. Increases in intermicellar interaction will promote the aggregation of the particles contained in the interacting reverse micelles. Therefore, it can be clearly concluded that in addition to the size of microemulsion droplets (described by parameter w), there must be other factors affecting the size of the produced nanoparticles. 261 Today, it is generally known that for the most surfactant-mediated synthesis applications, the connection between the morphology of the surfactant aggregates and the resulting particle structure is much more complex (than the simple connection between the sizes and shapes of the micelles to the sizes and shapes of the produced particles), being affected by hardly reducible conditions prevailing in the local microenvironment surrounding the growing particle. These include the pH, composition, and ionic strength, as well as surfactant headgroups spacing. As reaction takes place within these microenvironments, many of these factors are liable to change. The decoupling of each of these effects still remains a challenge, and provides interesting new opportunities in this field.112 8. Mechanism of Particle Formation Synthesis of materials in reverse micelles might be considered as an ordinary precipitation or (co-)precipitation method of synthesis, but which takes place within a very specific environment. Compared to many other wet syntheses such as sol-gel, within reverse micelle synthesis, the precursor ions incorporated in form of a colloidal precipitate and not a complex compound, must be obtained. One of the major problems related to the co-precipitation method of synthesis is the inability to control supersaturation of the precipitating system which in turn leads to very poor control over the distribution of particle sizes, that is normally much more broadened than in the cases of reverse micellar syntheses.29 Due to narrowly dispersed reverse micelles with templatelike properties, the synthesis of materials within reverse micelles very often overcome these problems. The particle formation in a microemulsion can occur either through collisions between microdroplets which carry reactants within their aqueous cores (multiple microemulsion approach) or by diffusion of the hydrophilic precipitating reactant to the microemulsion droplets containing the to-be-precipitated reactant (single microemulsion approach). Although the surfactant walls of the microdroplets act as cages for the growing particles and thereby reduce the average size of the particles, the exchange of the precursor ions between the microdroplets by either a diffusion process or June 2, 2005 12:33 00700 262 V. Uskoković & M. Drofenik a collision of microdroplets affects the particle formation, resulting in a similar narrow particle size distribution.185 Droplet collisions and rapid intermicellar exchange of their water content are processes which are mainly controlled by diffusion and are also dependent on the nature of the surfactant.135 Wide size distribution of the synthesized particles was ascribed to the fact that for reverse micelles with smaller size, nucleation will occur only in a small number of micelles at the beginning of the precipitation reaction because most of them do not contain enough ions to form a critical nucleus.53 Since the typical ion concentrations per reverse micelle are too small to form critical nuclei, the formation of the final precipitated particles requires the aggregation of precursor atoms from several different reverse micelles. Therefore, due to diffusion, new nuclei will form as a function of time, which will cause a different growth rate of particles, and subsequently a broad size distribution. Size distribution can be narrowed by increasing the concentration of reagents, as was proven in case of the synthesis of Ni nanoparticles when increased hydrazine concentration gave rise to enhanced reduction rate that led to the generation of a higher number of nuclei and, therefore, to the formation of smaller and more uniform particles.187 When using a co-surfactant to increase the fluidity of the interface and therefore the kinetics of the intermicellar exchange, a more homogeneous repartition of reactants among micelles could be ensured, thus promoting the formation of smaller, but more numerous particles.188 By using a Monte Carlo model, Jain et al. showed that as the efficiency of the intermicellar exchange increased, the resultant particle size increased,139 whereas within the synthesis of Ag nanoparticles in AOT-based reverse micelles, it was shown that the particle size decreases at higher intermicellar exchange rates.137 By studying the solvent effects on copper particles prepared in AOT reverse micelles, the authors concluded that growth rate and particle size are inversely related, so that a decrease in growth rate leads to a larger particle size at a specific w value.189 From this point of view, the dynamic intermicellar interaction might be considered as the major, but thoroughly complex influence on the properties of the synthesized particles. One of the often skipped questions with regard to materials synthesis in reverse micelles is the matter of influence of the nucleation and crystallization of particles, as well as of their subsequent interaction on the structure of reverse micelles. The tendency for the particles to agglomerate sometimes exceeds the forces which keep reverse micelles dispersed within the oil phase and complete or localized phase separation occurs. Researches based on steadystate and time-resolved emission spectroscopy as a function of water content came to the conclusion that encapsulation of a relatively small particle or molecule (around 1.2 nm) within a large reverse micelle (with diameter of 3 nm and w > 10) does not perturb the nanodroplet structure, but on the other hand, an incorporation of a similarly sized particle within a small reverse micelle (with diameter of 1.5 nm and w ∼ 5) results in a redistribution of surfactant and water.58 Three possible hypotheses were proposed in order to explain the final states of the particles synthesized within reverse micelles.190 According to the first hypothesis, nanoparticles are in the organic phase and are in direct contact with the polar heads of the surfactant, whereas the surfactant tails are in organic phase. According to the second hypothesis, nanoparticles are surrounded by a layer of water, whereas according to the third hypothesis, nanoparticles are surrounded by surfactant tails that point their polar heads toward the water. If the microemulsion becomes unstable, mostly by exceeding the limiting parameter w for given oil/surfactant ratio, the aggregation of particles takes place, which leads to the formation of nonuniform particles regarding their size and shape as well as to the formation of multicore particles in the process of composite synthesis.185 It was mentioned that in the case of AOT/isooctane/water microemulsion system, the extreme size limits of 5 nm and 30 nm exist, where the size distribution increases due to the problems in micelle stability and crystallization.151 Also, depending on the nature of the surfactant and the type of interaction, the amphiphilic molecule may either stabilize the inorganic particles or induce significant aggregation. It has been demonstrated, for instance, that positively charged particles flocculate upon addition of anionic surfactants and that, on further addition, the particles re-disperse because of the formation of surfactant bilayers.135 When hydrothermal and reverse micellar methods for producing zincophosphate crystals were June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles compared, it was noticed that the crystal growth process was about an order of magnitude slower within the reverse micellar procedure, as compared to the hydrothermal method.191 The surface of the hydrothermally obtained crystals was found to have multiple layers and terraces, whereas the crystals obtained from the reverse micellar method were atomically smooth, which suggests a much finer mechanism for the formation of particles within the reactions coupled with obviously morphologyregulating intermicellar exchange and surfactantmediated growth control. Ultrathin nanowires (Fig. 10) have recently been produced by using the so-called catanionic reverse micelles that are formed by mixed cationic–anionic surfactants, whereby the molar ratio between the mixed surfactants is expected to play a key role in the synthesis of such 1D nanostructures.103 Only mixtures with an excess of either cationic or anionic surfactants, when ratio of two oppositely charged head-groups is far from 1:1, have been shown to form catanionic micelles.192 The length of V2 O5 nanowires was shown to be possible to be tuned by controlling the aging time of the synthesized particles in microemulsion,193 which is consistent with the findings in the case of the formation of ZnS nanowires.194 Parameter w also largely influenced the process of the formation of ZnS nanowires in C12 E9 /cyclohexane/water reverse micelles. It was shown that acicular NiZn-ferrite particles are formed within CTAB/1-hexanol/water microemulsion only within the precipitating pH value range of 8–11.94 Colloidal particles with an electric or magnetic dipole moment were theoretically predicted to selfassemble into flexible chains, and these assemblies were recently experimentally observed.195 The production of such pearl-like microstructures today surely presents a significant challenge for materials design by using self-assembly templating amphiphile structures. Surfactant aggregates associate (or dissociate) with time scales that can range over several orders of magnitude. If the time scale associated with the particle formation is small compared to the time scale of such transformations, intermediate, nonequilibrium aggregate structures might afford additional opportunities for templated materials synthesis. Fast kinetics associated with several solution-based precipitation processes might be ideally suited for this 263 (a) (b) Fig. 10. BaCrO4 nanowires103 as obtained from catanionic reverse micelles consisting of (a) water, decane, undecylic acid and decyl amine and (b) molecular sieve fibers of AlPO4 -5183 as obtained from cetylpyridinium chloride/butanol/toluene/water reverse micellar microemulsion. application. The control of solution conditions and surfactant architecture to exploit nonequilibrium surfactant morphologies presents an interesting scientific challenge.112 The potential of reverse micellar microemulsions in the field of nanotectonics — controlled construction of organized matter using nanoparticle building blocks — is significant and is expected to bring about new challenges in materials research. June 2, 2005 12:33 00700 264 V. Uskoković & M. Drofenik 9. Synthesis of Composites within Reverse Micelles Reverse micellar microemulsion medium is a dynamic system where colliding micelles coalesce, exchanging their content and splitting. This property provides conditions for performing the desired chemical reactions of synthesis in a range of new and unpredictable manners. On the other hand, this property causes instability of colloidal solutions over time. Freshly formed nanocrystalline particles protected from further growth in micelles, have high surface energy and once micelles coalesce, particles readily aggregate.196 Nanoparticles in ferrofluids, for instance, have surface energies higher than 100 dyn/cm2 , so strong magnetic dipole–dipole attractions between particles tend to provoke their agglomeration.181 In order to prevent agglomeration of nanoparticles and retain their nature of nanocrystals, an inert silica coating (Fig. 11) or a novel mesoporous silica matrix is used, offering excellent control of nanocluster size distribution and cluster morphology as well as high thermal stability associated with the inorganic host structure.171,197 The insulating matrix greatly enhances the chemical stability of the particles, and the wear and corrosion resistance of the media.198 When present in small quantities (< 0.05%), the silica component can impede grain growth through impurity drag and precipitate drag mechanisms, whereas when silica is present in larger quantities (> 0.05%), a liquid phase may form.199 If the liquid phase is not properly distributed, discontinous grain growth is probable. However, if the liquid phase is properly distributed, grain-growth impedement can still occur if the diffusion path between grains has been increased sufficiently to offset the increased diffusion rates caused by the reactive liquid. The coating with silica improves the structural order at the surface of the magnetic nanoparticles in diminishing the surface magnetic anisotropy.77 By varying the thickness of the silica shell and the coated particle radius, an overlap of the wavefunction and band gap can be designed, which opens attractive new possibilities in the semiconductor field.10 The outer silica surface is biocompatible and functionalizable as well. DNA was successfully attached to silica-coated magnetic nanoparticles.76 To additionally protect the nanoparticles from agglomeration, (a) (b) Fig. 11. Silica-coated rhodium particles as synthesized from a reverse micellar C-15/cyclohexane/water microemulsion.101 silica surface can be covalently modified into a neutral or charged surface76 or additionally coated with some biomolecules. Hydrophilic organic coatings such as albumin, dextran or hydroxemethylmethacrylate may enhance biocompatibility.165 June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles Alkoxides tetraethyl orthosilicate [Si(OC2 H5 )4 — TEOS, known as tetraethoxysilane as well] or calcium isopropoxide [Ca(O3 H7 )2 ] are mostly used as precursors for SiO2 or CaO coatings.200–204 Immersing TEOS in water and ethanol results in a bubbly structure of solid silicon in a matrix of an organic compound (HOCH2 CH3 ), which is removed in a subsequent step.205 Because this process occurs slowly at room temperature, acid or base catalyst is added to the formulation. The amount and type of catalyst play key roles in the microstructural properties of the final coating. After addition of TEOS into a microemulsion mixture, the silica shell can be synthesized by gradually increasing the pH inside the droplets with ammonia to catalyze the condensation of TEOS.77 Optimal pH value is found to be between 11 and 12.116 When using ammonia as a catalyst, more spherical morphologies are obtained and therefore, ammonia is often referred to as a morphogenic catalyst.116 For higher ammonia concentrations, a minimum in particle size was observed at intermediate w values, and it was suggested that this trend might be associated with the aggregation of the nuclei.206 Ammonia concentration was found to have impact on particle sizes, where at relatively low w, the particle size goes to a minimum as ammonia concentration is increased, but at relatively high w, the particle size increases monotonically with ammonia concentration.6 However, in the presence of a large amount of strong base, the total ionic strength of the water pool increases, which causes instability of the microemulsion system. To avoid this problem, it was suggested76 that a very small amount of base be used, that is sufficient for the precipitation reaction followed by the polymerization reaction of TEOS. Ammonia also provides the particles with a negative, stabilizing surface charge, which does not enable aggregation of the particles.207 It was found that at low ammonia concentrations, the particle stability is higher. On one hand, an increase in the concentration of ammonia and water will increase the surface charge through the dissociation of silanol groups and will thus stabilize the particles. On the other hand, increasing the concentration of ammonia and/or water will increase the concen− and thereby decrease the tration of NH+ 4 and OH 207 If the particles are very double layer thickness. stable (e.g. at low concentrations of ammonia) the 265 aggregation stops relatively early and a larger number of silica-coated particles grow to a small final size. Because in this case the silica layer that is deposited after the aggregation is thin, the resulting particles will not yet be very spherical and the surface will be rough. At high ammonia concentration the aggregation continues longer and only a relatively small number of particles becomes stable. Silica-coated ceramic particles may be synthesized by the double-microemulsion approach, where the first microemulsion contains aqueous solution of the salt precursors and the second microemulsion, that is to be introduced dropwise in the first one, contains an alkali precipitant and TEOS. In this way, mixing and sonication for 2 h together with a day of aging and subsequent washing with ethanol and centrifugation yielded silica-coated magnetic particles.76 By the use of an external magnetic field during the solidification process, magnetic nanoparticles embedded in silica matrix could be preferably oriented in order to obtain an anisotropic macrostructure, known as a magnetically textured system,208 which has such properties as anisotropic magnetization and optical anisotropy. It was found that the diameter of silica-coated particles was highly dependent on the water-tosurfactant molar ratio. Not only did the amount of free water influence the relative rates of hydrolysis and condensation reactions, but it also played a determinant role in colloidal stabilization. Nucleation involves the condensation of monomers via intramicellar or intermicellar exchange polymerization reactions, whereas particle growth occurred either by the addition of TEOS to the nuclei or by particle aggregation.135 The major problem for the reverse micellar preparation of silica-coatings is that TEOS is an organophilic molecule and is therefore more readily dissolved in the oil-phase alcohols than in the aqueous droplets.77 TEOS is also known to be incompatible with strong oxidizing agents. Upon TEOS addition to the reverse micellar system, the formation of silica coating over alreadyformed nanoparticles involves a series of steps,206 which can be identified as: (i) association of TEOS molecules within the reverse micelles; (ii) TEOS hydrolysis and formation of monomers; (iii) nucleation; June 2, 2005 12:33 00700 266 (iv) (v) (vi) (vii) (viii) V. Uskoković & M. Drofenik particle growth; nuclei dissolution; intermicellar exchange of monomers; ionization of monomers; and particle surface ionization. The reaction temperature for hydrolysis and condensation of TEOS monomers should be kept below 25◦ C to avoid any complications due to the presence of a possible thermally induced phase inversion,77 and after 48 h the reaction is normally completed. It was suggested that the synthesis is silica-driven and that the micelle-initiated nucleation is not the dominant mechanism, but the nucleating site for the porous silica coating is either oligomeric or polymeric silica.116 In most cases, not all ethoxy groups from TEOS are hydrolyzed and therefore part of the solvent remain trapped in the silica pores. In fact, it was proven that several percent of the ethoxy groups never leave the TEOS molecules.207 Besides the water-to-surfactant molar ratio and concentrations of ammonia and water, the ionic strength can be used to influence the final particle size. The hydrolysis and subsequent condensation of TEOS can be affected by the precursor salts introduced to the initial solutions of the gels. In the case of nitrate salts, for instance, due to high solubility of nitrates, the hydrating water from salts allows a total hydrolysis of TEOS molecules and as a consequence of that, the oxygen presence is low. On the contrary, the chloride salts retain the water hydration retarding the hydrolysis process and promoting a rich oxygen ambient, sometimes convenient for the formation of a different phase.209 The number of ions determines the particle size at which colloidal stability is reached and when the aggregation stops; a smaller number of growing particles results in a larger final silica coating. In a certain experiment, the hydrolysis rate was unaltered by the addition of LiNO3 salt, while the final radius of synthesized particles was significantly increased.207 When comparing the number of silica particles with the number of surfactant aggregates, one silica particle was produced out of circa 104 to 106 surfactant aggregates, depending on the parameter w. The minimum number of hydrolyzed TEOS monomers required to form a stable nucleus was estimated to be 2, and the nucleation efficiency factor, i.e. the probability of effecting nucleation in a reverse micelle that contains enough monomers to produce a stable nucleus was found to increase with w for relatively low ammonia concentrations (1.6 wt%). A decrease in the apparent nucleation efficiency factor was observed at high w values with more concentrated ammonia, and this was attributed to silica nuclei aggregation promoted by enhanced intermicellar collisions and intermicellar exchange.206 The thickness of the SiO2 layer increases with increasing hydrolysis time, and therefore it can be designed by controlling the hydrolysis time. In order to terminate TEOS hydrolysis, the micellar structure of the solution containing SiO2 -coated nanoparticles can be destroyed by adding an alcohol such as propanol.101 Metallic nanoparticles synthesized in reverse micelles have been coated with gold layer by the reduction of Au3+ from HAuCl4 to Au via excess hydrazine.150 The gold shell on the metallic particles in this case provides the functionality required to form organized arrays on functionalized substrates. Such so-called nano-onion structures comprise a solid core grown in the first step, which serves as a nucleation source for a functionalized shell.119 The Au:Fe:Au nanocomposites produced by a sequential reverse micelle technique have been shown to exhibit the GMR effect.210 Gold has become a favored coating material because of the simple synthetic procedure and its chemical functionality.211 Ferrofluids as prosperious magnetic composites212 were prepared by using reverse micellar microemulsion.213 Magnetic nanoparticles in ferrofluid composites are coated with surfactant monolayers and are immersed within a carrier that might be an organic solvent such as heptane, xylene, toluene, or an inorganic solvent such as water, a hydrocarbon (synthetic or petroleum) or any ester, polyglycol, or styrene. Magnetic nanoparticles in such systems are unable to aggregate and thus the superparamagnetic nature of the whole system is retained.212 Oleic acid derivatives have proven to be of great benefit in the encapsulation reactions of colloidal magnetic particles through microemulsion polymerization. The surface of magnetite colloid was first coated by a monolayer of sodium monooleate and then stabilized by adsorption of SDBS surfactant.135 Reverse micellar synthesis is especially convenient for this sort of composites due to capability of tuning various magnetic properties — such as magnetization or June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles coercivity — by control of simple parameters within the given procedure of synthesis. An interesting approach has been found in the synthesis of ceramic — polymer composites, within which nanosized spherical crystals of the ceramic phase are incorporated into micron-sized polymer particles. Some of the essential properties of magneto–ceramic–polymer composites are its light weight, good formability toughness and flexibility, typical for plastics, that allow the preparation of the products for magnetic circuits with complicated shapes.214 The remarkable aspect of polymerization in reverse micelles is the fact that the polymer could form precipitates out from the solution with the morphology of interconnected submicron-sized spheres.215 It was claimed that micelles have a templating effect, to fold chains as they are formed to the resulting spheres. It was proposed that the polymer chains grow around the micelle interface and the collisions of reverse micelles with growing chains result in chain linkages and the formation of interconnected spherical particles. As a direct consequence of the assembly process of surfactants on mineral surfaces, a hydrophobic interlayer can form on the inorganic particles, into which monomers can solubilize, and polymerization can subsequently proceed. Whereas covalent binding through in situ polymerization techniques allows the formation of nanocomposite particles with hairy-like, polynuclear, or core-shell morphologies, electrostatic coupling or physicochemical adsorptions preferentially afford heterocoagulated nanostructures with raspberrylike morphologies.135 P-ethylphenol (PEP) is used as an organic precursor and polymerization of these monomers was achieved by adding hydrogen peroxide with an oxidative enzyme (horseradish peroxidase) in order to polymerize the phenols encapsulated in the reverse micelles.215 Polymerization within reverse micelles95 yields oligomers of controlled conjugation lengths and without occurrence of undesirable nano-oligomers coagulation or interconnecting, and, therefore, the products of such procedures might find a number of applications in electronics and photonics such as for controlled band gaps.216 In this sense, PPV [poly (p-phenylenevinylene)] is an interesting option because it is traditionally made by a base-catalyzed reaction of a water-soluble salt monomer precursor and has a number of applications due to its 267 electrical, nonlinear optical, electroluminiscent and lasing properties. The dispersion polymerization or microemulsion polymerization in the presence of surfactant-coated particles carries the risk of incomplete and nonuniform encapsulation. Reaction conditions must drive all the to-be-coated particles to transfer uniformly into the resulting polymer particle, or these particles must provide the only site for the precipitation of polymer. One way to potentially guarantee the uniform loading of core particles into latex was proposed: directly disperse the magnetic particles into the monomer of interest, emulsify, and then polymerize.165 Theoretically, the insolubility of the polymer and magnetic particles in the oil continuum should guarantee uniform loading of magnetite per latex; however, no proof has been provided yet. Upon polymerizing acrylamide monomers microemulsified in oil, 40-nm-diameter polyacrylamide latex results. Upon microemulsifying and precipitating iron salts within water droplets in microemulsions, magnetic nanoparticles result. A combination of the two approaches results in magnetic nanoparticles encapsulated in hydrophilic polymers, with average diameters ranging from 80 to 320 nm.185 However, a relatively polydisperse latex in this case contains only 3.3 wt% (0.6 vol%) of magnetic particles. A material with a greater concentrations of the core phase might arise using the larger droplets in reverse micellar microemulsions, created by using relatively low amounts of surfactant, as opposed to the nanodroplets of microemulsions, created with significantly larger amounts of surfactant.165 It was demonstrated that the presence of the coupling agent (MPS) was a prerequisite condition for the encapsulation reaction within the synthesis of silica/polystyrene colloids. The dispersion of core particles, which is a major drawback of encapsulation reactions in microemulsion polymerization, can be significantly improved with the aid of ultrasound. In miniemulsion polymerization, an effective surfactant/hydrophobe system was used, and the role of the hydrophobe was to prevent or retard coalescence of the droplets and Ostwald ripening.135 The particle size was shown to be possible to be controlled by water-soluble cross-linker concentration and surfactant/water ratio.217 When using micelle polymerization as a term, surfactant molecules are regarded as monomers, June 2, 2005 12:33 00700 268 V. Uskoković & M. Drofenik whereas when using microemulsion polymerization as a term, monomers are thought to be encapsulated within reverse micelles and polymerize under surfactant restrictions. Surfactant monomers (so-called surfmers) are used often in order to gain nanoparticle–polymer composites with narrow size and shape distribution. Since the pioneering surfmer synthesis performed by Freedman et al. who reported the first synthesis of vinyl monomers which also functioned as an emulsifying agent,218 vast amount of literature has been dedicated to the topic of polymerization of/in organized amphiphilic assemblies.219 Polymerization of surfmers was observed only at concentrations above CMC, proving that micelle formation is a necessary precondition for polymerization.219 Surfactants acting as surfmers are CVDAC (cetyl-p-vinylbenzyldimethylammonium chloride, that might be photopolymerized too), DDM-MEAC [N,N-didodecyl-N-methyl-N-((2methacryloyloxy)ethyl)ammonium chloride] and didecyldimethylammonium methacrylate.220 The first synthesis of 2–5-nm-sized latex particles was performed by using didecyldimethylammonium methacrylate, a polymerizable surfactant which forms reverse micelles in toluene.221 10. Survey of the Types of Materials Synthesized by Using Reverse Micelles Materials of many different chemical compositions were synthesized within reverse micelles and we shall list some of them. Applications of the produced materials by using reverse micelles are widespread in paints and surface coatings, catalysis, separation media, drug delivery systems, highfrequency electronic components, etc. The possibility of obtaining materials with high surface areas makes reverse micellar methods of synthesis viable for the production of catalytically active transition metals, including Pt,222 Au,90 Ag,186 Pd,169 Rh,1,223 Ir,1 PtPd,224 AuAg,225 AuPd,226 PtRu,227 PtCo,228 Zr,142 Si,6 Ni,187 Ti,229 Bi,230 Cu,4,231 Se,84 Ce,232 Ce1-x -ZrxO2 ,233 Ce–Tb mixed oxides,234 (Ce,Zr)O-x/Al2 O3 ,235 CeZrO4 ,236 CeZrO4 /Al2 O3 ,237 Pd/CexZr1-O-x(2)/Al2O3 threeway catalysts,238 LaMnO3 /CdS photocatalyst,239 and ferric sulfide catalyst for direct coal liquefaction (DCL).240 The metal nanoparticles produced within reverse micelles are characterized by a higher catalytic stability and effectiveness than classical metal catalysts.108 Nanosized and narrowly sizedistributed reverse micelle aqueous cores open a way to synthesize narrowly distributed semiconductors with marked quantum-size (Q-size) effects, such as: CdS,24,149 PbS,241 ZnS,242 CdE (E = S, Se, Te) semiconductors.243 BaF2 and BaF2 :Ce,87 BaF2 :Nd,244 AgBr,245 AgI,42,113 AgCl,246 Ag2 S,4 Cu2 S,247 MoSx ,188 CeF3 ,248 Bi2 Te3 ,249 ZnSe,250 B alloys,251 Ni-, Co- and Fe- borides,252 tungstic acid,253 Cu2 [Fe(CN)6 ],254 amorphous AlOOH255 were also synthesized within reverse micelles. Many composites have been prepared, which primarily include using silica coatings or matrices as additional component. Such materials are silica-coated Co nanoparticles,256 Pt/silica,257 Pd/SiO2 particles,78 Cu/SiO2 ,258 Ag/SiO2 ,259 SiO2 -coated Rh nanoparticles,101 SiO2 -coated CeO2 nanoparticles,260 SiO2 -coated Co-ferrite and magnetite,102 SiO2 -coated Pt, Pd and PtAg particles,261 SiO2 -coated NiFe permalloy nanoparticles,312 silica-coated Fe2 O3 ,76 silica-coated ZnFe2 O4 ,77 silica-coated CoFe2 O4 and MnFe2 O4 ,262 TiO2 /SiO2 composites,263 ZrO2 /SiO2 .264 The use of gold as an additional component of composites produced in reverse micelles — exemplified by the cases of preparation of gold-coated Fe nanoparticles,211 gold-coated FePtx and CoPtx particles,265 goldcoated Co, CoPt and CoPt3 magnetic alloys266 or superparamagnetic magnetite-gold composites267 — is noticed as well as the use of polymers in polymer ceramics nanocomposites,86,215,217 including magnetite polymer composites,185,268 nanocomposites of polyaniline with AgCl, BaSO4 and TiO2 .86,176 Composites of TiO2 /Fe2 O3 ,269 CdS particles coated with ZnS,270 Co–Ag and Fe–Cu–Fe nanoparticles with antiferromagnetic coupling between the Fe layers271 were also prepared within reverse micelles. Beside using reverse micelles and other dynamic nanostructures of self-assembling surfactant molecules in microemulsions to limit the particle size of dispersed systems formed by precipitation in the case of inorganic particles, various approaches have been used for lattices gained by polymerization reactions. Nanoscale polymerization reactions within the reverse micellar confinement yield products with controlled conjugation lengths and, thus, controlled band gaps for applications to electronics June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles and photonics. In addition, control of the length of the polymer blocks provides the prospects of monodispersity, improved processability, and designable preparation of nanocomposites. Polymeric nanosized particles (latexes) were first synthesized by using the reverse micelle method221 and since then many polymers, including poly(p-phenylenevinylene) (PPV),216 were synthesized within reverse micellar microemulsions. Many procedures of synthesis concerning encapsulation of nanoparticles with polymers,326 either by extracting the particles from the microemulsion or by in situ polymerization, such as in the case of the preparation of Bi/PMMA,160 CdS/polyacrilamide nanocomposites272 and BaSO4 in polymerized surfactants,273 were also reported in the literature. High-Tc superconductors YBa2 Cu3 O7-x 274–276 were prepared by this method as well as many materials with exotic morphologies, such as V2 O5 nanowires,193 ZnS semiconductor nanowires,194 PbCl2 nanowires,277 Sb2 O3 and Sb2 O5 nanorods,278 BaCO3 single-crystal nanowires,279 single-crystal SnO2 nanorods,280 CaCO3 , BaCO3 and CaSO4 nanowires,178 MnOOH nanorods,281 BaSO4 nanoparticles and nanofilaments,282 star-shaped CdS patterns,283 nanosized fluorescent hybrid silica (NFHS),284 fluorescent europium(III) chelatedoped silica nanoparticles,285 and silica-coated terbium(III) chelate fluorescent nanoparticles.286 Ca(OH)2 ,3 CaCO3 145 and materials for biomedical applications, such as calcium hydroxyapatite287,288 and Ca(PO4 )3 ,289 have also been produced within reverse micelles. The pioneering nanoengineering via biocompatible microemulsion was performed by Debuigne et al. and was related to the preparation of anti-inflammatory drug nimesulide by direct precipitation in reverse micellar microemulsions comprising either Epikuron 170 (lecithin)/isopropyl myristate/water/n-butanol or E170/isopropyl myristate/water/n-isopropanol190 and to the synthesis of cholesterol, Rhovanol and Rhodiarome in AOT/heptane/water, Triton/decanol/water, and CTAB/hexanol/water microemulsions.174 Many ceramic materials have been synthesized by using this method: ZrO2 –Y2 O3 nanoparticles,53 barium hexaaluminate nanoparticles,152,290 zirconolite CaZrTi2 O7 ,291 CuHo2 O5 and Cu2 Er2 O5 ,88 CaSO4 ,292 ZnAl2 O4 ,89 YVO4 ,293 perovskite-type 269 LaNiO3 , La2 CuO4 and BaPbO3 ,5 Y2 O3 : Eu3+ ,294 Gd2 O3 : Eu3+ and Gd2 O2 S : Eu3+ ,295 Na2 PO4 and K2 PO4 ,296 calcium thiophosphate,297 zincophosphates191,298 as well as many oxides, such as GeO2 ,299 ZnO,300 EuO,301 SnO,302 TiO2 ,303 ZrO2 ,85 CeO2 ,304 Fe2 O3 ,269 Al2 O3 ,305 iron oxide-doped alumina nanoparticles,306 Csdoped alumina nanoparticles,307 and bariumstabilized alumina nanoparticles.308 Reverse micelles have shown to be feasible in producing magnetic ceramics, such as manganites LaMnO3 89,309,69 and LaSr-manganites,310,311 Li(Mn,Co)2 O4 ,74 ferrites NiFe2 O4 ,79 MnZnFe2 O4 ,80,117,313,314 147,157,161 NiZnFe2 O4 , MnFe2 O4 ,75 SrFe2 O4 ,31,92 67,120 CoFe2 O4 , CoCrFe2 O4 ,315,316 BaFe2 O4 ,317 69 318 LaFeO3 , and Fe3 O4 . Many magnetic materials, as Fe,29,319 Co,23 FeCo nanocrystals,182 Co, CoPt and CoPt3 ,266 KMnF3 and NaMnF3 antiferromagnetic particles,83 were successfully synthesized within reverse micelles. 11. Future Directions After all that has been said, it useful to take a look once again at the major conclusion which emerges from this review. On one hand, reverse micelles are useful multimolecular structures that give significant opportunities for the design of nanostructured materials of well-defined and uniform properties. On the other hand, reverse micelles cannot be considered as nano-engineering reactor shells because their size and permeability cannot yet be tuned with sufficient precision.108 “Nanoreactor” does not sound as a convenient term at all for describing the function of reverse micelles in microemulsion-assisted syntheses of materials, since conventional microreactors imply enabling precise control over the reactions and properties of the resulting nanomaterials, and it is still difficult to achieve this aim at the nano level by using the reverse micellar technique. Combinations of the reverse micellar precipitation method or any other microemulsion method with some other methods of synthesis — such as solgel,291,320,321 hydrothermal synthesis,183 ultrasonic irradiation,322 UV irradiation,323 pH-shock wave method,324 and flame-spraying325 — that have been used lately, have presented many new potentialities and advantages. Merging of two or more preparation techniques into one can, due to synergy effects, lead June 2, 2005 12:33 00700 270 V. Uskoković & M. Drofenik to multiple advantages, such as improved control of the stoichiometry of the final product (taken from sol-gel method) and extremely fine and controllable grain size (taken from reverse micellar synthesis) in the examples290,327 of the combination of sol-gel and reverse micellar approaches to materials synthesis. Low yields, surfactant-contamination and difficulties arising from 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,325 at the same time transcending poor control of particle size and shape, which is a typical drawback of 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 (Fig. 12) of cobalt nanocrystals.144 Carbon nanotubes have as well, for instance, been prepared by direct introduction of in situ prepared catalytically active Co and Mo particles by a reverse micellar method.328,329 Silver nanorods encapsulated by polystyrene were prepared by a combination of reverse micellar, gas antisolvent, and ultrasound techniques.330 In order to overcome difficulties arising out of the fact that it is not possible to carry out sequential reactions inside the same reverse micelles in order to obtain multilayered composites, layerby-layer (LbL) techniques comprising adsorption of oppositely charged polyelectrolytes on a solid surface of synthesized particles in reverse micelles have been used.108 The assemblying of particles formed in the processes of reverse micellar and, in general, microemulsion-assisted syntheses into precisely tailored, supra-nanocrystalline 3D structures, clearly presents an important challenge. In situ reactions in well-organized and immobilized amphiphilic matrices present only one approach toward this goal. The use of shear on liquid crystalline systems for instance, can be used to produce the alignment of microscopic domains into macroscopic regions in the direction of shear, or other interesting morphological transitions. External stimuli provides great promise, and it is predicted that in the future days, progress in Fig. 12. Self-organized arrangements of cobalt nanoparticles obtained by combining the synthesis of the material within reverse micelles and evaporating deposition thereof.144 using light, electric and magnetic fields to organize particles into 3D matrices will be realized. A more active and potentially richer approach is the use of engineered peptides to direct the self-organization of nanoparticles into technologically important structures. The real test of self-assembly and materials synthesis will be the generation of novel functional structures with unique and useful properties.112 Future research in the field of the materials synthesis within reverse micelles may not only help in the fabrication of new inorganic materials and the development of spatially confined and highly controllable synthetic approaches, but may also provide us with new understanding of the chemical, physicochemical and biochemical processes that occur in a confined environment at the nanometer scale, which is significant for the fundamental understanding of nature and life. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles References 1. M. Boutonnet, J. Kizling and P. Stenius, Coll. Surf. 5(3) (1982) 209–225. 2. D. O. Yener and H. Giesche, J. Am. Ceram. Soc. 84(9) (2001) 1987–1995. 3. A. Nanni and L. Dei, Langmuir 19 (2003) 933–938. 4. M. P. Pileni, A. Hammounda, I. Lisiecki, L. Motte, N. Moumen and J. Tanori, Control of the size and shape of nanoparticles, in Fine Particles Science and Technology, ed. E. Pelizzetti (Kluwer Academic Publishers, 1996), pp. 413–429. 5. L. M. Gan, L. H. Zhang, H. S. O. Chan, C. H. Chew and B. H. Loo, J. Mater. Sci. 31 (1996) 1071–1079. 6. F. J. Arriagada and K. Osseo-Asare, J. Coll. Interface Sci. 211 (1999) 210–220. 7. Y. X. Chen, X. Z. Zhang, K. Zheng, S. M. Chen, Q. C. Wang and X. X. Wu, Enzyme Microbial Technol. 23(3–4) (1998) 243–248. 8. P. P. Sun and Y. M. Zhang, Synth. Commun. 27(23) (1997) 4173–4179. 9. K. K. B. Lee, L. H. Poppenborg and D. C. Stuckey, Enzyme Microbial Technol. 23(3–4) (1998) 253–260. 10. M. A. Lopez-Quintela, C. Tojo, M. C. Blanco, L. Garcia Rio and J. R. Leis, Curr. Opin. Coll. Interface Sci. 9 (2004) 264–278. 11. K. Carlile, G. D. Rees, B. H. Robinson, T. D. Steer and M. Svensson, J. Chem. Soc., Faraday Transact. 92(23) (1996) 4701–4708. 12. C. R. Babu, P. F. Flynn and A. J. Wand, J. Biomol. NMR 25(4) (2003) 313–323. 13. E. P. Melo, S. M. B. Costa, J. M. S. Cabral, P. Fojan and S. B. Petersen, Chem. Phys. Lipids 124 (2003) 37–47. 14. G. G. Chang, T. M. Huang and H. C. Hung, Proc. Natl. Sci. Counc. ROC B24(3) (1999) 89–100. 15. A. Tuck, Surv. Geophys. 23 (2002) 379–409. 16. P. A. Bachmann, P. Walde, P. L. Luisi and J. Lang, J. Am. Chem. Soc. 113(22) (1991) 8204–8209. 17. P. A. Bachmann, P. L. Luisi and J. Lang, Chimia 45(9) (1991) 266–268. 18. J. H. Schulman, W. Stoeckenius and L. M. Prince, J. Phys. Chem. 63 (1959) 1677–1680. 19. J. Liu, Y. Ikushima and Z. Shervani, Curr. Opin. Solid State Mater. Sci. 7 (2003) 255–261. 20. M. C. McLeod, R. S. McHenry, E. J. Beckman and C. B. Roberts, J. Phys. Chem. B107(12) (2003) 2693–2700. 21. C. J. O’Connor, V. Kolesnichenko, E. Carpenter, C. Sangregorio, W. Zhou, A. Kumbhar, J. Sims and F. Agnoli, Synth. Met. 122 (2001) 547–557. 22. M. P. Pileni, T. Zemb and C. Petit, Chem. Phys. Lett. 118(4) (1985) 414–420. 23. E. E. Carpenter, C. T. Seip and C. J. O’Connor, J. Appl. Phys. 85(8) (1999) 5184–5186. 271 24. U. Natarajan, K. Handique, A. Mehra, J. R. Bellare and K. C. Khilar, Langmuir 12 (1996) 2670–2678. 25. R. H. Kodama, J. Magn. Magn. Mater. 200 (1999) 359–372. 26. P. Tartaj and C. J. Serna, Chem. Mater. 14 (2002) 4396–4402. 27. K. K. Ghosh and L. K. Tiwary, J. Dispersion Sci. Technol. 22(4) (2001) 343–348. 28. B. Lindman and H. Wennerstrom, Micelles. Amphiphile Aggregation in Aqueous Solution, Topics in Current Chemistry, Vol. 87 (SpringerVerlag, Berlin, 1980). 29. F. Li and C. Vipulanandan and K. K. Moharty, Coll. Surf. A223 (2003) 103–112. 30. S. L. Watt, D. Tunaley and S. Biggs, Coll. Surf. A137 (1998) 25–33. 31. J. Fang, J. Wang, L. M. Gan, S. C. Ng, J. Ding and X. Liu, J. Am. Ceram. Soc. 83(5) (2000) 1049–1055. 32. H. R. Rabie, M. E. Weber and J. H. Vera, J. Coll. Interface Sci. 174 (1995) 1–9. 33. F. Li and C. Vipulanandan, Conductivity of waterin-oil microemulsion system, retrieved from University of Houston website at http://gem1.cive. uh.edu. 34. D. Brown and J. H. R. Clarke, J. Phys. Chem. 92 (1988) 2881–2888. 35. N. E. Levinger, Science 298 (2002) 1722–1723. 36. S. Senapati and M. L. Berkowitz, J. Chem. Phys. 118(4) (2003) 1937–1944. 37. D. Brown and J. H. R. Clarke, J. Phys. Chem. 92(10) (1988) 2881–2888. 38. S. Brunetti, D. Roux, A. M. Bellocq, G. Fourche and P. Bothorel, J. Phys. Chem. 87(6) (1983) 1028–1034. 39. V. K. Aswal and P. S. Goyal, Physica B245 (1998) 73–80. 40. C. C. Co, E. W. Kaler and S. R. Kline, Microstructure transformation during microemulsion and micellar polymerizations, Research Highlights in the monthly review paper of NIST Center for Neutron Research, world wide web. 41. V. K. Aswal, Bhabha Atom. Res. Centre Newslett. 237 (2003). 42. S. Tamura, K. Takeuchi, G. Mao, R. Csencsits, L. Fan, T. Otomo and M. L. Saboungi, J. Electroanal. Chem. 559 (2003) 103–109. 43. N. V. Sushkin, Ph.D. dissertation, Worcester Polytechnic Institute (1997). 44. Q. Zhong, D. A. Steinhurst, E. E. Carpenter and J. C. Owrutsky, Langmuir 18 (2002) 7401–7408. 45. G. Y. Xu, L. Zhang, S. L. Yuan, X. R. Huang and G. Z. Li, J. Dispersion Sci. Technol. 22(6) (2001) 563–567. 46. T. Patzlaff, M. Janich, G. Seifert and H. Graener, Chem. Phys. 261 (2000) 381–389. June 2, 2005 12:33 00700 272 V. Uskoković & M. Drofenik 47. R. H. Cole, S. Mashimo and P. Winson, J. Chem. Phys. 84(7) (1980) 786–793. 48. B. Gestblom and E. Noreland, J. Chem. Phys. 81(8) (1977) 782–788. 49. J. van Stam, S. Depaemelaere and F. C. De Schryver, J. Chem. Edu. 75(1) (1998) 93–98. 50. M. P. Pileni, J. Phys. Chem. 97 (1993) 6961–6973. 51. F. Reiss-Husson and V. Luzzati, J. Phys. Chem. 68(12) (1964) 3504–3511. 52. F. Li, G. Z. Li, H. Q. Wang and Q. J. Xue, Coll. Surf. A127 (1997) 89–96. 53. X. Fang and C. Yang, J. Coll. Interface Sci. 212 (1999) 242–251. 54. G. Lindblom, B. Lindman and L. Mandell, J. Coll. Interface Sci. 34(2) (1970) 262–271. 55. T. Kato and T. Seimiya, J. Phys. Chem. 90 (1986) 3159–3167. 56. L. Wang and R. E. Verrall, J. Phys. Chem. 98 (1994) 4368–4374. 57. I. Benito, M. A. Garcia, C. Monge, J. M. Saz and M. L. Marina, Coll. Surf. A125 (1997) 221–224. 58. J. J. Rack, T. M. McCleskey and E. R. Birnbaum, J. Phys. Chem. B106 (2002) 632–636. 59. Z. Lin, J. J. Cai, L. E. Scriven and H. T. Davis, J. Phys. Chem. 98 (1994) 5984–5993. 60. T. Kinugasa, A. Kondo, S. Nishimura, Y. Miyauchi, Y. Nishii, K. Watanabe and H. Takeuchi, Coll. Surf. A204 (2002) 193–199. 61. G. Montelvo, M. Valiente and E. Rodenas, Langmuir 12 (1996) 5202–5208. 62. D. Roux, A. M. Bellocq and P. Bothorel, Prog. Coll. Polymer Sci. 69 (1984) 1–11. 63. E. Rodenas and M. Valiente, Coll. Surf. 62 (1992) 289–295. 64. J. Kim and M. Lee, J. Phys. Chem. A103(18) (1999) 3378–3382. 65. P. Ekwall, L. Mandell and P. Solyom, J. Coll. Interface Sci. 35(2) (1970) 267–272. 66. Y. Liu, W. Wang, Y. Zhan, C. Zheng and G. Wang, Mater. Lett. 56 (2002) 496–501. 67. C. Liu, A. J. Rondinone and Z. J. Zhong, Pure Appl. Chem. 72(1–2) (2000) 37–45. 68. C. R. Vestal and Z. J. Zhang, J. Solid State Chem. 175 (2003) 59–62. 69. A. E. Giannakas, A. K. Ladavos and P. J. Pomonis, Appl. Catal. B49 (2004) 147–158. 70. D. Walsh and S. Mann, Chem. Mater. 8(8) (1996) 1944–1953. 71. T. Hirai, S. Hariguchi, I. Komasawa and R. J. Davey, Langmuir 13(25) (1997) 6650–6653. 72. H. Zhang and A. I. Cooper, Chem. Mater. 14 (2002) 4017–4020. 73. J. C. Linehan, J. L. Fulton and R. M. Bean, Process of forming compounds using reverse micelle or reverse microemulsion systems, US Patent 5,170,172 (1998). 74. C. H. Lu and H. C. Wang, J. Eur. Ceram. Soc. 23 (2003) 865–871. 75. C. Liu, B. Zuo, A. J. Rondinone and Z. J. Zhang, J. Phys. Chem. 104(6) (2000) 1141–1145. 76. S. Santra, R. Tapec, N. Theodoropoulou, J. Dobson, A. Hebard and W. Tan, Langmuir 17 (2001) 2900–2906. 77. F. Grasset, N. Labhsetwar, D. Li, D. C. Park, N. Saito, H. Haneda, O. Cador, T. Roisnel, S. Mornet, E. Duguet, J. Portier and J. Etourneau, Langmuir 18(21) (2002) 8209–8216. 78. D. S. Bae, K. S. Han and J. H. Adair, J. Mater. Chem. 12(10) (2002) 3117–3120. 79. J. Fang, N. Shama, L. D. Tung, E. Y. Shin, C. J. O’Connor, K. L. Stokes, G. Caruntu, J. B. Wiley, L. Spinu and J. Tang, J. Appl. Phys. 93(10) (2003) 7483–7485. 80. J. Wang, P. F. Chong, S. C. Ng and L. M. Gan, Mater. Lett. 30 (1997) 217–221. 81. T. Liu, L. Guo, Y. Tao, Y. B. Wang and W. D. Wang, Nanostruct. Mater. 11(4) (1999) 487–492. 82. Y. Y. Luk and N. L. Abbott, Curr. Opin. Coll. Interface Sci. 7 (2002) 267–275. 83. E. E. Carpenter, C. Sangregorio and C. J. O’Connor, Mol. Cryst. Liq. Cryst. 334 (1999) 641–649. 84. J. A. Johnson, M. L. Saboungi, P. Thiyagarajan, R. Csecsits and D. Meisel, J. Phys. Chem. B103 (1999) 59–63. 85. X. Li, C. K. Loong, P. Thiyagarajan, G. A. Lager and R. Miranda, J. Appl. Crystalog. 33(1) (2000) 628–631. 86. X. M. Sui, Y. Chu, S. X. Xing and C. Z. Liu, Mater. Lett. 58(7–8) (2004) 1255–1259. 87. R. Hua, C. Zang, C. Shao, D. Xie and C. Shi, Nanotechnology 14 (2003) 588–591. 88. F. Porta, C. Bifulco, P. Fermo, C. L. Bianchi, M. Fadoni and L. Prati, Coll. Surf. A160 (1999) 281–290. 89. A. E. Giannakas, T. C. Vaimakis, A. K. Ladavos, P. N. Trikalitis and P. J. Pomonis, J. Coll. Interface Sci. 259 (2003) 244–253. 90. F. Chen, G. Q. Xu and T. S. A. Hor, Mater. Lett. 57 (2003) 3282–3286. 91. G. Palazzo, F. Lopez, M. Giustini, G. Colafemmina and A. Ceglie, J. Phys. Chem. B107 (2003) 1924–1931. 92. D. H. Chen and Y. Y. Chen, J. Coll. Interface Sci. 236 (2001) 41–46. 93. J. P. Wilcoxon and P. P. Provencio, J. Phys. Chem. B103 (1999) 9809–9812. 94. V. Uskoković and M. Drofenik, Coll. Surf. A, submitted, 2005. 95. Y. Teraoka, S. Nanri, I. Moriguchi, S. Kagawa, K. Shimanoe and N. Yamazoe, Chem. Lett. 10 (2000) 1202–1203. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles 96. M. Mamak, G. S. Metraux, S. Petrov, N. Coombs, G. A. Ozin and M. A. Green, J. Am. Chem. Soc. 125(17) (2003) 5161–5175. 97. K. Kurumada, S. Itakura, S. Nagamine and M. Tanigaki, Langmuir 13(14) (1997) 3700–3705. 98. M. P. Pileni, N. Moumen, J. F. Hochepied, P. Bonville and P. Veillet, J. Phys. IV France 7, C1-505-8 (1997). 99. M. A. Lopez-Quintela, Curr. Opin. Coll. Interface Sci. 8 (2003) 137–144. 100. B. R. Smith and A. E. Alexander, J. Coll. Interface Sci. 34(1) (1970) 81. 101. T. Tago, Y. Shibata, T. Hatsuta, K. Miyajima, M. Kishida, S. Tashiro and K. Wakabayashi, J. Mater. Sci. 37 (2002) 977–982. 102. T. Tago, T. Hatsuta, K. Miyajima, M. Kishida, S. Tashiro and K. Wakabayashi, J. Am. Ceram. Soc. 85(9) (2002) 2188–2194. 103. H. Shi, L. Qi, J. Ma, H. Cheng and B. Zhu, Adv. Mater. 15(19) (2003) 1647–1651. 104. B. A. Simmons, S. Li, V. T. John, G. L. McPherson, A. Bose, W. Zhou and J. He, Nano Lett. 2(4) (2002) 263–268. 105. P. Somasundaran, J. Coll. Interface Sci. 256 (2002) 3–15. 106. G. Porte, D. Roux, J. F. Berret, S. Lerouge, J. P. Decruppe, P. Lindner and L. Noirez, Rheology of wormlike micelles, ITP Complex Fluid Program 3/27/02 (CNRS-Rhodia, 2002). 107. R. Zana and J. Lang, Dynamics of Microemulsions, in Microemulsions: Structure and Dynamics, eds. S. E. Friberg and P. Bothorel (CRC Press: Boca Raton, 1987), pp. 153–172. 108. D. G. Shchukin and G. B. Sukhorukov, Adv. Mater. 16(8) (2004) 671–682. 109. P. van Beurten and A. Vrij, J. Chem. Phys. 74(5) (1981) 2744–2748. 110. F. G. Sanchez and C. C. Ruiz, J. Lumin. 69 (1996) 179–186. 111. Z. Cui and R. J. Mumper, Pharm. Res. 19(7) (2002) 939–946. 112. V. T. John, B. Simmons, G. L. McPherson and A. Bose, Curr. Opin. Coll. Interface Sci. 7 (2002) 288–295. 113. S. Xu, H. Zhou, J. Xu and Y. Li, Langmuir 18 (2002) 10503–10504. 114. J. Yang, Curr. Opin. Coll. Interface Sci. 7 (2002) 276–281. 115. M. Bhat, Studies in surfactant science and hydrotropy, Ph.D. Thesis, University Institute of Chemical Technology, Mumbai, India. 116. R. I. Nooney, D. Thirunavukkarasu, Y. Chen, R. Josephs and A. E. Ostafin, Chem. Mater. 14 (2002) 4721–4728. 117. D. Makovec, A. Košak and M. Drofenik, Nanotechnol. 15 (2004) S160–S166. 273 118. A. M. Goncalves, A. P. Serro, M. R. Aires-Barros and J. M.S. Cabral, Biochim. Biophys. Acta 1480 (2000) 92–106. 119. E. E. Carpenter, Ferrites: Proc. 8th Int. Conf. Ferrites (ICF 8), Kyoto/Tokyo, Japan (2000). 120. V. Pillai and D. O. Shah, J. Magn. Magn. Mater. 163 (1996) 243–248. 121. J. Sjöblom, R. Lindberg and S. E. Friberg, Adv. Coll. Interface Sci. 95 (1996) 125–287. 122. S. Bhattacharjee, M. Elimelech and M. Borkovec, Croatica Chem. Acta 71(4) (1998) 883–903. 123. D. G. Grier, Nature 393(6686) (1998) 621. 124. T. Dichristina, D. Roux, A. M. Bellocq and P. Bothorel, J. Phys. Chem. 89(8) (1985) 1433–1437. 125. B. Lindman and P. Stilbs, in Microemulsions: Structure and Dynamics, eds. S. E. Friberg and P. Bothorel (CRC Press, Boca Raton, 1987), 119–152. 126. J. R. Hansen, J. Phys. Chem. 78 (1974) 256. 127. H. F. Eicke and P. E. Zinsli, J. Coll. Interface Sci. 65 (1978) 131. 128. M. Wong, J. K. Thomas and T. Nowak, J. Am. Chem. Soc. 99 (1977) 4730. 129. B. Lindman and H. Wennerstrom, in Solution Behavior of Surfactants, Vol. 1, eds. K. L. Mittal and E. J. Fendler (Plenum Press, New York, 1982). 130. H. Grunhagen, J. Coll. Interface Sci. 53 (1975) 282. 131. J. Biais, B. Clin, P. Lalanne and B. Lemanceau, J. Chim. Phys. 74 (1977) 1197. 132. H. Yoshioka, J. Coll. Interface Sci. 95 (1983) 81. 133. J. Lang, A. Djavanbakht and R. Zana, J. Phys. Chem. 84 (1980) 145. 134. J. M. DiMeglio, M. Dvolaitzky, R. Ober and C. Taupin, J. Phys. Lett. 44 (1983) L229. 135. E. Bourgeat–Lami, J. Nanosci. Nanotechnol. 2(1) (2002) 1–24. 136. Z. Shervani and Y. Ikushima, J. New Chem. 26 (2002) 1257–1260. 137. R. P. Bagwe and K. C. Khilar, Langmuir 16 (2000) 905–910. 138. B. H. Robinson and P. D. Fletcher, Ber. Bunsenges. Phys. Chem. 85 (1981) 863. 139. R. Jain and A. Mehra, Langmuir 20 (2004) 6507–6513. 140. S. Atik and J. K. Thomas, J. Phys. Chem. 85 (1981) 3921. 141. K. Inouye, R. Endo, Y. Otsuka, K. Miyashiro, K. Kaneko and T. Ishikawa, J. Phys. Chem. 86(8) (1982) 1465–1469. 142. J. Fang, J. Wang, S. C. Ng, C. H. Chew and L. M. Gan, Nanostruct. Mater. 8(4) (1997) 499–505. 143. L. Motte, C. Petit, L. Boulanger, P. Lixon and M. P. Pileni, Langmuir 8(4) (1992) 1049–1053. June 2, 2005 12:33 00700 274 V. Uskoković & M. Drofenik 144. J. Legrand, A. T. Ngo, C. Petit and M. P. Pileni, Adv. Mater. 13(1) (2001) 58–62. 145. K. Kandori, K. Kon-No and A. Kitahara, J. Coll. Interface Sci. 122(1) (1988) 78–82. 146. H. Sato and I. Komasawa, J. Chem. Eng. Jpn. 33(2) (2000) 262–266. 147. V. Uskoković and M. Drofenik, J. Magn. Magn. Mater. 284 (2004) 294–302. 148. A. Košak, D. Makovec, A. Žnidaršič and M. Drofenik, J. Eur. Ceram. Soc. 24 (2004) 959–962. 149. A. Agostiano, M. Catalano, M. L. Curri, M. Della Monica, L. Manna and L. Vasanelli, Micron 31 (2000) 253–258. 150. C. T. Seip and C. J. O’Connor, Nanostruct. Mater. 12 (1999) 183–186. 151. C. J. O’Connor, C. T. Seip, E. E. Carpenter, S. Li and V. T. John, Nanostruct. Mater. 12 (1999) 65–70. 152. A. J. Zarur and J. Y. Ying, Nature 403 (2000) 65–67. 153. A. Košak, D. Makovec and M. Drofenik, Phys. Status Solidi. C1(12) (2004) 3521–3524. 154. A. Košak, D. Makovec and M. Drofenik, J. Metastable Nanocryst. Mater. 23 (2005) 251–254. 155. A. Košak, D. Makovec, M. Drofenik and A. Žnidaršič J. Magn. Magn. Mater. 272–276 (2004) 1542–1544. 156. A. Košak, D. Makovec and M. Drofenik, Mater. Sci. Forum 453–454 (2004) 219–224. 157. V. Uskoković and M. Drofenik, Mater. Sci. Forum 453–454 (2004) 225–230. 158. V. Uskoković and M. Drofenik, Mater. Technol. 37(3–4) (2003) 129–131. 159. D. B. Zhang, L. M. Qi, H. M. Cheng and J. M. Ma, Chin. Chem. Lett. 14(1) (2003) 100–103. 160. J. Fang, L. Stokes, J. Wiemann and W. Zhou, Mater. Lett. 42 (2000) 113–120. 161. V. Uskoković and M. Drofenik, Surf. Rev. Lett. 12 (2005) 97–100. 162. M. C. Moran, A. Pinazo, L. Perez, P. Clapes, M. Angelet, M. T. Garcia, M. P. Vinardell and M. R. Infante, Green Chem. 6 (2004) 233–240. 163. Nano-Crystalline Materials Research, The Newsletter of the New York State Center for Advanced Ceramic Technology 11.3 (1999). 164. A. A. Date and V. B. Patravale, Curr. Opin. Coll. Interface Sci. 9 (2004) 222–235. 165. K. Wormuth, J. Coll. Interface Sci. 241 (2001) 366–377. 166. J. Sjoblom, R. Skurtveit, J. O. Saeten and B. Gestblom, J. Coll. Interface Sci. 141(2) (1991) 329–337. 167. K. Shinoda and H. Takeda, J. Coll. Interface Sci. 32(4) (1970) 642. 168. S. L. Watt, D. Tunaley and S. Biggs, Coll. Surf. A137 (1998) 25–33. 169. C. C. Wang, D. H. Chen and T. C. Huang, Coll. Surf. A189 (2001) 145–154. 170. M. L. Curri, A. Agostiano, F. Mavelli and M. Della Monica, Mater. Sci. Eng. C22 (2002) 423–426. 171. L. Zhang, G. C. Papaefthymiou and J. Y. Ying, J. Appl. Phys. 81(10) (1997) 6892–6900. 172. M. Hasegawa, Langmuir 17 (2001) 1426–1431. 173. J. Kizling and P. Stenius, J. Coll. Interface Sci. 118(2) (1987) 482–492. 174. F. Debuigne, L. Jeunieau, M. Wiame and J. B. Nagy, Langmuir 16 (2000) 7605–7611. 175. E. E. Carpenter, Synthesis of nanoparticles using reverse micelles, www.darpa.mil/dso/future/ metamaterials/presentations/carpenter.pdf. 176. X. M. Sui, Y. Chu, S. X. Xing, M. Yu and C. Z. Liu, Coll. Surf. A251(1–3) (2004) 103–107. 177. S. Vaucher, M. Li and S. Mann, Angew. Chem. Int. Ed. 39 (2000) 1793–1796. 178. D. Kuang, A. Xu, Y. Fang, H. Ou and H. Liu, J. Cryst. Growth 244 (2002) 379–383. 179. A. Courty, C. Fermon and M. P. Pileni, Adv. Mater. 13(4) (2001) 254–258. 180. C. E. Bunker, B. A. Harruff, P. Pathak, A. Payzant, L. F. Allard and Y. P. Sun, Langmuir 20 (2004) 5642–5644. 181. D. K. Kim, Y. Zhang, W. Voit, K. V. Rao and M. Muhammed, J. Magn. Magn. Mater. 225 (2001) 30–36. 182. N. C. Tansil, R. V. Ramanujan and H. F. Li, Trans. Ind. Inst. Met. 56(5) (2003) 509–512. 183. J. C. Lin, J. T. Dipre and M. Z. Yates, Chem. Mater. 15 (2003) 2764–2773. 184. D. C. Steytler, A. Gurgel and R. Ohly, Langmuir 20 (2004) 3509–3512. 185. P. A. Dresco, V. S. Zaitsev, R. J. Gambino and B. Chu, Langmuir 15 (1999) 1945–1951. 186. P. Barnickel, A. Wokaun, W. Sager and H. F. Eickel, J. Coll. Interface Sci. 148(1) (1992) 80–90. 187. D. H. Chen and S. H. Wu, Chem. Mater. 12 (2000) 1354–1360. 188. K. E. Marchand, M. Tarret, J. P. Lechaire, L. Normand, S. Kasztelan and T. Cseri, Coll. Surf. A214 (2003) 239–248. 189. C. Kitchens, M. C. McLeod and B. Roberts, J. Phys. Chem. B107 (2003) 11331–11338. 190. F. Debuigne, J. Cuisenaire, L. Jeunieau, B. Masereel and J. B. Nagy, J. Coll. Interface Sci. 243 (2001) 90–101. 191. R. Singh, J. Doolittle, M. A. George and P. K. Dutta, Langmuir 18(21) (2002) 8193–8197. 192. H. Chakraborty and M. Sarkar, Langmuir 20 (2004) 3551–3558. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles 193. N. Pinna, M. Willinger, K. Weiss, J. Urban and R. Schlogl, Nano Lett. 3(8) (2003) 1131–1134. 194. Q. S. Wu, N. W. Zheng, Y. P. Ding and Y. D. Li, Inorg. Chem. Commun. 5(9) (2002) 671–673. 195. S. A. Safran, Nature Mater. 2 (2003) 71–72. 196. J. Sims, A. Kumbhar, J. Lin, F. Agnoli, E. Carpenter, C. Sangregorio, C. Frommen, V. Kolesnichenko and C. J. O’Connor, Mol. Cryst. Liq. Crys. 379 (2002) 121–130. 197. L. P. Li and G. S. Li, Hyperfine Interact. 128 (2000) 437–442. 198. S. H. Liou and C. L. Chien, Appl. Phys. Lett. 52(6) (1988) 512–514. 199. C. R. Hendricks, V. W. R. Amarakoon and D. Sullivan, Ceram. Bull. 70(5) (1991) 817–823. 200. Y. S. Cho, D. Schaffer, V. L. Burdick and V. R. W. Amarakoon, Mater. Res. Bull. 34(14/15) (1999) 2361–2368. 201. K. Mandal, S. P. Mandal, P. Agudo and M. Pal, Appl. Surf. Sci. 182 (2001) 386–389. 202. K. G. Brooks and V. R. W. Amarakoon, J. Am. Ceram. Soc. 74(4) (1991) 851–853. 203. F. A. Selmi and V. R. W. Amarakoon, J. Am. Ceram. Soc. 71(11) (1988) 934–937. 204. A. Chatterjee, D. Das, S. K. Pradhan and D. Chakravorty, J. Magn. Magn. Mater. 127 (1993) 214–218. 205. G. Beckingham, Aerogels: their history, structure and application, http://members.tripod.com/ ∼geobeck/frontier/aerogels.html. 206. F. J. Arriagada and K. Osseo-Asare, Coll. Surf. A154 (1999) 311–326. 207. A. Van Blaaderen, J. Van Geest and A. Vrij, J. Coll. Interface Sci. 154(2) (1992) 481–501. 208. F. Bentivegna, J. Ferre, M. Nyvlt, J. P. Jamet, D. Imhoff, M. Canva, A. Brun, P. Veillet, S. Visnovsky, F. Chaput and J. P. Boilot, J. Appl. Phys. 83(12) (1998) 7776–7788. 209. S. Ponce-Castaneda, J. R. Martinez, F. Ruiz, S. Palomares-Sanchez and O. Dominguez, J. SolGel Sci. Technol. 25 (2002) 29–36. 210. J. Wiggins, E. E. Carpenter and C. J. O’Connor, J. Appl. Phys. 87(9) (2000) 5651–5653, Part 2. 211. J. Lin, W. Zhou, A. Kumbhar, J. Wiemann, J. Fang, E. E. Carpenter and C. J. O’Connor, J. Solid State Chem. 159 (2001) 26–31. 212. K. Raj, B. Moskowitz and R. Casciari, J. Magn. Magn. Mater. 149 (1995) 174–180. 213. J. A. Lopez-Perez, M. A. Lopez-Quintela, J. Mira and J. Rivas, IEEE Trans. Magn. 33(5) (1997) 4359–4362. 214. A. Kosturiak, J. Polavka, L. Valko, J. Slama, A. Gruskova and M. Miglierini, J. Magn. Magn. Mater. 153 (1996) 184–188. 215. N. S. Kommareddi, M. Tata, V. T. John, G. L. McPherson, M. F. Herman, Y. S. Lee, 216. 217. 218. 219. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 275 C. J. O’Connor, J. A. Akkara, and D. L. Kaplan, Chem. Mater. 8 (1996) 801–809. M. Lal, N. D. Kumar, M. P. Joshi and P. N. Prasad, Chem. Mater. 10 (1998) 1065–1068. Y. Deng, L. Wang, W. Yang, S. Fu and A. Elaissari, J. Magn. Magn. Mater. 257 (2003) 69–78. H. H. Freedman, J. P. Mason and A. I. Medalia, J. Org. Chem. 23 (1958) 76–82. M. Summers and J. Eastoc, Adv. Coll. Interface Sci. 100–102 (2003) 137–152. T. Hirai, T. Watanabe and I. Komasawa, J. Phys. Chem. B104 (2000) 8962–8966. A. Hammouda, T. Gulik and M. P. Pileni, Langmuir 11 (1995) 3656–3659. O. P. Yadav, A. Palmqvist, N. Cruise and K. Holmberg, Coll. Surf. A221 (2003) 131–134. T. Hanaoka, T. Hatsuta, T. Tago, M. Kishida and K. Wakabayashi, Appl. Catal. A190 (2000) 291–296. R. Touroude, P. Girard, G. Maire, J. Kizling, M. Boutonnet, M. Kizling and P. Stenius, Coll. Surf. 67 (1992) 9–19. D. H. Chen and C. J. Chen, J. Mater. Chem. 12 (2002) 1557–1562. M. L. Wu, D. H. Chen and T. C. Huang, Langmuir 17 (2001) 3877–3883. X. Zhang and K. Y. Chan, Chem. Mater. 15 (2003) 451–459. X. Zhang and K. Y. Chan, J. Mater. Chem. 12 (2002) 1203–1206. W. E. Stallings and H. H. Lamb, Langmuir 19 (2003) 2989–2994. J. Fang, K. L. Stokes, J. A. Wiemann, W. L. Zhou, J. Dai, F. Chen and C. J. O’Connor, Mater. Sci. Eng. B83 (2001) 254–257. J. P. Cason and C. B. Roberts, J. Phys. Chem. B104 (2000) 1217–1221. S. Patil, S. C. Kuiry, S. Seal and R. Vanfleet, J. Nanoparticle Res. 4(5) (2002) 433–438. J. A. Rodriguez, J. C. Hanson, J. Y. Kim and G. Liu, J. Phys. Chem. B107 (2003) 3535–3543. A. B. Hungria, A. Martinez-Arias, M. FernandezGarcia, A. Iglesias-Juez, A. Guerrero-Ruiz, J. J. Calvino, J. C. Conesa and J. Soria, Chem. Mater. 15(22) (2003) 4309–4316. M. Fernandez-Garcia, A. Martinez-Arias, A. B. Hungria, A. Iglesias-Juez, J. C. Conesa and J. Soria, Phys. Chem. Chem. Phys. 4(11) (2002) 2473–2481. A. Martinez-Arias, M. Fernandez-Garcia, A. B. Hungria, J. C. Conesa and G. Munuera, J. Phys. Chem. B107(12) (2003) 2667–2677. A. Martinez-Arias, M. Fernandez-Garcia, A. B. Hungria, J. C. Conesa and J. Soria, J. Alloy. Compound. 323 (2001) 605–609. M. Fernandez-Garcia, A. Martinez-Arias, A. Iglesias-Juez, A. B. Hungria, J. A. Anderson, June 2, 2005 12:33 00700 276 239. 240. 241. 242. 243. 244. 245. 246. 247. 248. 249. 250. 251. 252. 253. 254. 255. 256. 257. 258. 259. 260. 261. V. Uskoković & M. Drofenik J. C. Conesa and J. Soria, Appl. Catal. B31(1) (2001) 39–50. T. Kida, G. Q. Guan and A. Yoshida, Chem. Phys. Lett. 371(5–6) (2003) 563–567. D. B. Dadyburjor, T. E. Fout and J. W. Zondlo, Catal. Today 63(1) (2000) 33–41. A. J. I. Ward, E. C. O’Sullivan, J. C. Rang, J. Nedeljkovic and R. C. Patel, J. Coll. Interface Sci. 161(2) (1993) 316–320. T. Hirai, H. Sato and I. Komasawa, Ind. Eng. Chem. Res. 33 (1994) 3262. C. B. Murray, D. J. Norris and M. G. Bawendi, J. Am. Chem. Soc. 115(19) (1993) 8706–8715. C. M. Bender, J. M. Burtlich, D. Baber et al., Chem. Mater. 12(7) (2000) 1969–1976. M. I. Freedhoff, W. Chen, J. M. Rehm, C. Meyers, A. Marchetti and G. McLendon, in Fine Particles Science and Technology, ed. E. Pelizzetti (1996), pp. 281–293. T. Sugimoto and K. Kimijima, J. Phys. Chem. B107 (2003) 10753–10759. S. G. Dixit, A. R. Mahadeshwar and S. K. Haram, Coll. Surf. A133(1–2) (1998) 69–75. S. Q. Qiu, J. X. Dong and G. X. Chen, Powder Technol. 113(1–2) (2000) 9–13. E. E. Foos, R. M. Stroud and A. D. Berry, Nano Lett. 1(12) (2001) 693–695. F. T. Quinlan, J. Kuther, W. Tremel, W. Knoll, S. Risbud and P. Stroeve, Langmuir 16(8) (2000) 4049–4051. J. B. Nagy, I. Bodart-Ravet and E. G. Derouane, Faraday Discuss. 87 (1989) 189–198. J. B. Nagy, Coll. Surf. 35(2–4) (1989) 201–220. A. K. Panda, S. P. Moulik, B. B. Bhowmik and A. R. Das, J. Coll. Interface Sci. 235 (2001) 218–226. S. P. Moulik, G. C. De, A. K. Panda, B. B. Bhowmik and A. R. Das, Langmuir 15 (1999) 8361–8367. Y. Berkovich, A. Aserin, E. Wachtel and N. Garti, J. Coll. Interface Sci. 245 (2000) 58–67. M. Wu, Y. D. Zhang, S. Hui, T. D. Xiao, S. Ge, W. A. Hines and J. I. Budnick, J. Appl. Phys. 92(1) (2002) 491–495. D. S. Bae, K. S. Han and J. H. Adair, J. Am. Ceram. Soc. 85(5) (2002) 1321–1323. D. S. Bae, K. S. Han and J. H. Adair, J. Mater. Sci. Lett. 21(1) (2002) 53–54. D. S. Bae, S. W. Park, K. S. Han and J. H. Adair, Met. Mater. Int. 7(4) (2001) 399–402. T. Tago, S. Tashiro, Y. Hashimoto, K. Wakabayashi and M. Kishida, J. Nanoparticles Res. 5 (2003) 55–60. K. M. K. Yu, C. M. Y. Yeung, D. Thompsett and S. C. Tsang, J. Phys. Chem. B107 (2003) 4515–4526. 262. C. Vestal and Z. J. Zhang, Nano Lett. 3 (2003) 1739–1743. 263. S. S. Hong, M. S. Lee, S. S. Park and G. D. Lee, Catal. Today 87(1–4) (2003) 99–105. 264. P. Tartaj and L. C. De Jonghe, J. Mater. Chem. 10(12) (2000) 2786–2790. 265. C. J. O’Connor, J. A. Sims, A. Kumbhar, V. L. Kolesnichenko, W. L. Zhou and J. A. Wiemann, J. Magn. Magn. Mater. 226–230 (2001) 1915–1917. 266. A. Kumbhar, L. Spinu, F. Agnoli, K. Y. Wang, W. L. Zhou and C. J. O’Connor, IEEE Trans. Magn. 37(4) (2001) 2216–2218. 267. T. Kinoshita, S. Seino, K. Okitsu, T. Nakayama, T. Nakagawa and T. A. Yamamoto, J. Alloy. Compound. 359 (2003) 46–50. 268. S. A. Gomez-Lopera, R. C. Plaza and A. V. Delgado, J. Coll. Interface Sci. 240 (2001) 40–47. 269. T. Hirai, J. Y. Mizumoto, S. Shiojiri and I. Komasawa, J. Chem. Eng. Jpn. 30(5) (1997) 938–943. 270. L. M. Qi, J. M. Ma, H. M. Cheng and Z. G. Zhao, Chin. Chem. Lett. 6(11) (1995) 1013–1016. 271. M. A. Lopez-Quintela, J. Rivas, M. C. Blanco and C. Tojo, in Nanoscale Materials, eds. L. M. L. Marzan and P. V. Kamat, (Kluwer Academic Plenum Publ, 2003), pp. 135–155. 272. Y. Ni, X. Ge, H. Liu, Z. Zhang and Q. Ye, Chem. Lett. 9 (2001) 924–925. 273. M. Summers, J. Eastoe and S. Davis, Langmuir 18 (2002) 5023–5026. 274. P. Ayyaub, A. N. Maitra and D. O. Shah, Physica C168(5–6) (1990) 571–579. 275. P. Ayyub and M. S. Multani, Mater. Lett. 10(9–10) (1991) 431–436. 276. F. Li and C. Vipulanandan, IEEE Trans. Appl. Supercond. 13(2) (2003) 3196–3198. 277. Q. S. Wu, N. W. Zheng and Y. P. Ding, Chem. J. Chin. Uni. 22(6) (2001) 898–900. 278. L. Guo, Z. Wu, T. Liu, W. Wang and H. Zhu, Chem. Phys. Lett. 318 (2000) 49–52. 279. L. M. Qi, J. M. Ma, H. M. Cheng and Z. G. Zhao, J. Phys. Chem. B101(18) (1997) 3460–3463. 280. Y. Liu, C. Zheng, W. Wang, Y. Zhan and G. Wang, J. Cryst. Growth 233 (2001) 8–12. 281. X. Y. Dong, X. T. Zhang, G. Cheng, Y. C. Li and Z. L. Du, Acta Chim. Sin. 62(24) (2004) 2441–2443. 282. J. D. Hopwood and S. Mann, Chem. Mater. 9(8) (1997) 1819–1828. 283. X. Mo, C. Wang, L. Hao, M. You, Y. Zhu, Z. Chen and Y. Hu, Mater. Res. Bull. 36 (2001) 1925–1930. 284. H. H. Yang, H. Y. Qu, P. Lin, S. H. Li, M. T. Ding and J. G. Xu, Analyst 128(5) (2003) 462–466. 285. Z. Q. Ye, M. Q. Tan, G. L. Wang and J. L. Yuan, J. Mater. Chem. 14(5) (2004) 851–856. June 2, 2005 12:33 00700 Synthesis of Materials within Reverse Micelles 286. Z. Q. Ye, M. Q. Tan, G. L. Wang and J. L. Yuan, Anal. Chem. 76(3) (2004) 513–518. 287. S. Bose and S. K. Saha, Chem. Mater. 15 (2003) 4464–4469. 288. J. Z. Liu, B. D. Gou, S. J. Xu and K. Wang, Prog. Natural Sci. 12(8) (2002) 615–617. 289. S. Sarda, M. Heughebaert and A. Lebugle, Chem. Mater. 11(10) (1999) 2722–2727. 290. A. J. Zarur, H. H. Hwu and J. Y. Ying, Langmuir 16 (2000) 3042–3049. 291. C. J. Barbe, R. Graf, K. S. Finnie, M. Blackford, R. Trautman and J. R. Bartlett, J. Sol-Gel Sci. Technol. 26 (2003) 457–462. 292. G. D. Rees, R. G. Evans, S. J. Hammond and B. H. Robinson, Langmuir 15(6) (1999) 1993–2002. 293. L. Sun, Y. Zhang, J. Zhang, C. Yan, C. Liao and Y. Lu, Solid State Commun. 124 (2002) 35–38. 294. M. H. Lee, S. G. Oh and S. C. Yi, J. Coll. Interface Sci. 226(1) (2000) 65–70. 295. T. Hirai, T. Hirano and I. Komasawa, J. Coll. Interface Sci. 253(1) (2002) 62–69. 296. B. Delfort, L. Normand, P. Dascotte and L. Barre, J. Coll. Interface Sci. 207(2) (1998) 218–227. 297. B. Delfort, A. Chive and L. Barre, J. Coll. Interface Sci. 186(2) (1997) 300–306. 298. K. S. N. Reddy, L. M. Salvati, P. K. Dutta, P. B. Abel, K. I. Suh and R. R. Ansari, J. Phys. Chem. 100(23) (1996) 9870–9880. 299. K. Kon-no, in Fine Particles Science and Technology, ed. E. Pelizzetti (1996), pp. 431–42. 300. J. Y. Liang, GuoLin, X. H. Bin, L. Jing, L. X. Dong, W. Z. Hua, W. Z. Yu and J. Weber, J. Cryst. Growth 252 (2003) 226–229. 301. G. Wakefield, H. A. Keron and P. J. Dobson et al. J. Coll. Interface Sci. 215(1) (1999) 179–182. 302. K. C. Song and J. H. Kim, J. Coll. Interface Sci. 212(1) (1999) 193–196. 303. M. S. Lee, G. D. Lee and S. S. Hong, J. Indust. Eng. Chem. 9(4) (2003) 412–418. 304. Z. Wu, J. Zhang, R. E. Benfield, Y. Ding, D. Garndjean, Z. Zhang and X. Ju, J. Phys. Chem. B106 (2002) 4569–4577. 305. Y. X. Pang and X. Bao, J. Mater. Chem. 12 (2002) 3699–3704. 306. P. Tartaj and J. Tartaj, Chem. Mater. 14 (2002) 536–541. 307. Z. You, I. Balint and K. Aika, Chem. Lett. 11 (2002) 1090–1091. 308. I. Balint, Z. You and K. Aika, Phys. Chem. Chem. Phys. 4 (2002) 2501–2503. 309. M. Hayashi, H. Uemura, K. Shimanoe, N. Miura and N. Yamazoe, Electrochem. Solid State Lett. 1(6) (1998) 268–270. 277 310. V. Uskoković, D. Makovec and M. Drofenik, Mater. Sci. Forum 494 (2005) 155–160. 311. V. Uskoković and M. Drofenik, J. Mater. Sci., submitted (2005). 312. I. Ban, M. Drofenik and D. Makovec, Mater. Sci. Forum 494 (2005) 161–166. 313. S. A. Morrison, C. L. Cahill, E. E. Carpenter, S. Calvin and V. G. Harris, J. Appl. Phys. 93(10) (2003) 7489–7491. 314. M. Drofenik, D. Lisjak and D. Makovec, Mater. Sci. Forum 494 (2005) 129–136. 315. M. Han, R. Vestal and Z. J. Zhang, J. Phys. Chem. B108 (2004) 583–587. 316. C. R. Vestal and Z. J. Zhang, Chem. Mater. 14(9) (2002) 3817–3822. 317. M. Han, R. Vestal and Z. J. Zhang, J. Phys. Chem. B108 (2004) 583–587. 318. S. Bandow, K. Kimura, K. Kon-no and A. Kitahara, Jpn. J. Appl. Phys. 26(5) (1987) 713–717. 319. C. Y. Wang, W. Q. Jiqng, Y. Zhou, Y. N. Wang and Z. Y. Chen, Mater. Res. Bull. 35(1) (2000) 53–58. 320. F. Meyer, A. Dierstein, C. Beck, W. Hartl, R. Hempelmann, S. Mathur and M. Veith, Nanostruct. Mater. 12 (1999) 71–74. 321. P. Kluson, P. Kacer, T. Cajthaml and M. Kalaji, J. Mater. Chem. 11(2) (2001) 644–651. 322. J. Huang, Y. Xie, B. Li, Y. Liu, J. Lu and Y. Qian, J. Coll. Interface Sci. 236 (2001) 382–384. 323. K. Mallick, Z. L. Wang and T. Pal, J. Photochem. Photobiol. A140(1) (2001) 75–80. 324. G. C. Koumoulidis, A. P. Katsouilidis, A. K. Ladavos, P. J. Pomonis, C. C. Trapalis, A. T. Sdoukos and T. C. Vaimakis, J. Coll. Interface Sci. 259(2) (2003) 254–260. 325. M. Bonini, U. Bardi, D. Berti, C. Neto and P. Baglioni, J. Phys. Chem. B106 (2002) 6178–6183. 326. H. S. Wu and E. W. Kaler, Microemulsion polymerization systems and coated materials made therefrom, US Patent 5,539,047 (1996). 327. H. Althues and H. Kaskel, Langmuir 18 (2002) 7428–7435. 328. H. Ago, S. Ohshima, K. Uchida, T. Komatsu and M. Yumura, Physics B323 (2002) 306–307. 329. H. Ago, S. Ohshima, K. Tsukuagoshi, M. Tsuji and M. Yumura, Curr. Appl. Phys. 5(2) (2005) 128–132. 330. J. L. Zhang, Z. M. Liu, B. X. Han, T. Jiang, W. Z. Wu, J. Chen, Z. H. Li and D. X. Liu, J. Phys. Chem. B108(7) (2004) 2200–2204.