ZeePedia

TEMPLATE BASED SYNTHESISSynthesis, Mechanism and Pathway

<< SOL-GEL TECHNIQUES:DEFINITIONS, GENERAL MECHANISM, INORGANIC ROUTE
MICROEMULSION TECHNIQUES:Significance of Packing Parameter >>
img
Chapter - 6
TEMPLATE BASED SYNTHESIS
B.Kuppan
INTRODUCTION
To make a jar, a piece of wood of the desired shape is first fabricated, and layer of
clay is applied to the wood. Heat treatment of the clay/wood composite at high
temperature generates a ceramic jar. During the heat treatment, clay is transformed to
a ceramic material and the wood is burned off leaving the empty space in the
resulting jar. When this process is scale down to nanometer regime, it is basically the
template synthesis process [1].
Fig. 6.1. Schematic representation of template synthesis (reproduced from ref. 1)
One class of templates are surfactants that are used to produce mesoporous materials.
It is true to say one of the most exciting discoveries in the field of material synthesis
over the last 15 years is the formation of mesoporous silicate and alumino silicate
molecular sieves with liquid crystal templates. This family materials generally called
as M41S family (MCM-41, MCM-48 and MCM-50) [2].
These M41S family
materials can be prepared by using cationic surfactants as the templates, and other
kind of mesoporous materials also can be prepared by using non-ionic surfactants as
the templates. These family materials are called as SBA-n family [3].
The above mentioned mesoporous molecular sieves also can be used as templates
to prepare ordered mesoporous carbon by using sucrose as the carbon precursors.
These mesoporous silicates, aluminosilicates and mesoporous carbon materials are
6.2
Template Based Synthesis
more important in the field of catalysis because of its high surface area, narrow pore
size distribution and large number of surface functional groups. The surfaces of these
materials can also be tuned depending on the applications.
Another kind of template based synthesis is to prepare freestanding, non-oriented
and oriented nanowires, nanorods or nanotubes. The fabrication of metal nanowires
had potential applications in the microelectronic industry and in particular, for
interconnection in electronic circuits. The procedure is based on metal displacement
reaction leading to the growth of metal nanowires into the pores. The galvanic
displacement reaction for the synthesis of core /sheet nanostructured materials has
been investigated in the literature [4].
Nanostructured materials have attracted much interest because of their unique
properties. In fact, due to their structure features and size effects, they show physical
properties that are different from bulk materials. Many methods have been developed
for the fabrication of nanowires. Among these methods template synthesis is
considered as quite useful, because it can be used for the preparation of different
types of nanostructures. The high order degree of porous structure of anodic alumina
membrane (AAM) (consisting in a close packed array of columnar hexagonal cells,
each containing a central cylindrical pore normal to the surface), makes it an ideal
template for the fabrication of nanostructured materials, suitable for applications in
optoelectronics, sensors, magnetic memories and electronic circuits [5].
Synthesis, Mechanism and Pathway
A large number of studies have been carried out to investigate the formation and
assembly of mesostructures on the basis of surfactant self-assembly. The initial liquid
crystal template mechanism first proposed by Mobil's scientists is essentially always
"true" because the pathways basically include almost all possibilities. Two main
pathways that is cooperative self assembly and "true" liquid ­crystal templating
process, seems to be effective in the synthesis of ordered mesostructures.
Surfactants for the formation of templates
The term surfactant is a blend of "surface acting agent" surfactants are usually
organic compounds that are amphophilic in nature. Amphophilic means they contain
img
Synthetic Strategies in Chemistry
6.3
both hydrophobic groups (tails) and hydrophilic groups (heads). Therefore they are
soluble in both organic solvents and water.
Generally a clear homogeneous solution of surfactants in water is required to get
ordered mesostructures. Frequently used surfactants can be classified into cationic,
anionic and non-ionic surfactants.
Table 6.1. Cationic surfactants
(Reproduced from ref.7)
img
6.4
Template Based Synthesis
Quaternary cationic surfactants, CnH2n+1N (CH3)3Br (n = 8-22), are generally efficient
for the synthesis of ordered mesoporous silicate materials. Commercially available
CTAB (cetyltrimethylammonium bromide) is often used. Gemini surfactants,
multiheadgroup surfactants, and recently reported cationic fluorinated surfactants can
also be used as templates to prepare various mesostructures. Frequently used cationic
quaternary ammonium surfactants are shown in the table below
First reports of mesoporous silica from Mobil's company, cationic surfactants
were used as structure directing agents (SDAs). Cationic surfactants have excellent
solubility, have high critical micelle temperature (CMT) values and can be widely
used in acidic and basic media. But they are toxic and expensive.
Anionic surfactants include carboxilates, sulphates, sulfonates, and phosphates;
recently a kind of lab-made anionic surfactant terminal carboxylic acids is used to
template the synthesis of mesoporous silicas with the assistance of aminosilanes or
quaternary
aminosilanes
such
as
3-aminoprpoyltrimethoxysilane
(APS)
N-
trimethoxylsilylpropyl-N, N, N-trimethylammonium chloride (TMAPS) as co-
structure directing agents (CSDAs).
Anionic surfactants
Fig. 6.2. Representation of anionic surfactants (reproduced from ref. 7)
img
Synthetic Strategies in Chemistry
6.5
Table 6.2. Non-ionic surfactants (reproduced from ref. 7)
Non-ionic surfactants are available in a wide variety of different chemical structures.
They are widely used in industry because of attractive characteristics like low price,
nontoxicity, and biodegradability. In addition the self assembling of non-ionic
surfactants products mesophase with different geometries and arrangements. They
become more and more popular and powerful in the syntheses of mesoporous solids.
The syntheses that largely promote the development of mesoporous materials are
simple and reproducible. A family of mesoporous silica materials has been prepared
with various mesoporous packing symmetries and well defined pore connectivity.
img
6.6
Template Based Synthesis
Cooperative Self-Assembly of Surfactant and Silica Source to Form
Mesostructure
This pathway is established on the basis of the interaction between the silicates
surfactants to from inorganic-organic mesostructure composites.
Cooperative Self-Assembly Mechanism
Fig. 6.3. schematic diagram of the co-operative self assembly of silicate-surfactant
mesophase. (reproduced from ref. 6)
A layer to hexagonal mechanism (folded sheet mechanism) was postulated by Kuroda
and co-workers, according to which the mesostructure is created from a layer
kanemite precursors. Silicate polyanions such as silicate oligomers interact with
positively charged groups in cationic surfactants driven by Coulomb forces. The
silicate species at the interface polymerized and cross-link and further change the
Synthetic Strategies in Chemistry
6.7
charge density of the inorganic layers. With the proceeding of the reaction, the
arrangement of the surfactants and the charge density between inorganic and organic
species influence each other. Hence the compositions of the inorganic-organic
hybrids differ to some degree. It is the matching of charge density at the
surfactants/inorganic species interfaces that govern the assembly process. The final
mesosphere is the ordered 3D arrangement with the lowest interface energy. The
transformation of the isotropic micellar solutions of CTAB into hexagonal or lamellar
phase when mixed with anionic silicate oligomers in highly alkaline solutions were
indeed detected through a combination of correlated solution state 2H,
13
29
C, and
Si
nuclear magnetic resonance (NMR) spectroscopy, small angle x-ray scattering
(SAXS), and polarized optical microscopic measurements. The mechanism in
different surfactant systems has been studied using NMR technique.
This cooperative formation mechanism in non-ionic surfactant system was
investigated by in-situ techniques. Goldfarb and co-workers investigated the
formation mechanism of mesoporous silica SBA-15, which are templated by triblock
copolymer P123 (EO20PO20EO20)
by using direct imaging and freeze-fracture
replication cryo-TEM techniques, in situ electron paramagnetic resonance (EPR)
spectroscopy, and electron spin-echo envelope modulation (ESEEM) experiments.
They found a continuous transformation from speriodial micelles into threadlike
micelles. Bundles were then formed with dimension that are similar to those found in
the final materials. The elongation of micelles in a consequence of the reduction of
polarity and water contact within the micelles due to the adsorption and
polymerization of silicate species. Before the hydrothermal treatment, the majority of
PEO chains insert into silicate frameworks, which generates micropores after removal
of templates. Moreover they found that the extent of the PEO chains located within
the silica micropores depended on the hydrothermal aging temperature and Si/P123
molar ratio. The formation dynamics of SBA-15 studied by Flodstrom et al. on the
basis of time-resolved in situ 1H NMR and TEM investigations. They observed four
stages during the cooperative assembly, which are the adsorption of silicates on
globular micelles, the association of globular micelles into floes, the precipitation of
floes, and the micelle-micelle coalescence. Khodakov et al. proposed a structure with
6.8
Template Based Synthesis
a hydrophobic PPO core and a PEO ­water ­silicate corona in the first stage. The
cylindrical micelles pack into the domains. At the same time, solvents are replaced by
condensed silicate species.
These mechanisms consider the interactions on the surfactants/inorganic species
interfaces. Monnier and Huo et al. gave a formula of the free energy in the whole
process.
G=
Ginter +
Gwall +
Gintra +
Gsol
In which
Ginter associated with interaction between the inorganic walls and
Gwall is the structural free energy for the inorganic frameworks,
surfactant micelles,
Gintra is the van der Waals force and conformational energy of the surfactant, and
Gsol is the chemical potential associated with the species in solution phase.
Gsol can be
For surfactant-templating assembly mesostructured silicates,
regarded as a constant in a given solution system. Therefore, the key factor is the
interaction between surfactant and inorganic species, such as the matching of charge
density. The more negative
Ginter is, the more easily the assembly process can be
proceeded.
Elaborate investigations on mesoporous materials have been focused on
understanding and utilizing the inorganic-organic interactions. Table 6.3 lists the
main synthesis routes and the corresponding surfactants and classical products.
Stucky and co-workers proposed four general synthetic routes, which are S+I-, S-
I+, S+X-I+, and S-X+I- (S+ = surfactant cations, S- = surfactant anions, I+ = inorganic
precursors cations, I- = inorganic precursors anions, X+ = cationic counter ions, and
X- = anionic counterions). To yield mesoporous materials, it is important to adjust the
chemistry of the surfactants headgroups, which can fit the requirement of the
inorganic components. Under basic conditions, silicate anions (I-) match with the
surfactant cations (S+) through Coulomb forces (S+I-). The assembly of the polyacid
anions and surfactant cations to "salt"-like mesostructures also belongs to S+I-
interaction. In contrast to this, one of the examples of S+I- interaction occurs between
cationic keggin ion (Al137+) and anionic surfactants like dodecyl benzenesulfonate
salt.
img
Synthetic Strategies in Chemistry
6.9
The organic-inorganic assembly of surfactants and inorganic precursors with the same
charge is also possible. However, counter ion is necessary. For example, in the
syntheses of mesoporous silicates by the S+X-I+ interaction, S+ and I- are cationic
surfactants and inorganic anionic precursors, and X- can be halogen ions (Cl-, Br-, and
I-), SO42-, NO3-, etc. In strongly acidic medium, the initial S+X-I+ interaction through
Coulomb forces or more exactly, double-layer hydrogen bonding interaction,
gradually transforms to the (IX)-S+ one. It was the first time that the mesoporous
silica was synthesized under a strongly acidic condition. Here anions affect the
structures, regularity, morphologies, thermal stability, and properties of mesoporous
silicas. The Hofmeister series of the anions are one of the possible reasons that
change the hydrolysis rates of the silicate precursors and the micellar structures.
Schematic illustration of the two types of interactions between APS (A) or TMAPS
(B) and anionic surfactant headgroups.
Fig. 6.4. interactions between APS (A) or TMAPS (B) and anionic surfactant
headgroups (reproduced from ref. 7)
Compared with those cationic surfactants, the repulsive interaction between anionic
surfactants and silicate species fails to organize ordered mesostructures. Concerning
the charge matching effect, Che et al. demonstrated a synthetic route to create a
family of mesoporous silica structures under basic conditions by employing anionic
img
6.10
Template Based Synthesis
surfactants as SDAs APS or TMAPS as CSDAs. This route can be described as an "S-
N+I-" pathway, where N+ are cationic amino groups of organoalkoxysilanes. Fig. 6.4
gives the schematic illustration of interactions between amino groups and anionic
surfactant head groups. The negatively charged headgroups of the anionic surfactants
interact with the positively charged ammonium sites of APS or TMAPS electrostatic
ally through neutralization. The most efficient surfactant is possibly terminal
carboxylic acid. The co-condensation of tetraethylorthosillane (TEOS) with APS or
TMAPS and assembly with surfactants occur to form the silica framework.
Table 6.3. Synthesis Routes to Mesoporous Materials with the Emphasis on Silicates
(reproduced from ref.7)
route
interactions
symbols
conditions
classical products
S+I-
S+, cationic surfactants
electrostatic
basic
MCM-41, MCM-48
I-, anionic silicate species
Coulomb
MCM-50,
SBA-6
force
SBA-2, SBA-8
FDU-2, FDU-11
FDU-
13
S-I+
S-, anionic surfactants
electrostatic
aqueous
mesoporous alumina
I+, Transition metal ions
Coulomb
Force
S+X-I+ electrostatic
S+, cationic surfactants
acidic
SBA-1, SBA-2
I+, cationic silicate species
SBA-3
Coulomb
force, double X-, Cl-, Br-, I-, SO4-, NO3-
layer H bond
S-N+I- electrostatic
S-, anionic surfactants
basic
ASM-n
N+, cationic amino groups
Coulomb
img
Synthetic Strategies in Chemistry
6.11
I-, anionic silicate species
Force
S-X+I- electrostatic S-, anionic phosphate
basic
W, Mo oxides
Coulomb
surfactants
force, double I-, transition metal ions,
layer H bond X+, Na+, K+, Cr3+, Ni2+
SoIo  H bond
So, nonionic surfactants
neutral
HMS, MSU,
disordered
(NoIo)
No, organic amines
Worm-like
Io, Silicate species, aluminate
mesoporous
species
silicates
SoH+X-I+ electrostatic So, nonionic surfactants
acidic
SBA-n (n=11,
Coulomb
12, 15, 16)
I+, silicate species
FDU-n (n=1, 5)
Force,
X-, Cl-, Br-, I-, SO42-, NO3-1
KIT-5 and KIT-6
double,
layer
H bond
No...I+
No, organic amines
coordination
acidic
Nb, Ta oxides
I+, transition metal
bond
S+-I-
S+, cationic surfactants
covalent
basic
mesoporous
I-, silicate species
silica
bond
Liquid Crystal Template Pathway
In this pathway true or semi-liquid crystal mesophases are involved in the surfactant
assembly to synthesize ordered mesoporous solids. Attard and co-workers
synthesized mesoporous silicas using high concentrations of non-ionic surfactants as
img
6.12
Template Based Synthesis
templates. The condensation of inorganic precursors is improved owing to the
confined growth around the surfactants and thus ceramic-like frameworks are formed.
After the condensation, the organic templates can be removed by calciation,
extraction, etc. The inorganic materials "cast" the mesostuructures, pore sizes, and
symmetries from the liquid crystal scaffolds. Direct templating of microemulsion
liquid-crystal mesophases were used to synthesis mesoporous silicas from butanol-
water-copolymer ­silica ternary system [6, 7].
Silica/surfactant = 1/0.27
calcination
As-synthesized
As-synthesized MCM-41
(hexagonal structure)
MCM-41
Silica/surfactant =
1/0.60
calcination
MCM-48
As-synthesized MCM-48
(cubic structure)
Fig. 6.4. possible mechanism pathways for the formation of M41S family by liquid
crystal template method (reproduced from ref. 2).
Evaporation Induced Self Assembling Techniques (EISA)
Evaporation induced self-assembly mechanism
The evaporation induced self assembling technique is one of the best techniques to
prepare the nanomaterials. The successful synthesis of mesoporous silica films, the
EISA method is engaged to prepare ordered mesoporous polymer and carbon
materials. The EISA method is strategy that avoids the cooperative assembling
img
Synthetic Strategies in Chemistry
6.13
process between the precursors and the surfactant template. Therefore, the cross-
linking and thermopolymerization process of the resols separate from the assembly.
Fig. 6.5. Scheme for the preparation of ordered mesoporous polymer resins and
carbon frameworks from the surfactant templating process of EISA. (reproduced from
ref. 8)
Compare to hydrothermal synthesis the EISA method is easier and can produce the
mesoporous resins and carbon in a wider synthetic range, including pH values,
surfactant and phenol/template ratio.
The choice of organic precursors is essential for the EISA of the organic ­organic
templating process. The polymerization of inorganic precursors should be low enough
to form a moldable inorganic-organic framework at the initial assembly stage of
inorganic species with organic surfactants. Highly ordered mesostructures can be
formed. The inorganic framework is rigid. Therefore, the mesophase can be
img
6.14
Template Based Synthesis
solidified, and the surfactant can be easily removed by calcinations or extracted with
ethanol.
The synthesis procedure includes five major steps (Fig. 6.5), which are the
preparation of resol precursors, the formation of ordered hybrid mesophases by
organic-organic self assembly during the solvent evaporation, thermopolymerization
of the resols around the templates to solidify the ordered mesophases, the removal of
the templates, and carbonization of the resin polymer frameworks to the homologous
carbons [8].
Template Synthesis of Metal Nanostructures
Fig. 6.6. Scheme of the arrangement used for the fabrication of copper nanowires
into the AAM template (reproduced from ref. 5)
Copper nanowires are prepared by using aluminium foil as active metal having a
thickness of 1.5 mm. In order to deposit copper inside the channels of anodic alumina
membrane (AAM), prior to cementation a thin conductive layer of Au was sputtered
on one side of the AAM using a conventional sputter coater to make this surface
electrically conductive. A portion of AAM was mounted onto the aluminum support
by means of a conductive paste and delimited by an insulating film. Since the
displacement deposition process needs an electrolytic contact of the active metal
img
Synthetic Strategies in Chemistry
6.15
surface, a small area of the aluminum support was exposed to the deposition solution.
A scheme of the arrangement used for the fabrication of copper nanowires is reported
in Fig. 6.5. This arrangement was placed horizontally in a beaker and covered with
25 ml of a 0.2 M copper sulphate and 0.1 M boric acid solution having pH 3. The
experiments were conducted at room temperature. The surface area of the AAM
exposed to the deposition solution was of the order of 1 cm2. A fresh solution was
used for each experiment. Experiments were carried out for different times of
deposition (from 7 h to 7 days) [5].
Preparation of Carbon Nanotubes
Fig. 6.7. Schematic representation of carbon nanotubes (reproduced from ref. 9)
Alumina membrane is used as template for the preparation of carbon nanotubes; here
polyphenyl acetylene is used as a carbon source for the preparation of carbon
nanotubes.
It
contains
only
carbon-hydrogen
bonds.
The
polyphenyl
acetylene/alumina composite was prepared by adding 10 ml of 5% w/w polyphenyl
acetylene in dichloromethane to the alumina membrane applying vacuum from the
bottom. The entire polymer solution penetrates inside the pores of the membrane was
dried in vacuum at 373 K for 10 min. the composite was then polished with fine
neutral alumina powder to remove the surface layers and ultrasonicated for 20 min to
img
6.16
Template Based Synthesis
remove the residual alumina powder used for polishing. The composite was
carbonized by heating in Ar atmosphere at 1173 K for 6 h at heating rate of 10 K/min.
This resulted in the deposition of carbon on the channel walls of the membrane. The
carbon/alumina composite was then placed in 48 % HF to free the nanotubes. The
nanotubes were washed with distilled water to remove HF [9].
Preparation of Porous Solids by Template Method
Ordered mesoporous carbon has been prepared by using mesoporous silica as the
template, and this mesoporous silica can be prepared by using CTAB, P123,
surfactants as the templates.
Mesoporous silica SBA-15 is prepared by the following method, P123 non-ionic
surfactant was dissolved in 2 M HCl solution, and it was stirred for 1 h then TEOS
was added as the silica source, this mixture was stirred until TEOS were completely
dissolved. The mixture was placed in oven at 373 K for 48 h. The product was filtered
and washed with distilled water and dried at 333 K for 6 h and then SBA-15 was
calcined at 823 K for 6 h in air atmosphere.
Fig. 6.8. XRD patterns of SBA-15 and CMK-3 (reproduced from ref. 10)
img
Synthetic Strategies in Chemistry
6.17
The calcined SBA-15 was impregnated with aqueous solution sucrose containing
sulfuric acid, this mixture was heated to 373 k to 433 K for 24 h. then the
carbonization was completed by pyrolysis with heating to typically 1173 K under N2
atmosphere for 6 h at the rate of 5 K/min. the carbon-silica composite obtained was
washed with 1 M NaOH solution or 5 Wt % hydrofluoric acid at room temperature, to
remove the silica template. The template free carbon product thus obtained was
filtered, washed with ethanol and dried at 393 K.
In Fig. 6.9 the low angle XRD conforms the mesoporous structure and to support this
mesoporous structure the transition electron microscopy images shows the presence
of ordered mesoporous structure [10].
Fig. 6.9 TEM and selected area electron diffraction (SAED) images of CMK-3
(reproduced from ref.10)
SUMMARY
Strategy aimed at the controllable synthesis has been focused on the control of micro,
meso and macroscale, including synthetic methods, architecture concepts, and
fundamental principles that govern the rational design and synthesis. In this chapter,
synthesis mechanisms and the corresponding pathways are first demonstrated for the
synthesis of mesoporous silicates from the surfactant-templating approach. Virtually
all mesoporous silicates begin with an understanding of the interaction between
organic surfactants and inorganic species, as well as among themselves.
6.18
Template Based Synthesis
Soft templating approach is one of the most general strategies now available for
crating nanostructures. The assembly of surfactants and silicates species is normally
carried in solutions or the interface to allow the required driving force for the
formation of nanostructures. Relying on sol-gel, solution and surface chemistry, there
is great potential to explore novel strategies for mesostructures, especially a strategy
that can utilize interfacial tension. Items that attract attention also include the control
of weak interaction such as the hydrophobic interaction between the assembling
components. In view of the fact that surfactant self-assembly can occur with
components larger than 1 nm continuous to be challenge and an interest in condensed
matter science.
REFERENCES
1. J. Lee, J. Kim, T. Hyeon, Adv. Mater., 18 (2006) 2073.
2. C. T. Kresge, M. E. Leonowicz, W. J. Roth, J. C. Vartuli, J. S. Beck,
Nature, 359 (1992) 710.
3. D. Zhao, J. Feng, Q. Huo, N. Elosh, G. H. Fredirckson, B. F. Chemlka,
G. D. Stucky, Science, 279 (1998) 548.
4. G. Cao, D. Liu, Adv. Colloid. Interface. Sci. 136 (2008) 45
5. R. Inguanta, S. Piazza, C. Sunseri, Eleectrochem. Commun., 10 (2008) 506
6. A. Corma, Chem. Rev. 97 (1997) 2373.
7. Y. Wan, D. Zhao, Chem. Rev. 107 (2007) 2821.
8. Y. Meng, D. Gu, F. Zhang, Y. Shi, L. Cheng, D. Feng, Z. Wu, Z. Chen,
Y. Wan, A. Stein, D. Zhao, Chem. Mater., 18 (2006) 4447.
9. M. Sankaran, Ph.D Thesis, Indian Institute of Technoloyg Madras, 2007.
10. S. Jun, S. Joo, R. Ryoo, M. Kruk, M. Jaroniec, Z. Liu, T. Ohsuna,
O. Terasaki, J. Am.Chem. Soc., 122 (2000) 712.
Chapter - 7
MICROEMULSION TECHNIQUES
Ch. Venkateswara Rao
The main focus of this chapter is to give brief introduction to the microemulsion systems,
formation of microemulsions, how those microemulsions can be used for the preparation of
various types of nanoparticles and the dominant factors that influence the nanoparticles
preparation in the microemulsion.
INTRODUCTION
In 1943, Hoar and Schulman first reported that oil can be dissolved in bulk water or water in
bulk oil with the aid of surfactant to produce a clear homogeneous solution. The oil phases are
simple long-chain hydrocarbons and the surfactants are long-chain organic molecules with a
hydrophilic head (usually an ionic sulfate or quarternary amine) and lipophilic tail. The generally
used oil phases are cyclohexane, decane, heptane and surfactants are cetyltrimethylammonium
bromide (CTAB), sodium bis(2-ethylhexyl) sulfosuccinate (AOT).
The clear homogeneous
solution is called as microemulsion. According to the nature of the bulk solvent used in the
microemulsion, they are designated as water-in-oil or oil-in-water microemulsions. The
microemulsion is primarily distinguished from the emulsion not by being composed of smaller
droplets but by being subjected to a restrictive condition that it is thermodynamically stable.
MICROEMULSION FORMATION AND NANOPARTICLES PREPARATION
In general, water and oil are not miscible. But the amphiphilic nature of the surfactants such as
CTAB makes them miscible. The surfactant molecules form a monolayer at the interface
between the oil and water, with the hydrophobic tails of the surfactant molecules dissolved in the
oil phase and the hydrophilic head groups in the aqueous phase. It leads to the formation of
microemulsion. So the microemulsion is defined as a system of water, oil and surfactant. This
system is an optically isotropic and thermodynamically stable. At macroscopic scale, a
microemulsion looks like a homogeneous solution but at molecular scale, it appears to be
heterogeneous. Even though it is optically isotropic; it cannot be properly described as a solution.
The internal structure of the microemulsion at a given temperature is determined by the ratio of
its constituents. The structure consists either of nanospherical monosized droplets or a
img
7.2
Microemulsion Techniques
bicontinuous phase. The different structures of a microemulsion at a given concentration of
surfactant are shown in Fig 1. It indicates that
(i)
At high concentration of water, the internal structure of the microemulsion consists of
small oil droplets in a continuous water phase (micelles), known as o/w
microemulsion.
(ii)
With increased oil concentration, a bicontinuous phase without any defined shape is
formed.
(iii)
At high oil concentration, the bicontinuous phase is transformed into a structure of
small water droplets in a continuous oil phase (reverse micelles), known as a w/o
microemulsion.
Depending on the type of surfactant used to form microemulsion, size of the different
droplets will be varied from 5-100 nm. It is also evident that the microemulsion system is
sensitive to temperature. It can be seen in Fig. 7.1 that the increase in temperature will
destroy the oil droplets while the decrease in temperature will destroy the water droplets.
Fig. 7.1 The microscopic structure of a microemulsion at a given concentration of surfactant as
function of temperature and water concentration (reproduced from ref. 20).
Significance of Packing Parameter
The shape of micellar aggregates and the formation of microemulsion can be understood from
the packing parameter of surfactant molecule used for the microemulsion formation. Packing
The Evolving Synthetic Strategies in Chemistry
7.3
parameter is defined as v/a.l, where v is the volume of hydrocarbon of the surfactant, a is the
polar head area and l is the fully extended chain length of the surfactant. The packing parameter
value can be 1, <1 and >1. When the packing parameter value (or ratio v/a.l) is greater than 1, the
aggregate curvature will be toward the water. This corresponds to a situation where the oil is
penetrating the surfactant tails and/or the electrostatic repulsion between the charged head group
is low. When the packing parameter value (or ratio v/a.l) is less than 1, it corresponds to a
situation where the electrostatic repulsion is larger and/or the oil is not penetrating the surfactant
tails. It implies that (i) when oil is solubilized in hydrophilic micelles, one can observe the
formation of o/w microemulsions for v/a.l<1; (ii) when water is solubilized in hydrophobic
micelles, one can observe the formation of w/o microemulsions for v/a.l>1.; and (iii) When
v/a.l≈1, lamellar phases or bicontinuous microemulsions are observed.
Based on the experimental results, it is observed that the micelles will be in spherical shape
when the packing parameter is less than 1/3. The packing parameters for cylinders and planar
bilayers are 0.5 and 1, respectively. In the case of reverse micellar structures, the packing
parameter is greater than 2 for cylinders as well as spherical micelles. It was observed that
reverse micelles will be in cylinder shape up to v/a.l≤2 and spherical shape when v/a.l>3.
Oil-in-water (o/w) microemulsions are monodisperse. Water-in-oil (w/o) microemulsion
solutions are mostly transparent, isotropic liquid media with nanosized water droplets that are
dispersed in the continuous oil phase and stabilized by surfactant molecules at the water/oil
interface. These surfactant-covered water pools offer a unique microenvironment for the
formation of nanoparticles. They not only act as micro-/nano-reactors for processing reactions
but also exhibit the process aggregation of particles because the surfactants could adsorb on the
particle surface when the particle size approaches to that of the water pool. As a result, the
particles obtained in such a medium are generally uniform in size and shape (i.e., monodisperse).
APPLICATIONS OF MICROEMULSIONS
1. Synthesis of Metal Nanoparticles
Precipitation of metal particles in the Water-in-oil (w/o) microemulsion solutions (reverse
micellar system) has been found to be the simple methodology for the preparation of
nanoparticles.
img
7.4
Microemulsion Techniques
The method of nanoparticle preparation consists in mixing of two microemulsions carrying the
appropriate reactants (generally metal precursor and reducing agent) in order to obtain the
desired particles. It is represented in Scheme 7.1.
Scheme 7.1. Proposed mechanism for the formation of metal particles by the microemulsion
approach (reproduced from ref. 24)
At the beginning of reaction, the attractive van der Waals force and the repulsive osmotic force
between reverse micelles lead to collisions of micelles. It results in the interchange of the
reactants (in general, metal ions and reducing species) solubilized in two different reverse
micelles respectively. As a result, the initial monomeric metal nuclei begin to form and grow.
When the exchange of reactants is fast in the water droplets, metal ions reduce and the metal
nuclei grow quickly. It remains unchanged when the particle size reaches a certain size. It
corresponds to the thermodynamically stabilized species in the presence of microemulsion. Due
The Evolving Synthetic Strategies in Chemistry
7.5
to the possible aggregation, the final size of silver particles is generally larger than that of the
water cores. Once the particles attain the final size, the surfactant molecules are attached to the
surface of particles and stabilize and protect them against further growth. The dynamic exchange
of reactants such as metallic salts and reducing agents between droplets via the continuous oil
phase is strongly depressed due to the restricted solubility of metal salts in the oil phase. This is a
reason why the attractive interactions (percolation) between droplets play a dominant role in the
particle nucleation and growth in the water-in-oil (w/o) microemulsion reaction medium.
2. Synthesis of Metal Alloys
Various metal alloys can be prepared in the similar way as described in scheme 1. It consists of
mixing the reverse micelle containing two or three metal precursors and another reverse micelle
contain reducing species. Example: formation of Pt-Ru (1:1) alloy. In order to prepare the alloy
nanoparticles, both Pt4+ and Ru3+ will be taken in 1:1 atomic ratio to form reverse microemulsion
I. Reverse microemulsion II will be prepared by using NaBH4 reducing agent. When both the
reverse microemulsions are mixed together, reduction of Pt4+ and Ru3+ takes place by BH4- ions
and results in the formation of Pt-Ru (1:1) nanoparticles (Xiong and Manthiram, 2005). In the
similar way, Pt-Fe nanoparticles in different ratios (1:1, 1:2 and 1:3) can be prepared (Carpenter
et al., 2000). In the same manner, various metal chalcogenides also can be prepared by taking
two micelle solutions containing the desired ions prepared separately and then rapidly mixed.
In conventional methods of preparation, the reaction temperature should be high to promote
alloy formation. As a result, the formed particles will be large in size. Moreover, the shape of the
particle cannot be controlled. Since the reaction takes place inside the micellar cores,
(i) the enormous amount of heat that is generated within the micellar cores during the reduction
process is enough to form alloy of desired composition.
and (ii) both size, shape and composition can be controlled.
3. Synthesis of Metal Oxides
The synthesis of oxides from reverse microemulsion relies on the co-precipitation of one or more
metal ions. It is almost similar in many respects to the precipitation of oxides from aqueous
solutions. But the precipitation occurs within the micellar cores so that particle size as well as
shape can be controlled. In a typical process, precipitation of hydroxides is induced by addition
of a reverse micelle solution containing dilute NH4OH to a reverse micelle solution containing
aqueous metal ions at the micellar cores. Alternatively, dilute NH4OH can simply be added
img
7.6
Microemulsion Techniques
directly to a micelle solution of the metal ions. The precipitation of the metal hydroxides is
typically followed by centrifugation and heating in the presence of oxygen to remove water
and/or improve crystallinity. In general, it is represented as follows:
A2+ + 2B2+ + OH- (excess) → AB2O4 + xH2O↑
Where A and B are metals.
In this way, simple metal oxides and multicomponent metal oxides can be prepared. Some of the
examples of nanomaterials synthesized in w/o microemulsions are given in Table 7.1.
Table 7.1. Nanomaterials formed in w/o microemulsions
Nanoparticle system
Example
Reference
Pt
Metals/Metal alloys
Rojas et al., 2005
Pd
Boutonnet et al., 1982
Ag
Taleb et al., 1997
Au
Herrera et al., 2005
Pt-Ru
Xiong and Manthiram, 2005
Fe-Pt
Carpenter et al., 2002
Pd-Co-Au
Raghuveer et al., 2006
ZnS
Metal chalcogenides
Khiew et al., 2005
PbS
Eastoe et al., 1995
RuSe
Venkateswara Rao and Viswanathan, 2007
Pb2Ru2O7
Metal oxides
Raghuveer et al., 2002
CeO2
Bumajdad et al., 2004
CoFe2O4
Liu et al., 2003
LiNi0.8Co0.2O2
Lu et al., 2000
Core-shell nanoparticles
Fe3O4/SiO2
Tago et al., 2002
Fe/Au
Zhou et al., 2000
The Evolving Synthetic Strategies in Chemistry
7.7
4. Synthesis of Core-Shell Nanoparticles
Some of the metallic nanoparticles like Fe, Co, Ni are susceptible to rapid oxidation. This
problem can be largely circumvented by coating the nanoparticles with gold or other inert
metals. The technique for applying gold coatings on metal nanoparticles is reasonably
straightforward and simply adds an additional step to the reverse micelle synthesis. In this
process, a water-soluble gold salt (HAuCl4) is dissolved and dispersed in a separate reverse
micelle solution that is then added to the metal-containing reverse micelle solution that has
already been reduced with an excess of BH4-. The aqueous AuCl4- ions encapsulate the metal
particles and are subsequently reduced by forming a metallic gold shell around the metal
particles.
8AuCl4- + 3 BH4- + 9H2O → 8Au + 3B(OH)3 + 21H+ + 32Cl-
The preparation of core-shell type structures is not limited to metals as the core or shell
materials. Combinations of precipitation, reduction and hydrolysis reactions can be performed
sequentially to produce oxides coated with metals, oxides coated with oxides, and so forth.
EFFECTS OF THE PARAMETERS ON THE FORMATION OF NANOPARTICLES IN
MICROEMULSION
The main parameters that influence the size of nanoparticle are molar ratio, W =
[water]/[surfactant], type of solvent employed, surfactant or co-surfactants used and
concentration of reagents. Recent investigations suggest that the particle shape also can be
affected by the influence of micellar template, added anions and molecular adsorption. But a
general method for controlling nanocrystal shapes through soft chemistry has not yet been found.
The major factors that influence the formation of nanoparticles are explained with examples
below.
Water-to-Surfactant Ratio, W
The size of the metallic particle will depend on the size of the droplets in the microemulsion. The
droplet size will be influenced by the water-to-surfactant ratio, W.
In general, the volume of water in microemulsions is proportional to the cubic radius of the
water core. And also the amount of surfactant which is present as a film around the water cores is
proportional to the surface area of the micro/nanodroplets. So the molar ratio of water to
surfactant (W) has a linear relationship with the radii of the water cores (Rw). It has been
observed in AOT microemulsion system that the water-pool radius increases with the water
img
7.8
Microemulsion Techniques
content. An increase of molar ratio, W at constant concentration of surfactant will increase the
average diameter of the droplets. Consequently, the obtained nanoparticles will be large in size.
Table 7.2. Characteristic of silver particles obtained in w/o microemulsion (data taken from ref.
23)
W
Reducing agent
Conc. of reducing
Particle size (nm)
agent (M)
2.5 x 10-4
-
NaBH4
2
2.5 x 10-4
2.7
NaBH4
5
1.0 x 10-5
-
NaBH4
7.5
1.0 x 10-4
4.5
7.5
NaBH4
2.5 x 10-4
5.5
NaBH4
7.5
1.0 x 10-4
6.0
NaBH4
15
2.5 x 10-4
7.0
NaBH4
15
For example, by dynamic light scattering technique, it has been confirmed that the water core
radius, Rw, is coincident with the equation Rw(nm) = 0.18W + 1.5 in a wide range of alkanes. For
example in the AOT-water-isooctane microemulsion system, it was found that the linear
relationship between Rw and W was Rw (Å) = 1.5W. Thus the linear relationship may be different
in various systems. Generally, low water content is favorable to form smaller microemulsion
droplets (reverse micelles) which yield fine nanoparticles with a narrow size distribution. On the
contrary, high water content is favorable to form bigger microemulsion droplets which are easy
to fabricate larger particles. The reason for the behaviour is explained as follows: Actually, at
low water content, the water solubilized in the polar core is bound by the surfactant molecules,
which increases the boundary strength and decreases the intermicellar exchange rate. As a result,
a decrease in water content induces formation of monodisperse nanoparticles with small particle
size. However, the bound water would turn into bulk water with an increase in water content,
which is benefit for the water pools to exchange their contents by collisions and makes the
chemical reaction or co-precipitation between compounds solubilized in two different reverse
micelles complete more quickly. Since the nature of bulk water is drastically different from that
of bound water, reactants would be rapidly transferred from one water core to another, and thus
img
The Evolving Synthetic Strategies in Chemistry
7.9
the resultant particle size is relatively big and the size distribution becomes relatively wide. The
variation of silver and Fe nanoparticles size with W is given in Tables 7.2 and 7.3 respectively.
Table 7.3. Characteristic of iron particles obtained in w/o microemulsion (Data taken from ref.
24)
[FeCl2]x104
[NaBH4]x103
Particle size
S. No
[AOT]
W
(mol/dm3)
(mol/dm3)
(nm)
3.9
3.5
7.5
22
0.1
1
4.1
0.68
1.6
22
0.1
2
2.5
0.35
0.93
10
0.05
3
1.9
24
48
5
0.025
4
2.5
24
48
9
0.025
5
3.3
24
48
13
0.025
6
4.3
24
48
18
0.025
7
4.7
24
48
22
0.025
8
5.4
24
48
26
0.025
9
5.8
24
48
31
0.025
10
Solvent Effect
Particle size is affected by solvent type. This was shown initially by Pileni (2003) in a study on
silver nanoparticles, in which larger particles were seen (by TEM) to be formed in isooctane than
in cyclohexane. This is probably due to the significant difference in intermicellar exchange rate
constant between the two solvents - a factor of 10.
The change in growth rate has been explained in the following way:
(i)
Smaller and less bulky solvent molecules with lower molecular volumes such as
cyclohexane, can penetrate between surfactant tails and increase the surfactant
curvature and rigidity. The increased rigidity at the interface may lead to a slower
growth rate.