You are on page 1of 12

View Online

TUTORIAL REVIEW

www.rsc.org/csr | Chemical Society Reviews

Microemulsion-based synthesis of nanocrystalline materials


Ashok K. Ganguli,* Aparna Ganguly and Sonalika Vaidya
Received 23rd February 2009 First published as an Advance Article on the web 22nd September 2009 DOI: 10.1039/b814613f Microemulsion-based synthesis is found to be a versatile route to synthesize a variety of nanomaterials. The manipulation of various components involved in the formation of a microemulsion enables one to synthesize nanomaterials with varied size and shape. In this tutorial review several aspects of microemulsion based synthesis of nanocrystalline materials have been discussed which would be of interest to a cross-section of researchers working on colloids, physical chemistry, nanoscience and materials chemistry. The review focuses on the recent developments in the above area with current understanding on the various factors that control the structure and dynamics of microemulsions which can be eectively used to manipulate the size and shape of nanocrystalline materials.

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

1. Introduction
The large number of unusual properties and applications associated with nanomaterials has triggered enormous interest among scientists from varied elds of research especially due to the interdisciplinary nature of this subject. Nanoscience and nanotechnology today is practised by chemists, biologists, physicists, material scientists and engineers who have put in tremendous eorts to understand new phenomena and develop technologies in this eld. A major contribution to the development of this eld has been made by chemists working primarily on the theme to design and control of nanostructures and also to functionalize them using both low-temperature solution-based routes and high-temperature

Department of Chemistry, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India. E-mail: ashok@chemistry.iitd.ernet.in; Fax: 91-11-26854715; Tel: 91-11-26591511

(thermodynamic) methods. The microemulsion method is one among the various low-temperature routes to tailor nanoparticles. The term microemulsion was rst coined by J. H. Schulman in 1959,1 and since then its use has grown considerably and has received justied acclaim from the nanomaterials community. There have been several important reviews published on this subject especially during 19932006 by Pileni,2 Eastoe,3 Lopez-Quintela,4 Capek,5 Holmberg,6 and Uskokovic.7 In this review, we have given a brief introduction to the concepts and principles involved in microemulsions and their applications to nanomaterial synthesis. We then build on the information available in the previous reviews and focus on the developments in the past ten years, especially discussing the current understanding on the various factors controlling the structure and dynamics of microemulsions, and their manipulation to control the synthesis of nanocrystalline powders and related systems. In spite of the signicant work carried out earlier in this eld, many aspects of microemulsion

Prof. Ashok Kumar Ganguli obtained his PhD from the Indian Institute of Science, Bangalore in 1990. He subsequently worked at Dupont Company, Wilmington, USA and Ames Laboratory, Iowa State University, USA. before joining IIT Delhi in 1995 where currently he is a full professor. His interests are in the synthesis and properties of nanocrystalline materials, complex metal oxides with dielectric and superconducting Ashok K. Ganguli properties and polar intermetallics. He has published over 125 papers in international journals and around 15 in conference proceedings and books, and was awarded the Materials Research Society of India Medal for 2006, and the Chemical Research Society of India medal for 2007.
474 | Chem. Soc. Rev., 2010, 39, 474485

Aparna Ganguly

Ms Aparna Ganguly obtained her BSc (Hons) in chemistry from Sri Venkateshwara College, University of Delhi in 2002. Later she obtained her MSc in chemistry from University of Delhi with specialisation in physical chemistry in 2004. Currently she is working as a joint PhD student of Prof. A. K. Ganguli (IIT Delhi) and Dr T. Ahmad (Jamia Millia Islamia) on microemulsion routes to synthesize functionalised nanostructures.

This journal is

 c

The Royal Society of Chemistry 2010

View Online

based synthesis are yet to be understood, to cater to the increasing demands of precisely tailored nanomaterials. Hence a lot of excitement prevails among scientists to explore its vast horizon. This review attempts to put into perspective the understanding available at present, and to foresee future developments, which may become popular and meaningful in the coming years. 1.1 Colloidal solutions

The use of colloidal gold had started from the ancient Roman period to colour glasses red, mauve or yellow by varying the concentration of gold. Paracelsus, the alchemist of 16th century is claimed to be the rst to have prepared a gold colloid (Aurum Potabile). Inspired by his work, Michael Faraday8 prepared a gold colloid solution in 1857. The nely divided gold particles exhibited dierent optical properties which Faraday recognized as being dependent on the size of the particles. Colloidal solutions have since become a topic of intense research due to their interesting properties and applicability. To prevent the particles from aggregating, stabilizers such as citrate ions are added which are adsorbed on the surface of the particles inducing a surface charge and hence repulsion from other particles to prevent agglomeration. This is a kind of electrostatic stabilisation. Steric stabilisation can be achieved with bulky organic molecules being present on the metal surface providing a protective shield, as is the case with surfactants. The stabilizer should coordinate to the particle strongly enough to prevent agglomeration but should also be easily removable from the metal surface. Colloidal solutions have found application in synthesis of novel materials, plastics and ceramics, and more recently in nanotechnology. Due to their optical and electronic properties they nd use as biosensors, especially gold colloids, which are being extensively studied for this purpose.9 Among other important applications, silver colloids have been found to have anti-bacterial activity, which is being exploited in textiles. 1.2 Surfactant aggregates

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

immiscible phases. They are mostly organic molecules with a polar head group (hydrophilic) and a long alkyl chain (hydrophobic part). Depending on the size of these two chains, an empirical number, HydrophilicLipophilic Balance (HLB), has been assigned to the surfactants. It is a measure of the degree to which it is hydrophilic or lipophilic. Grin proposed an HLB scale for non-ionic surfactants and the HLB number 1 was assigned to the most lipophilic molecule while 20 was assigned to the most hydrophilic molecule. Various methods have been described in literature to calculate the HLB number of the surfactants. For instance, HLB values of polyhydric alcohol fatty acid esters can be calculated using eqn (1) in which Mh is the weight of the hydrophobic group and Mw is the molecular weight.   Mh HLB 20 1 Mw 1

For fatty acid esters (Tween type), the HLB value can be calculated using eqn (2) where E is the weight percentage of oxyethylene and P is the weight percentage of polyhydric alcohol. HLB = (E + P)/5 (2)

The word surfactant is derived from surface active agent and is known to reduce the interfacial tension between two

Sonalika Vaidya

Ms Sonalika Vaidya obtained her BSc (Hons) in chemistry from Hindu College, University of Delhi in 2002. She obtained second position in the university at her undergraduate level. She has been awarded the Rastogi Award in all the three years of her undergraduate studies for holding rst position at college level. Later she obtained her MSc in Chemistry from IIT Delhi in 2004 after which she joined for her PhD and is currently working on core shell nanostructures in Professor Gangulis group.

Pasquali et al. in their studies has developed dierent equations to calculate the HLB number for various kinds of surfactants.10 The HLB value of the surfactant depends on its structure and thus decides its action in the solution. The application of the surfactant can be predicted from its HLB number, for example w/o type of emulsions can be formed using a surfactant with low HLB number while o/w emulsions can be formed with surfactants having a high HLB number. The other factor which is important when surfactants are discussed is the critical micellar concentration (CMC). At low concentrations, the surfactant dissolves in the aqueous phase but when the concentration exceeds the critical micellar concentration (CMC), the surfactant molecules organize spontaneously to form aggregates such as micelles, vesicles etc. Formation of such micelles is an entropy driven process. Water molecules in the liquid state can be considered to have a 3-D structure of hydrogen bonds similar to ice with cavities. In liquid water, there is always an equilibrium existing between the destruction and formation of hydrogen bonds, which results in movement of free water molecules through cavities. In the presence of a hydrocarbon, the cavities are occupied by the hydrocarbon molecules that results in restricted movement of water; consequently water molecules surrounding the hydrophobic solute become more ordered. During micellization, there is a transfer of non-polar surfactant chains from an ordered aqueous environment to the hydrocarbon-like environment of the micelles, resulting in the disordering of water molecules surrounding the non-polar molecules, thereby increasing the entropy of the system and stabilizing the microemulsion. In a micelle the hydrophobic tail of the surfactant points towards the core while the polar head group forms an outer shell. Such an assembly maintains a favourable contact with water. Micelles can thus solubilize signicant amounts of non-polar molecules attributed to the hydrophobic core inside. Similarly, surfactants or amphiphiles may
Chem. Soc. Rev., 2010, 39, 474485 | 475

This journal is

 c

The Royal Society of Chemistry 2010

View Online

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

aggregate in non-polar organic solvents (bulk), with less amount of water, wherein the structural organization is just the opposite of what is observed for micelles and thus they are referred to as reverse or inverse micelles. The aqueous interior of these reverse micelles allows the dissolution of polar moieties. The third factor associated with surfactant is the surfactant packing parameter (Ns) which depends mainly on the volume of the polar head group and the length of the hydrocarbon chain and is given mathematically by the ratio Ns = V/alc where, V is the eective hydrocarbon volume, a is the surface area of the headgroup of the surfactant, and l is the fully extended chain length. These parameters control the varying forces which play a key role in the formation of surfactant aggregates. The collective eect of the forces acting simultaneously on dierent molecules (water, surfactant and oil) and also on dierent parts of the same molecule ultimately determine the structure of the microemulsions. The eect is more pronounced on the surfactant molecules as they are located at the interface of the immiscible oil and water mixture. Repulsive hydrophilic forces on the head group of the amphiphile are balanced by the attractive hydrophobic forces acting at the waterhydrocarbon interface and the repulsive steric forces between the chains. The surface area of the headgroup of the surfactant, a, can be determined by the rst two forces. The steric chainchain and oil penetration interactions acting within the hydrocarbon interior determine the eective hydrocarbon volume, V, and fully extended chain length l. By simple geometry, the critical radius of curvature R can be determined in which the molecules pack together within the aggregate. Calculations of packing parameter require constraints to be included for proper treatment of the assembly of the surfactant aggregates.11 For a spherical micelle the radius of the hydrocarbon core, R is given by eqn (3). R = 3V/a (3)

prohibited, giving a critical condition for the formation of sphere as V/alc = 1/3. For cylinders, planar bilayers and inverse aggregates, this parameter is 0.5, 1 and 41, respectively. Fig. 1 shows schematic representations of various surfactant aggregates. The knowledge of the factors discussed above (HLB number, CMC and Ns) enables one to choose surfactants for desired applications especially in the synthesis of nanomaterials with controlled size and shape. 1.3 Microemulsions A microemulsion is a thermodynamically stable dispersion of two immiscible liquids in the presence of an emulsier or surfactant. They are characterized by ultra-low interfacial tension, large interfacial area and capacity to solubilize both water and oil components. Microemulsions are of use in oil recovery, pharmaceutics, cosmetics, detergency, lubrication etc. They are categorized as water-in oil (w/o) microemulsions when the water is dispersed homogenously in an organic media with the help of the surfactant and oil-in-water (o/w) microemulsions, where oil is dispersed in water. The water-insc-CO2 (sc = supercritical) microemulsion is another class of microemulsion added more recently.3 Though a microemulsion appears to be homogenous macroscopically; distinct phases can be seen at the microscopic levels. Among the various classes of microemulsion, the w/o microemulsion has been extensively studied. These are important due to their application in the synthesis of inorganic nanoparticles. There is a subtle dierence between the terms w/o microemulsion and reverse (inverse) micelles. Fig. 2 shows a schematic diagram of a typical reverse micelle. Aggregates containing a small amount of water (below Wo = 15; where Wo = [H2O]/[surfactant]) are usually called reverse micelles whereas microemulsions correspond to droplets containing a large amount of water (Wo 4 15).2 Since the reverse micellar region is stabilized only in certain regions of the ternary phase diagram, a complete understanding of the phase diagram is required to exploit its advantages for synthesis purposes. Keeping the relative concentration of any two constituents (oil/surfactant/water/co-surfactant) xed, a four-component pseudo-ternary phase diagram (Gibbs triangle) can be obtained using the titration method. Three parameters are required to dene completely the four component microemulsion system, [water]/[surfactant] ratio (Wo), [co-surfactant]/[surfactant]

Since the radius of a spherical micelle cannot exceed a certain critical length, lc (fully extended chain length of the hydrocarbon), so from this equation, it can be deduced that when V/alc 4 1/3, the formation of spherical micelles is

Fig. 1 Schematic representation of organized aggregates of surfactants: (a) normal micelles, (b) reverse micelles, (c) cylindrical micelle, (d) planar lamellar phases, (e) onion-like lamellar phases and (f) interconnected cylinders.

Fig. 2

A typical structure of a reverse micelle in a non-polar phase.

476 | Chem. Soc. Rev., 2010, 39, 474485

This journal is

 c

The Royal Society of Chemistry 2010

View Online

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

ratio (Po) and [solvent]/[surfactant] ratio (No). Information regarding the dierent surfactant aggregates forming the phases can be obtained by using experimental tools such as small angle neutron scattering (SANS), small angle X-ray scattering (SAXS), NMR self diusion, freeze fracture TEM (FFTEM) or conductivity measurements. To preserve the w/o microemulsion structure, it is critical to determine the water emulsication failure boundary (wefb). It is near the failure boundary where the formation of discrete water droplets in a continuous oil phase is expected. The boundary (wefb) is a measure of the maximum amount of water that can be solubilised in the oil phase at a constant temperature. Here the saturated w/o microemulsion coexists with the excess of water. We will discuss in detail how various parameters can be controlled for the design of the nanostructures. From the point of view of applications, reverse micelles are important due to their role as nanoreactors where the synthesis of nanoparticles can be carried out with sucient precision by controlling the components of the ternary/quaternary system.
Fig. 3 Mechanism showing the intermicellar exchange for the formation of nanoparticles.

1.4

Microemulsions as nanoreactors tr (occurring inside the reverse micelles) is also critical. The ratio tr/tex determines the kinetics of the chemical reaction inside the micelle. An encounter rate factor, g, depending on the lm exibility, aects the exchange rate constant kex. This value varies from 103 for a rigid interface (for AOT) to 101 for more exible lms. Thus, the characteristic exchange time, tex for the reverse micelles fall in the range of 10 ms o tex o 1 ms. This rate can be controlled via the interfacial uidity of the surfactant membrane which will be discussed later. Simulations to elucidate mechanisms aecting the droplet exchange, growth and size have been studied for a better understanding.16 From the simulations we understand that both the droplet exchange and the chemical reaction are important to understand the underlying mechanism. A variety of experiments may be designed in order to determine the intermicellar exchange rate. Quenching of a cytochrome in presence of dyes can be studied for the determination of the exchange rate which is based on the fact that quenching is exchange limited. The rate is dependent on the water content and the type of quencher used. Other techniques which measure the micellar diusion coecient are light scattering, quasi-elastic light scattering, neutral scattering, voltametry, and 1H pulse-gradient-stimulated-echo NMR. Based on the use of w/o microemulsions our group has successfully synthesized metal nanoparticles,17 dielectric15 and magnetic oxides,18,19 and more recently ternary and quaternary oxides such as Ca- and Sr-doped LaMnO3.20 The methodology involves the use of as many microemulsions as the number of reacting ions. For example in the synthesis of Sr doped LaMnO3, a quaternary oxide, four microemulsions are required. Three of these microemulsions contain a metal ion each and the fourth contains the precipitating agent. An alternative to this method for binary compounds is the single microemulsion method. One of the desired reactants is solubilised in the reverse micelle while the other one is added directly to it. A variety of nanomaterials have been synthesized by the above microemulsion methods and studied for their
Chem. Soc. Rev., 2010, 39, 474485 | 477

Microemulsion-based synthesis is a powerful method where expensive or specialized instruments are not needed, contrasting to the case for several physical methods such as plasma synthesis, ball milling, chemical vapour deposition etc. The product obtained is microhomogeneous as the desired stoichiometry is maintained inside the water pools. Metallic nanoparticles,12,13 semiconductor quantum dots,14 polymeric nanoparticles,3 ceramics15 etc are a few examples of nanomaterials synthesized using reverse micelles. The reverse micelles collide among themselves to exchange the reactants and then again break apart. This coalescence process is critical since it is only through this mechanism that the reactants, solubilised in individual reverse micelles (nanoreactors), come in close contact and undergo homogenous mixing. While decoalescence ensures the presence of the protective coating of the amphiphile for the controlled nucleation and growth it also prevents aggregation. On mixing the microemulsions, the reverse micelles containing the reactants, collide with each other forming a water channel which results in the formation of a transient dimer. Once such a dimer is formed, intermicellar exchange of the reactants take place and thus nucleation starts at the micellar edges with the well known growth process from the boundary to core. It is known that most ionic reactions are very fast compared to the lifetime of a dimer and hence the reaction starts instantly which can account for the nucleation starting at the micellar edge. Further growth occurs around this point, with more reactant fed in via intermicellar exchange. The boundary for the core growth mechanism was experimentally shown by Li et al. using TEM and has been illustrated in a review by Eastoe and co-workers.3 These reverse micelles (referred to as nanoreactors) favor the formation of small crystallites with a narrow size distribution. A schematic diagram of the reaction dynamics for a binary system is given in Fig. 3. The intermicellar exchange rate can be characterized by a parameter, tex, which is specic to the type of microemulsion chosen.4 Along with intermicellar exchange time, the time required for the chemical reaction,
This journal is
 c

The Royal Society of Chemistry 2010

View Online

properties.3 Single molecule magnets or coordination polymers are yet another class of nanomaterials which have been successfully obtained by this method. Uniform cubes of size B15 nm of Prussian blue were synthesized with the help of AOT reverse micelles.21 The ability to ne tune the particle size and morphology by using this method is largely due to the various parameters which can be altered to provide exibility for suitable tailoring of the products.

2. Synthesis of nanocrystalline materials using reverse micelles


Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

2.1

Metal nanoparticles

Boutonnet et al. in 1982 used for the rst time reverse micelles as a template to synthesize metal nanoparticles of Pt, Rh, Pd and Ir.12 Many dierent materials comprising deagglomerated and monodispersed metal nanoparticles have been synthesized thereafter due to their interesting optical, magnetic and electrical properties. Colloids of coinage metals such as Au, Ag and Cu have especially been of interest due to their interesting optical properties attributed to the surface plasmon resonance. Since the surface plasmon resonance (SPR) of Au lies in the visible region and is easy to visualize, the chemistry of gold nanoparticles and its applications (imaging, biosensing etc.) have been of tremendous interest. In an earlier review on metal nanoparticles by Capek et al.,5 some aspects of the synthesis of nanomaterials using microemulsions and the parameters (such as the reducing agent concentration/water content) that aect the nal particle size for the metal nanoparticles, have been discussed. In an investigation by Pileni et al. on metallic Pd nanoparticles, transformation of spherical nanoparticles to worm-like nanostructures was observed on increasing the water content.13 Recently Boutonnet has reviewed22 the developments in the microemulsion synthesis of nanoparticles especially for catalytic applications. Detailed investigation by Isabelle23 on metal nanoparticles elucidates the role of water content, capping agent, and concentration of reducing agent on the shape and size of copper nanoparticles.23 Control over the size of gold nanoparticles (from 2.2 to 6.6 nm) formed by using AOT-based reverse micelles has been achieved by controlling the reaction temperature from 15 to 40 1C24 (Fig. 4). The advantages oered by the microemulsion method over the other methods have been highlighted in the study on the synthesis of nanosized particles.25 In order to explore the scope of other polar organic solvents as the reaction media, methanol has been employed instead of water in an AOT/heptane system.26 Though solvation dynamics (through steady-state absorption and uorescence spectroscopy) and light scattering studies have been carried out on some of these reverse micelles, there exists a need to better understand the design and stability of such complex microemulsion systems, which are necessary for more intelligent tailoring of nanostructures. It should be noted that the properties of microemulsions with polar organic cores do not depend on the Wo value, making them very dierent in comparison to those with aqueous cores. Bimetallic alloy nanoparticles exhibit many improved properties over their single counterparts, which makes them commercially
478 | Chem. Soc. Rev., 2010, 39, 474485

Fig. 4 Transmission electron micrographs of gold nanoparticles obtained at dierent temperatures. Reprinted with permission from ref. 24. Copyright 2007, American Chemical Society.

important, and nanoparticles of Fe/Pt,27 and Cu/Ni17 have also been synthesized in low sizes using the microemulsion method. The scope of the microemulsion method can thus be expanded and employed for a wide range of metal (elemental) and alloy nanostructures. 2.2 Metal carboxylate nanorods An important and challenging area of research in nanoscience and nanotechnology is to obtain anisotropic nanostructures (nanorods, nanobres and nanowires) which have received increasing interest due to their potential applications in nanodevices. The template method is very eective for the fabrication of one-dimensional nanostructures of a desired material. Two dierent classes of templates are normally dened (hard and soft templates). Carbon nanotubes and porous alumina fall in the category of hard templates and can be used to control the size, shape and alignment of the synthesized material. The soft template method uses various types of microemulsions and micelles in which the reaction is restricted inside the micellar core and the shape and size can be tuned by the structure of the polar core. Most of such syntheses lead to spherical particles, which however under certain conditions may aggregate in the form of ellipsoids, cubes, triangles or higher-order nanostructures (needles, bres and nanotubes). In addition single-crystalline anisotropic nanostructures may also be obtained by the microemulsion method. Although nanorods or nanowires have smaller surface area as compared with nanoparticles, they exhibit a great advantage in device fabrication. Soft chemistry using surfactants and bidentate ligands has been exploited to synthesize such one-dimensional nanostructures. Nanorods of transition-metal oxalates21,28 have been obtained by the reverse micellar route using CTAB (cetyltrimethylammonium bromide) as the surfactant. Details of these nanorods have been given in Table 1. It was observed that the metal oxalate rods have a negative surface charge. It is thus proposed that the formation of these nanorods is facilitated by the templating eect of the cationic surfactant (CTAB) molecules, which align themselves on the linear arrangement of the metal and the ligand formed, as shown schematically in Fig. 5. It is observed that in the absence of the cationic surfactant, only spherical particles could be obtained. Most of these nanorods
This journal is
 c

The Royal Society of Chemistry 2010

View Online

spherical particles (not rods) with the divalent carboxylate ligand.15,18 The metalcarboxylate nanorods were found to be suitable precursors for the formation of metal and metal oxide nanoparticles. 2.3 Metal oxide nanoparticles Metal oxide nanoparticles have applications in various industries which include paints, pigments, cosmetics, batteries, electronics, pharamaceutics, magnetic and optical devices. Microemulsion synthesis of oxides, especially used as heterogeneous catalysts, has been studied in detail.22 Zarur et al. developed processes to synthesize nanocrystalline metal (non noble) oxides of high surface area as catalysts for combustion. Despite the dierent hydrolysis rates of the metal alkoxides, chemical homogeneity could be maintained in presence of the reverse micelles.29 Microemulsions have also been used to obtain ultra-low sized (B5 nm) tungsten oxide with high surface area at a much lower temperature compared to the conventional techniques.30 Eastoe et al.31 have obtained nanocrystalline oxides using mixed surfactants (DDAB and Brij 35) systems. The mixed surfactant systems were found to have improved thermal stability and large water solubilisation capacity of the microemulsion compared to the commonly used anionic surfactant, AOT. The stability of the microemulsion structure on addition of additives such as mono-, di- and trivalent metal ions, along with high concentrations of the precipitating agents, was studied from scattering studies. The microemulsion structure remains unperturbed with both soluble and insoluble additives. A complete study elucidating the eect of mixing time, surfactant concentration, water-to-surfactant molar ratio and precursor concentration on copper oxide nanoparticles was carried out by Nassar et al.32 The ionic strength of the water pools and the degree of interaction of the surfactant head group was found to aect the uptake of nanoparticles.32 The increase in the nanoparticle uptake was attributed to higher occupancy number, coupled with a rigid interface which promotes the intermicellar nucleation and growth, resulting in higher uptake. The reverse-micellar approach was used to synthesize simple binary oxides such as CeO2,15 ZrO2,15 Fe2O328 to complex ternary oxide nanoparticles such as BaTiO3,15 SrZrO3,15 LaMnO320 etc. Our group has been actively involved in the synthesis of ternary metal oxides with interesting dielectric properties such as strontium titanates15 (SrTiO3 and Sr2TiO4), barium titanates15 (BaTiO3 and Ba2TiO4), lead titanate15 (PbTiO3) and many other titanates and zirconates15 using the microemulsion method with Tergitol as a surfactant (non-ionic). The route developed avoids the use of expensive alkoxides such as Ba-alkoxides, Sr-alkoxide etc. The particle size for BaTiO3 at 900 1C was found to be B35 nm (Fig. 6) A weak tetragonal distortion was concluded based on Raman studies (inset of Fig. 6) and a weak ferroelectric transition was observed in the dielectric measurements. The dielectric constant (e) was found to increase with sintering temperature for BaTiO3 being 520 after sintering at 1100 1C. The dielectric constant for 3040 nm sized particles of SrTiO3 was found to be 90 with a dielectric loss of 0.08 at 100 kHz. Dielectric oxides such as BaZrO3, SrZrO3, PbZrO3
Chem. Soc. Rev., 2010, 39, 474485 | 479

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

Fig. 5 Mechanism depicting the formation of nanorods in the presence of a cationic surfactant. The TEM image shows nanorods of nickel oxalate dihydrate. Table 1 M Cu Ni Mnb Zn Co Fe
a

Transition metal oxalate hydrate (MC2O4nH2O) nanorods Diameter/nm 130 250 100 120 300 70 Length/nm 480 2500 2500 600 6500 470 TN/K TIPa 45 15 21 27

TIP = temperature independent paramagnetism. b n = 0 (anhydrous).

exhibit properties similar to their bulk counterparts with lowering of the magnetic transition temperature (Table 1). A TN of 45 K has been observed for nickel oxalate dihydrate, contrary to the bulk value of 50 K, and a transition at 15 K for manganese oxalate dihydrate which is somewhat higher than the bulk value of 2.6 K. Detailed analysis of the above studies show that the 1 : 1 stoichiometric ratio of ligand (carboxylate) to the metal ion is critical for the formation of nanorods. Metal ions with higher oxidation states (Ce3+, Zr4+) require larger content of the carboxylate ions (41 : 1) and result in
This journal is
 c

The Royal Society of Chemistry 2010

View Online

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

Fig. 6 TEM micrograph of BaTiO3 sintered at 900 1C. The inset shows the Raman spectrum of 35 nm BaTiO3 indicating weak tetragonal distortion.

Fig. 7 TEM image of PbSe QDs in silica spheres. The inset shows magnication of the same sample. Reprinted with permission from ref. 33. Copyright 2008, American Chemical Society.

and a solid solution of Ba1xPbxZrO3 were synthesized at low temperature.15 An increase in grain size was observed with increase in lead content for Ba1xPbxZrO3 (0 r x r 1). The increase in dielectric constant with Pb content was also observed up to x = 0.50, however, a further increase in lead content decreases the dielectric constant. Using metal carboxylates as precursors (optimized in our laboratory) several commercially important binary oxides have been obtained. Nanorods of cobalt and iron oxalate dihydrate were decomposed in dierent environments (nitrogen, hydrogen and vacuum) to obtain the nanoparticles of various cobalt19 (Co3O4, CoO and metallic Co nanoparticles) and iron oxides28 (Fe2O3 and Fe3O4). Thus the microemulsion method proves to be a versatile route for the synthesis of various types of ceramics (including ternary and quaternary oxides). 2.4 Coreshell nanostructures

in the water content. The diameter was also found to increase with the precursor concentration. An increase of the absorption and photoluminescence intensity with increase in ZnS shell thickness was observed which suggested better quantum eciency of the coreshell. There are numerous examples in the literature where w/o microemulsions have been used to synthesize magnetic coreshell37 nanostructures and bimetallic coreshells.38 Our group has recently shown an increase in the PL intensity indicating surface enhanced Raman scattering (SERS) activity for Ag@TiO2 coreshell nanostructures synthesized using w/o microemulsions. The homogeneity of the shell was more pronounced when the shell forming agent was changed from titanium isopropoxide to titanium hydroxyacylate.39 2.5 Metal chalcogenide nanoparticles Pileni and co-workers are among the rst few to synthesize metal chalcogenides using microemulsions.14 Since then, microemulsions have been extensively used to synthesize quantum dots, as the size distribution of these nanoparticles can be successfully controlled. Of all the known chalcogenides, CdS has been most conveniently synthesized using reverse micelles and widely studied for its properties. CdSe is another important semiconductor widely used in photoelectric devices. Nanorods of CdSe have been synthesized using the anionic surfactant AOT.40 A more recent study shows that quantum dots of 1-hexanethiolate capped a-Cu2xSe, a superionic conducting phase, were synthesized at room temperature in Triton X-100 water-in-oil microemulsions.41 2.6 Interesting morphologies using reverse micelles Pileni and Eastoe in their reviews have thrown light on the various parameters that govern particle size and shape.2,3 The micellar template is found to inuence the particle growth. Along with the control over the size of the nanoparticle synthesized, a control over the morphology is equally viable. Compositional changes in the water-in-oil surfactant systems can thus bring about a change in the shape of the surfactant aggregates. Spherical (reverse micelles or micelles), cylinders, interconnected cylinders and planes, termed as lamellar phases, can thus be obtained. From Pilenis illustrations on Ag and Cu nanocrystals it is certainly evident that water-in-oil
This journal is
 c

The water-in-oil (W/O) microemulsion system in conjunction with the Stober synthesis and silane coupling method has been used for the preparation of silica-coated, metallic, magnetic and semiconductor nanocrystals. Synthesis of coreshell nanostructures with semiconducting cores enables one to overcome the drawback that arises due to the presence of bent surfaces, dangling bonds, photo-oxidation. Dierent approaches have been followed by scientists to form shell over semiconducting nanoparticles. The rst approach is to coat these nanoparticles with an insulator such as silica or titania. Recently silica coating over PbSe33 (Fig. 7), quantum dots of CdS and CdTe34 and coreshell shell nanostructures of CdSe@ZnS@silica35 have been synthesized. The coating of silica causes a red shift in the PL maximum of the semi conductor. The second approach involves coating the semiconductor with another semiconductor of wide band gap. The most widely studied coreshell material of the above type is the CdS/ZnS coreshell nanostructures.36 The shell thickness was varied by using two dierent methodologies, one by increasing water content, at a given precursor concentration and second by increasing the precursor concentration at a given water content. It was found that the nanoparticle diameter and the shell thickness increase linearly with increase
480 | Chem. Soc. Rev., 2010, 39, 474485

The Royal Society of Chemistry 2010

View Online

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

microemulsions can be used to obtain specic shapes. Additives such as NaCl and KCl favour specic adsorption on the facets of the nuclei leading to various shapes. Not only the concentration but also the type of ions play a major role. For instance chloride ions induce remarkable changes during the growth process of elongated copper crystals whereas I and F ions do not cause any change in the nanocrystal shape.23 The ionic selective adsorption is a complex process not only related to the crystallographic nature of the surface but also related to the surface energies, that vary with the precursor involved, which give rise to varied morphologies. The coupled eect of self assembly and the modied synthesis leading to the designing of dierent types of nanostructures of barium chromate has been reported by Mann et al.42 They report the formation of ordered prismatic nanoparticles, nanolaments and a two-dimensional superlattice by simply altering the molar ratio of reactants. The synthesis of anisotropic structures of metal dicarboxylates with the aid of cationic surfactants has been studied in detail.19,28 Rods, cubes and spheres of Ni-oxalate were designed by choosing appropriate microemulsion systems.43 On varying the oxidation state of the metal ion from +2 to +3

and the ligand (succinate instead of oxalate) led to, spherical nanoparticles of iron succinate pentahydrate18 as opposed to the rod shaped nanostructures obtained for the corresponding metal oxalate. An unusual sh-bone type nanostructured BaWO4 has also been synthesized44 (Fig. 8(a)). Other unusual variations of structures have been achieved by controlling the surfactant aggregates. Equilateral nanotriangles of CdS (Fig. 8(b))45 and nanopetals of manganese dioxide46 (Fig. 8(c)) have been designed using the templating eect of the anionic surfactants. Studies have revealed the importance of the reaction conditions on the morphology of the nal product. The concentration of the surfactant, water content and reaction temperature were found to have a pronounced eect on PbWO4 nanostructures (spheres, ellipsoids, bipyramids and rod-like bundles) (Fig. 8(d)(f)) using the same anionic surfactant.47 Hexabranched germanium oxide with carambola shape has been synthesized with the help of surfactant template, however, the authors have reasoned such formation owing to its crystal structure, the role of microemulsion is still limited in controlling its size.48

3. Control of size and shape of nanocrystalline structures


There have been may recent reviews in the literature which focus on the parameters aecting the size and shape of reverse micelles and thereby their role in governing the size and shape of the synthesized product.24,6,7,25,49 Michaels et al. in their review49 have given a quantitative model on the dependence on size of non-ionic reverse micelles as a function of molecular structure of the surfactant, the type of oil, the total concentration of surfactant [NP], the ratio of surfactant to total surfactant (r), the water to surfactant molar ratio (o), temperature, salt concentration, and polar phase. It is to be noted that there are no simple rules that decide the morphology of the synthesized product. However, the lm rigidity of the reverse micelles can be controlled by changing the components involved in the formation of w/o microemulsions, which consequently gives nanomaterials of dierent size and shape. Several mechanisms have been proposed which reect the roles of solvent and surfactant in controlling the size and shape of the designed nanomaterials.44,50,51 In this review we have concentrated on recent developments in the synthesis of nanomaterials that are governed by four major factors viz. Wo, surfactant, solvent and co-surfactant. 3.1 Eect of water-to-surfactant ratio (Wo) The aqueous core of the reverse micelles plays a crucial role in determining the size of the nal product formed. The water pool solubilizes the reactants and provides the stage where the reaction occurs. An equation which relates the radius of the reverse micelles (assuming the water-in-oil droplets to be spherical) with the water content (Wo) is R = 3V/S where R, V and S are the radius, the volume and the surface area of the sphere, respectively. Thus, the particle size can be controlled by varying the aqueous content. The relation of the aqueous core to the surfactant concentration is given by Wo = [H2O]/[surfactant]. By varying the Wo, one eectively varies the concentration of the reactants. The water
Chem. Soc. Rev., 2010, 39, 474485 | 481

Fig. 8 (a) SEM image of BaWO4 shbone-like nanostructures. Reprinted with permission from ref. 44. Copyright 2004, Elsevier. (b) TEM image of nanotriangles of CdS nanocrystals, Reprinted with permission from ref. 45. Copyright 2001, American Chemical Society. (c) SEM image of nanopetals of MnO2, Reprinted with permission from ref. 46. Copyright 2007, Springer-Verlag. FESEM images of (d) ellipsoids, (e) bipyramids and (f) rodlike bundles of PbWO4. Reprinted with permission from ref. 47. Copyright 2004, American Chemical Society.

This journal is

 c

The Royal Society of Chemistry 2010

View Online

solubilisation capacity of reverse micelles increases linearly with an increase in surfactant concentration as reected by the relation Wo = [H2O]/[surfactant]. This was also theoretically proposed by Michaels et al.,49 where they found that the micellar size increases with increase in Wo, provided all the other factors that govern the micellar size are kept constant. The eect of Wo has been discussed in earlier reports.6,52 Uskokovic et al.52 discuss the parameters that govern the size of the reverse micelles. The radius of reverse micelles (r) at constant surfactant concentration S is given by eqn (5). r = (4.98 103)j/AsS
Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

(5)

where j is the volume fraction of the dispersed phase which can be controlled easily and As is the area occupied by the surfactant at the droplet surface. This relation is applicable where the microemulsion structure is not perturbed during the reaction. Pilenis group pioneered the studies2 related to the control of size and shape of nanomaterials by varying Wo. An increase in the size of copper nanoparticles with increasing water content was observed, which saturates at Wo = 1015. Many reports exist that show the dependence of size of the nanoparticles on the Wo parameter, though this correspondence in the increase of particle size with increasing Wo (found by many researchers) cannot be generalized. The surfactant/nanomaterial anity also inuences the particle size. A low anity of the surfactant towards the metal centre implies its ineciency to control the growth, thus the increase in the nal size.23 Recently, the dependence of the size and shape of the copper oxalate nanostructures on the aqueous content of the system was studied wherein the nanostructures were synthesized using a non-ionic surfactant (TX-100). An increase in the particle size was observed with increase in the aqueous content.50 3.2 Nature of surfactants: size and charge

Surfactants play the crucial role in stabilizing the immiscible oil/water phase by lowering the interfacial tension to form microemulsions. The review articles by Pileni,2 and Eastoe et al.3 are an informative source on the experimental and theoretical aspects pertaining to the eect of surfactants in controlling both the size and the shape. A variety of surfactants categorized as cationic, anionic, non-ionic and zwitterionic (having both positive and negative charges) depending on the type of charge on their head group are known. Two of most commonly used surfactants are the anionic surfactant sodium bis(2-ethylhexyl) sulfosuccinate (AOT) and the cationic surfactant cetyltrimethylammonium
Table 2 Classication of surfactants Examples

bromide, (CTAB). Tergitol, TX-100, Igepal, NP-5 are a few common non-ionic surfactants. Dierent types of surfactant along with representative examples are given in Table 2. CTAB (C16H33(CH3)3N+Br) provides a very exible lm and thus a high exchange dynamics is observed in its microemulsions allowing high reactant loading compared to the AOT-based systems. Quite a large number of nanocrystalline materials have been synthesized with desired morphology and size using CTAB as the surfactant. The anionic surfactant, AOT (bis(2-ethylhexyl) sulfosuccinate) has two hydrophobic chains and is widely used in the synthesis of nanomaterials via the microemulsion route. The versatility of this surfactant is mainly due to the large stability region of the reverse micelles in the ternary phase diagram. In the AOT reverse micellar system it has been found that short chain hydrocarbons penetrate the layer of AOT, forming a reverse micellar shell, due to which the inter-micellar exchange is reduced. Long chain hydrocarbons have strong intermolecular interactions and hence get embedded in the AOT layer to form an extra shell which allows easy inter-micellar exchange. The structure of colloidal templates thus depends upon the surfactant used and also controls the nanocrystal growth.2 We have earlier investigated the eect of various surfactants on the morphology of nanomaterials. It is observed that cationic surfactants lead to anisotropy and that nanorods of several divalent metal carboxylates could be obtained.43 Fig. 9(a) shows the formation of nanorods of nickel oxalate synthesized using the cationic surfactant CTAB. An isotropic growth occurs on using the non-ionic surfactant TX-100 leading to spherical nanoparticles of (B5 nm). On changing the non-ionic surfactant from TX-100 to Tergitol, larger cubes (Fig. 9(b)) of size B50 nm are formed. The rigidity of the surfactant plays a crucial role in guiding the morphology of the product formed. The rigidity of the surfactants decreases in the order TX-100 4 Tergitol 4 CTAB. Also it is expected that the positively charged surfactants assemble on the surface of the negatively charged nickel oxalate and thus favor the anisotropic growth (rods). In the absence of such positively charged surfactants, an isotropic growth leads to spheres and nanocubes. Thus the choice of surfactant becomes critical to the size, shape and stability of the particles synthesized. Recent studies53 show the eect of dierent types of surfactant used in controlling the morphology and crystal structure of calcium carbonate. A high concentration of non-ionic surfactant resulted in the formation of oblate sphere-like crystals of vaterite, while reduction in the amount of Brij causes the formation of a mixture of oblate sphere and needle-like crystals of vaterite and aragonite, respectively.53 In the

Surfactant type Cationic Anionic Non-ionic Zwitterionic Natural/biosurfactant Switchable

Cetyltrimethylammonium bromide (C16H33N(CH3)3Br) SDS: Sodium dodecylsulfate (C12H25SO4Na), AOT: Sodium bis(2-ethylhexyl) sulfosuccinate (C20H37O7SNa) Tergitol: C9H19(C6H4)(OCH2CH2)9OH, TX-100: (CH3)3CCH2C(CH3)2C6H4(C2H4O)9.5OH Hexadecylsulfobetaine SB3-16: (C16H33N(CH3)2(CH2)3SO3) Rhamnolipids: e.g. RLL (a-L-rhamnopyranosyl-b-hydroxydecanoyl-b-hydroxydecanoate (C26H48O9) Amidines: e.g. C16H33NC(CH3)N(CH3)2

482 | Chem. Soc. Rev., 2010, 39, 474485

This journal is

 c

The Royal Society of Chemistry 2010

View Online

when the head group was changed to pyridinium ion (cetylpyridinium bromide) instead of ammonium (CTAB) which was attributed to the rigid surfactant layer formed by the restricted orientation of the pyridinium ion. 3.3 Eect of solvent The solvent plays an important role in the assembly of the surfactant molecules and hence plays an important part in the synthesis using w/o microemulsions. This has been discussed by Pileni, Cason, Bagwe and Khilar and reviewed3 by Eastoe and co-workers. The various interactions between the solvent and the surfactant tails control the dynamics of nanoparticle formation. A pronounced eect of the bulkiness of the solvent molecules has been observed on the growth rate of the nanoparticles. This is attributed to the variation in the intermicellar exchange rate which is given by the degree of interaction of the solvent molecules with the surfactant tails. In particular, less bulky solvent molecules with lower molecular volumes such as cyclohexane can penetrate between the surfactant tails, increasing surfactant curvature and rigidity. This increase in rigidity leads to a slower growth rate.43 Isooctane is an example of bulky molecule (large molecular volume) and is unable to penetrate the surfactant tails eciently. This would lead to more uid interface and thus faster growth rates. Cyclohexane is a less bulky solvent and the micellar exchange rate constant is estimated to be lower by a factor of 10. However, there are some conicting reports shown by the studies done on surface rigidity by Eastoe et al., where the solvents have a minimal eect.54 It appears that the conclusions on the basis of lm rigidity and interfacial tension have to be considered along with the role played by the surfactant. A comparative study of microemulsion-mediated synthesis of nickel oxalate using solvents with varying lengths of the hydrophobic tails shows that the length and the aspect ratio of the nanorods are dependent on the bulkiness of the solvent.43 With the increase in bulkiness of the solvent molecule, the length of the nanorod increases. The solvent bulkiness and the intermicellar exchange follows the order n-hexane o cyclohexane o isooctane. Consequently the size of the particles obtained also follows the same order. The role of viscosity of the solvents on the growth kinetics of AgI in AOT reverse micelles has been clearly observed by Spirin et al.55 where the chain length (n) of the solvent molecule has been varied from hexane to dodecane. The rate of formation of nanoparticles is known to depend on the collision frequency and eciency of intermicellar exchange. The collision frequency is dened by the diusion constant K (K = 8000RT/3Z, R is the universal gas constant, T is the temperature and Z the dynamic viscosity of the medium).55 The intermicellar exchange is dependent on the rate of two consecutive processes: (1) formation of the pair of the colliding micelles and (2) formation of a channel between the two for the exchange of reactants. One in 103104 collisions results in an eective exchange. The viscosity of the solvents increases with increasing chain length of the solvent (0.307 cP for hexane to 1.492 cP for dodecane). Thus according to the above formula, a ve-fold decrease in K and hence in the number of the intermicellar collisions, N is expected as we go
Chem. Soc. Rev., 2010, 39, 474485 | 483

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

Fig. 9 TEM micrographs for nickel oxalate dihydrate synthesized using (a) CTAB/1-butanol/n-hexane and (b) Tergitol/1-octanol/ cyclohexane.

presence of DTAB, needle-like crystals in conjunction with very few oblate sphere-like crystals and aggregates of undened shape are seen. In contrast to the above result, the presence of anionic surfactant resulted in aggregates of undened shape.53 Not only the type, but also the surfactant concentration also plays an important role in controlling the shape of the synthesized product. Chen et al. in their study51 have tuned the morphologies of Ni complexes by changing the concentration of the surfactant (AOT). On increasing the AOT content, the shape evolved from spindles to ellipse-like to cuboidal and nally cubes. Such formations were explained on the basis of collisions and fusion of primary spherical particles thereby increasing the intermicellar nucleation rate. As a result, rearrangement of AOT molecules occur, which lead to the assembly of surfactant bilayers between the complex particles and thus the anisotropy. Self assembly and Ostwald ripening further guides the morphology to spindles and cubes. Recently we have also studied that the morphology of copper oxalate monohydrate can be changed from rods to cubes by changing the surfactant from cationic (CTAB) to non-ionic (Tergitol).50 The aspect ratio of the nanorods also depends on the length of the surfactant chain of the cationic surfactant. The length and the diameter of the nanorods were found to decrease with the decrease in the chain length of the cationic surfactant50 from C-16 to C-14. The polar head group also plays an important role in controlling the size of the nanorods.50 Nanorods of lower aspect ratio were obtained
This journal is
 c

The Royal Society of Chemistry 2010

View Online

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

from hexane to dodecane. Thus a decrease in the growth rate with increasing chain length n is expected, which would yield smaller nanoparticles. However, the expected decrease in size from the viscosity relation was not observed experimentally. Instead an increase in particle size was found with increasing chain length n. This increase in size is attributed to the presence of free water in the micellar pools which is further dependent on the micelle size. At constant Wo, the size grows with increasing chain length of the solvent molecule. However, there is practically no free water for the small sized micelles and the rate of particle growth is independent of n. Short chain alkanes penetrate deeper into the micellar shell and spread apart the surfactant molecules. This allows binding of a greater amount of water molecules. As a result the relative amount of free water inside the micelles decreases. Thus the growth is largely controlled by the penetrability of the micellar shells at initial stages while the size depends on the amount of water in the micellar pool for higher Wo values. The particle size of silver nanoparticles increased on changing the solvent from decane to heptane but decreases for cyclohexane.25 The aspect ratio of copper oxalate nanorods also changes when the solvent was changed to cyclohexane and hexane using CTAB when compared to our studies of copper oxalate nanorods using isooctane as the solvent. 3.4 Eect of co-surfactant

For the appropriate packing of amphiphiles (surfactant molecules) at the wateroil interface, surface active substances, in addition to surfactants are often added. These are generally short-chain alcohols or amines commonly referred to as co-surfactants. The role of a co-surfactant is to lower the interfacial tension between oil and water for the spontaneous formation of surfactant aggregates. The addition of co-surfactant is expected to reduce the surfactant concentration in the microemulsion. Low molecular weight alcohols such as butanol, due to its short hydrophobic chain and terminal hydroxyl group, are expected to increase the interaction with surfactant layers at the interface and thus inuence the curvature of the interface and hence the internal energy.56 In a dierent study by Marchand et al.,57 the eect of co-surfactant on the nal particle size has been discussed elaborately. However, contrary to the short chain alcohols or amines generally used as co-surfactants, they have used NP-5 (non-ionic surfactant) instead. The addition of NP-5 in small amounts for the synthesis of MoSx using AOT as a surfactant, leads to a substantial decrease in the average micellar size. This is attributed to the higher uidity of the interfacial lm, and a higher mean curvature of radius, which in turn inuences the intermicellar exchange. A high exchange rate implies the higher consumption of reactants at the nucleation stage, thus reducing the eective concentration for further growth, and results in smaller nanoparticles. Direct implication of the change of co-surfactant on the morphology of ZnS has been studied by Charinpanitkul.56 However, no attempt has been made to explain the results obtained. Curri et al. have studied the eect of pentanol as a co-surfactant on the synthesis of CdS nanoparticles using the CTAB/hexane/ water system.58 A pronounced eect was found on the size,
484 | Chem. Soc. Rev., 2010, 39, 474485

size distribution and stability of crystallites on addition of co-surfactant. Pentanol is expected to increase the lm exibility, thereby aecting the particle growth, and also its absorption on the semiconductor surface stabilizes the particles in solution by acting as a capping agent. The nature as well as the amount of co-surfactant is found to be critical while choosing the appropriate reverse micellar system. From the time-dependent absorption studies for growth of copper nanoparticles, Cason et al. found that at higher Wo, addition of more than 1% octanol as a cosurfactant created an unstable system by increasing the uidity of the interface.59 As a result the system ends up in phase separation with broken micelles. On replacing octanol with 1-benzyl alcohol as a co-surfactant in the same synthesis, an increased growth rate (at low Wo) is observed with the addition of co-solvent, but the terminal particle size is found to decrease. Our studies on the eect of co-surfactant50 on the size of copper oxalate nanorods show that the aspect ratio could be increased to 6.3 : 1 on changing the co-surfactant from 1-butanol to 1-hexanol. Cubes of dimension 80100 nm and nanoparticles of size 810 nm assembling to form nanorods, were obtained with 1-octanol and 1-decanol as the co-surfactant, respectively. The increase in the aspect ratio with increase in chain length is attributed to the decrease in surfactant lm exibility which resulted in rods with high aspect ratio. The change in morphology with further increase in the chain length is possibly due to the eective interaction of the alkyl chain of the co-surfactant with the surfactant tail resulting in a more rigid structure.50

Conclusions
We have reviewed several aspects of microemulsions, their stability, versatility and exibility towards the synthesis of nanostructured materials. The eld has grown considerably from the initial synthesis of spherical metal nanoparticles in the 1980s to the highly complex and multifunctional nanostructures of today. We nd that the understanding of the subject has beneted tremendously due to the availability of several new techniques to follow the dynamics of the processes underlying the synthesis carried out in microemulsions. Since there are several applications in high growth industries such as cosmetics, food and pharmaceuticals, the future of microemulsion based synthesis appears bright. However, there are challenges which will dominate the research in the next decade. The major challenge is the utilization and recycling of the used solvents involved in these microemulsions systems which currently restrict one to certain well understood and common microemulsion systems in industry, even though technically superior microemulsions (developed for academic purposes) may be available. There is of course no doubt for high-end ne chemicals. with stringent size and shape restrictions, for which the microemulsion based synthesis will always be more appropriate. Another direction of future development in this area is foreseen in the use of natural and biosurfactants. Finally the ability to control the release of drugs (pharmaceuticals and cosmeceuticals) under appropriate stimulus will always remain as a major theme in the future of microemulsion based synthesis.
This journal is
 c

The Royal Society of Chemistry 2010

View Online

Acknowledgements
A. K. G. thanks CSIR and DST (Govt. of India) for nancial assistance and IIT Delhi for facilities; S. V. thanks CSIR and A. G. thanks UGC for fellowships.

References
1 T. P. Hoar and J. H. Schulman, Nature, 1943, 152, 102103. 2 M. P. Pileni, J. Phys. Chem. C, 2007, 111, 90199038. 3 J. Eastoe, M. J. Hollamby and L. Hudson, Adv. Colloid Interface Sci., 2006, 128130, 515. 4 M. A. Quintela, C. Tojo, M. C. Blanco, L. Garc a Rio and J. R. Leis, Curr. Opin. Colloid Interface Sci., 2004, 9, 264278. 5 I. Capek, Adv. Colloid Interface Sci., 2004, 110, 4974. 6 K. Holmberg, J. Colloid Interface Sci., 2004, 274, 355. 7 V. Uskokovic and M. Drofenik, Adv. Colloid Interface Sci., 2007, 133, 2334. 8 M. Faraday, Philos. Trans. R. Soc. London, 1857, 147, 145181. 9 S. Chah, M. R. Hammond and R. N. Zare, Chem. Biol., 2005, 12, 323328. 10 R. C. Pasquali, M. P. Taurozzi and C. Bregni, Int. J. Pharm., 2008, 356, 4451. 11 J. N. Israelachvili, D. J. Mitchell and B. W. Ninham, J. Chem. Soc., Faraday Trans. 2, 1976, 72, 15251568. 12 M. Boutonnet, J. Kizling, P. Stenius and G. Maire, Colloids Surf., 1982, 5, 209225. 13 K. Naoe, C. Petit and M. P. Pileni, Langmuir, 2008, 24, 27922798. 14 C. Petit, P. Lixon and M. P. Pileni, J. Phys. Chem., 1990, 94, 15981603. 15 A. K. Ganguli, T. Ahmad, S. Vaidya and J. Ahmed, Pure Appl. Chem., 2008, 80, 24512477. 16 O. Myakonkaya and J. Eastoe, Adv. Colloid Interface Sci., 2009, 149, 3946. 17 J. Ahmed, K. V. Ramanujachary, S. E. Loand, A. Furiato, G. Gupta, S. M. Shivaprasad and A. K. Ganguli, Colloids Surf., A, 2008, 331, 206212. 18 A. Ganguly, R. Kundu, K. V. Ramanujachary, S. E. Loand, D. Das, N. Y. Vasanthacharya, T. Ahmad and A. K. Ganguli, J. Chem. Sci., 2008, 120, 521528. 19 J. Ahmed, T. Ahmad, K. V. Ramanujachary, S. E. Loand and A. K. Ganguli, J. Colloid Interface Sci., 2008, 321, 434441. 20 T. Ahmad, K. V. Ramanujachary, S. E. Loand and A. K. Ganguli, J. Chem. Sci., 2006, 118, 513518. 21 S. Vaucher, M. Li and S. Mann, Angew. Chem., Int. Ed., 2000, 39, 17931796. 22 M. Boutonnet, Curr. Opin. Colloid Interface Sci., 2008, 13, 270286. 23 I. Lisiecki, J. Phys. Chem. B, 2005, 109, 1223112244. 24 A. B. Smetana, J. S. Wang, J. Boeckl, G. J. Brown and C. M. Wai, Langmuir, 2007, 23, 1042910432. 25 W. Zhang, X. Qiao and J. Chena, Mater. Sci. Eng., B, 2007, 142, 115. 26 P. Setua, A. Chakraborty, D. Seth, M. U. Bhatta, P. V. Satyam and N. Sarkar, J. Phys. Chem. C, 2007, 111, 39013907. 27 A. R. Malheiro, L. C. Varanda, J. Perez and H. M. Villullas, Langmuir, 2007, 23, 1101511020. 28 A. K. Ganguli and T. Ahmad, J. Nanosci. Nanotechnol., 2007, 7, 20292035. 29 A. J. Zarur and J. Y. Ying, Nature, 2000, 403, 6567. 30 L. Xiong and T. He, Chem. Mater., 2006, 18, 22112218.

31 A. Bumajdad, J. Eastoe, M. I. Zaki, R. K. Heenan and L. Pasupulety, J. Colloid Interface Sci., 2007, 312, 6875. 32 N. N. Nassar and M. M. Husein, J. Colloid Interface Sci., 2007, 316, 442450. 33 R. Koole, M. M. Schooneveld, J. Hilhorst, C. M. Donega, D. C. Hart, A. Blaaderen, D. Vanmaekelbergh and A. Meijerink, Chem. Mater., 2008, 20, 25032512. 34 Y. Yang, L. Jing, X. Yu, D. Yan and M. Gao, Chem. Mater., 2007, 19, 41234128. 35 M. Darbandi, R. Thomann and T. Nann, Chem. Mater., 2005, 17, 57205725. 36 A. R. Loukanov, C. D. Dushkin, K. I. Papazova, A. V. Kirov, M. V. Abrashev and E. Adachi, Colloids Surf., A, 2004, 245, 914. 37 C.-W. Lu, Y. Hung, J.-K. Hsiao, M. Yao, T.-H. Chung, Y.-S. Lin, S.-H. Wu, S.-C. Hsu, H.-M. Liu, C.-Y. Mou, C.-S. Yang, D.-M. Huang and Y.-C. Chen, Nano Lett., 2007, 7, 149154. 38 I. Lisiecki, M. Walls, D. Parker and M. P. Pileni, Langmuir, 2008, 24, 42954299. 39 S. Vaidya, A. Patra and A. K. Ganguli, J. Nanopart. Res., 2009, DOI: 10.1007/s11051-009-9663-5. 40 L. F. Xi and Y. M. Lam, J. Colloid Interface Sci., 2007, 316, 771778. 41 P. P. Ingole, P. M. Joshi and S. K. Haram, Colloids Surf., A, 2009, 337, 136140. 42 M. Li, H. Schnablegger and S. Mann, Nature, 1999, 402, 393395. 43 S. Vaidya, P. Rastogi, S. Agarwal, S. K. Gupta, T. Ahmad, A. M. Antonelli, K. V. Ramanujachary, S. E. Loand and A. K. Ganguli, J. Phys. Chem. C, 2008, 112, 1261012615. 44 X. Zhang, Y. Xie, F. Xu and X. Tian, J. Colloid Interface Sci., 2004, 274, 118121. 45 N. Pinna, K. Weiss, H. Sack-Kongehl, W. Vogel, J. Urban and M. P. Pileni, Langmuir, 2001, 17, 79827987. 46 S. Devaraj and N. Munichandraiah, J. Solid State Electrochem., 2007, 12, 207211. 47 D. Chen, G. Shen, K. Tang, Z. Liang and H. Zheng, J. Phys. Chem. B, 2004, 108, 1128011284. 48 Y.-W. Chiu and Michael H. Huang, J. Phys. Chem. C, 2009, 113, 60566060. 49 M. A. Michaels, S. Sherwood, M. Kidwell, M. J. Allsbrook, S. A. Morrison, S. C. Rutan and E. E. Carpenter, J. Colloid Interface Sci., 2007, 311, 7076. 50 R. Ranjan, S. Vaidya, P. Thaplyal, M. Qamar, J. Ahmed and A. K. Ganguli, Langmuir, 2009, 25, 64696475. 51 M. Chen, Y. Wu, S. Zhou and L. Wu, J. Phys. Chem. B, 2008, 112, 65366541. 52 V. Uskokovic and M. Drofenik, Surf. Rev. Lett., 2005, 12, 239277. 53 A. Szczes , J. Cryst. Growth, 2009, 311, 11291135. 54 J. Eastoe and D. Sharpe, Langmuir, 1997, 13, 32893294. 55 M. G. Spirin, S. B. Bricklin and V. F. Razumov, J. Colloid Interface Sci., 2008, 326, 117120. 56 T. Charinpanitkul, A. Chanagul, J. Dutta, U. Rangsardthong and W. Tanthapanichakoon, Sci. Technol. Adv. Mater., 2005, 6, 266271. 57 K. E. Marchand, M. Tarret, J. P. Lechaire, L. Normand, S. Kasztelan and T. Cseri, Colloids Surf., A, 2003, 214, 239248. 58 M. L. Curri, A. Agostiano, L. Manna, M. D. Monica, M. Catalano, L. Chiavarone, V. Spagnolo and M. Lugara, J. Phys. Chem. B, 2000, 104, 83918397. 59 J. P. Cason, M. E. Miller, J. B. Thompson and C. B. Roberts, J. Phys. Chem. B, 2001, 105, 22972302.

Downloaded by University of Groningen on 19 October 2010 Published on 22 September 2009 on http://pubs.rsc.org | doi:10.1039/B814613F

This journal is

 c

The Royal Society of Chemistry 2010

Chem. Soc. Rev., 2010, 39, 474485 | 485

You might also like