Вы находитесь на странице: 1из 14

1.

1 Introduction
Nanoparticles play a vital role in high performance materials in high
technology
industries. The studies of nanoparticles started in the early 1980's and have
now become
one of the hottest worldwide research fields (Pui and Chen, 1997).
There are four main processing approaches for the preparation of
nanoparticles by
chemical method (Riman, 1993): (1) chemistry in liquid phase including direct
strike
(Murata, et al., 1976), nonsolvent addition (Mulder, 1970), solvent removal
(Cheng, et
al., 1986), gel drying (sol-gel) (Perthuis, And Colomban, 1984) and
precipitation from
homogeneous solution (Gordon, et al., 1959); (2) chemistry between
heterogeneous
phase including hydrothermal synthesis (Adair, et al., 1987), molten salt
synthesis
(Arendt, et al., 1979), pyrolysis (Wada, et al., 1987) and spark erosion
(Berkowitz, et al.,
1987); (3) chemistry in a droplet including emulsions (Woodhead, et al.,
1980), micelles
(Gobe, et al., 1983) or microemulsions (Kandori, et al., 1988) and aerosols
(Balboa, et
al., 1987); (4) chemistry in the vapor phase including heating method
(Mazdiyasni, et al.,
1965), vapor precursors (Iwama, et al., 1982), liquid precursors (Kagawa, et
al., 1983)
and solid precursors (Watanabe, et al., 1986). The most attractive methods
are those
which synthesize in the liquid medium, including methods of precipitation,
reduction,
dehydration, solvent evaporation, reversed micelle technology and
microemulsion
polymerization, etc. In this chapter, we will focus on the nanoparticles made
from both
W/O microemulsion (reversed micelles) and O/W microemulsion procedures.
Hence it is necessary to introduce the definition of micelles and
microemulsions before
dealing with the principles and practices of forming nanoparticles from
micelles and
microemulsions. Micelles are aggregates of surfactants in a liquid medium
which are
formed when the surfactant concentration exceeds the critical micelle
concentration
(CMC) (McBain and Salmon, 1920). It must be mentioned that this definition
is only for
normal micelles; for the case of reversed micelles it is not necessary to have
a CMC. In
the normal micelle the surfactant is orientated in such a way that the
hydrophobic
hydrocarbon chains are towards the interior of the micelle, leaving the
hydrophilic
groups in contact with the aqueous medium. Above the CMC, the physical
state of the
surfactant molecules dissolved in water changes dramatically, and additional
surfactant
exists as aggregates or micelles. Thus, the bulk properties of the surfactant,
such as
osmotic pressure, turbidity, solubilization, surface tension, conductivity and
selfdiffusion,
change around the critical micelle concentration (Fig. 1.1).
Figure 1.1 Changes in concentration dependence of a wide range of
physicochemical
quantities around the critical micelle concentration (After Lindman,
1980).
2
If the micelles are formed in non-aqueous medium, the aggregates are called
reversed
micelles, as in this case the hydrophilic head groups are now towards the
core of the
micelle while leaving the hydrophobic groups outside of the micelles. The
driving force
for formation of reversed micells is the dipole-dipole interactions of the
surfactant. The
number of aggregates is usually small and not sensitive to the surfactant
concentration
and thus there is no obvious CMC (Zhao, 1991; Gutmann and Kertes, 1973;
Kertes and
Gutmann, 1976). In both cases (micelles and reversed micelles), only a small
amount of
solubilized hydrophobic (usually oil) or hydrophilic (usually water) material
exists in the
micelles (Fig. 1.2). However, the solubilization can be enhanced if the
concentration of
surfactant is increased further. As the inside pool of water or oil is enlarged or
swollen,
the droplet size increases up to a dimension much larger than the monolayer
thickness of
the surfactants. In this case, we call them microemulsions or swollen
micelles. What we
now describe as the preparation of nanoparticles from the reversed micelles
may be
better described as preparation from swollen reversed micelles or water-in-oil
microemulsions.
Figure 1.2 The structure of micelles and microemulsions (O/W and W/O) (After
Overbeek et al., 1983).
As the surfactant concentration increases further, micelles can be deformed
and can
change their shapes to rodlike micelles, hexagonal micelles and lamellar
micelles or
3
liquid crystals (Fig. 1.3). It is these changes that make it possible to prepare
different
shapes of nanoparticles from micelle synthesis microreactors.
1.2 Formation Mechanisms of Micelles and Microemulsions
1.2.1 Simple Geometric Factors
The structures of micelles can be simply determined by the geometric factors
of the
surfactant at the interface, including head group area a0, the alkyl chain
volume v and the
maximum length lc (to which the alkyl chain can extend). According to
Israelachvili
(Israelachvili, et al., 1976), the packing considerations govern the geometry
of
aggregation into micelles, vesicles and liposomes.
Figure 1.3 A schematic phase diagram of surfactant-oil-water systems showing
a variety of self-assembled structures (After Liu, J., et al., 1996).
These obey the following rules:
1. Spherical micelles require v/a0lc < 1/3,
2. Non-spherical micelles require 1/3 < v/a0lc < 1/2,
3. Vesicles or bilayers require 1/2 < v/a0lc < 1, and
4. Inverted micelles require 1 < v/a0lc.
In each case, the limits for the packing parameter v/a0lc can be evaluated from
simple
geometry (Fig. 1.4) (Israelachvili, 1985). However, the change of environment
will
affect these parameters, and thus dictate the molecular packing at the
interface.
4
Figure 1.4 The relationship between aggregate type and geometry on the
packing requirements of surfactant head group and chains (Israelachvili,
1985).
5
1.2.1.1 Spherical Micelles
Spherical micelles are usually formed by anionic surfactants with or without
cosurfactants.
For an O/W micelle, this can be done by adjusting the repulsion (double
layers) between adjacent head groups, resulting in large values for a0. In this
case, the
micelle radius is approximately equal to the maximum stretched out length of
the
surfactant molecule and therefore the aggregates are very small. Bellare et
al. (1988),
using small-angle neutron scattering (SANS), have successfully visualized a
spherical
micelle of radius (3.0 ± 0.3) nm for a cryo-TEM image of a 10 mmol • dm-3
solution of
ditetradecyl-dimethyl-ammonium acetate.
1.2.1.2 Cylindrical Micelles
It is a quite common phenomenon that micelles grow as the preferred surface
curvature
decreases. Any change that reduces the effective head group area will lead
to the growth
of micelles. There are basically three ways to form cylindrical micelles: (1)
addition of a
co-surfactant with a very compact head group, i.e. n-alkanol for which the–OH
group is
small in comparison with a charged sulfate group, (2) changing the
counterion, i.e.,
changing Na+ to Mg2+ will significantly reduce the electric double layer
thickness, and
hence reduce the effective volume (size) of the head groups, (3) changing
the
hydrophilicity of non-ionic head groups by electrolyte addition or temperature
change;
i.e., for micelles formed by surfactants with poly(oxyethylene)(PEO) head
groups, the
head groups are sensitive to changes of solvency (Tadros, 1987).
1.2.2 The Critical Micelle Concentration (CMC) for Surfactants
The CMC of a surfactant system depends on the minimum value of the
interaction free
energy per molecule μN
0. The minimum arises from the hydrophilicity of the head group,
tending to increase the area per molecule, while the hydrophobicity of the
alkyl tail
tends to cause a decrease due to the hydrophobic bonding. From this
concept, one is able
to predict how various structural features of surfactant molecules will affect
their CMC
values.
Table 1.1 Typical CMC values for ionic surfactants at 25 °C
Surfactant CMC/mmol • dm-3
C12H25SO4Na 8.1
C12H25SO4Li 8.9
C12H25SO3Na 10
C12H25CO3K 12.5
C12H25NH3Cl 14.7
6
C12H25NC2H5Cl 15
C12H25N(CH3)3Br 16
C12H25N(CH3)3Cl 17
For these ionic surfactants, there is little difference between anionic and
cationic head
groups, since both have comparatively high CMC values, provided that the
counterion is
monovalent. Usually, the CMC values for these systems are 1–20 mmol • dm-3
(Table
1.1). However, to change the counterion to a multivalent one tends to
decrease the CMC
considerably.
For non-ionic surfactants, such as CxEy type, where x is the carbon number in
the range
of 8–18, and y is the ethylene oxide group in the range of 3–20, the CMC
value is
extremely low, i.e., 0.04–3 mmol • dm-3, depending on the structure of the
molecules
(Table 1.2).
Table 1.2 Typical CMC values for non-ionic surfactants at
25 °C
Surfactant CMC/mmol · dm-3
C12H25(OCH2CH2)4OH 0.046
C12H25(OCH2CH2)6OH 0.087
C12H25(OCH2CH2)8OH 0.109
C12H25(CH3)NO 2.1
1.2.3 Solubilization and Formation of Microemulsions
1.2.3.1 Solubilization
The term solubilization in this chapter refers to the dissolution of hydrophobic
(hydrophilic) materials into water (or oil) to an extent greatly exceeding their
normal
solubilities in water (oil). The interior of a micelle provides a hydrophobic
(hydrophilic)
environment in which non-polar (or polar) compounds can be accommodated.
As a
result, the solubility of hydrophobic (or hydrophilic) material increases
dramatically
with increasing surfactant concentration when it reaches the CMC as shown
in Fig. 1.1.
The solubility behavior of surfactants is anomalous as the temperature is
increased to a
value at which there is a sudden increase in solubility and the material then
becomes
very highly soluble (Krafft, 1899). This is illustrated in Fig. 1.5.
7
Figure 1.5 Schematic representation of solubility versus temperature showing
location of the Krafft point (After Shinoda, 1974).
The process of solubilization has many applications in industrial preparations,
for
example, in solubilization of insoluble drugs for intravenous injection. The
process of
solubilization by micellar systems is also important in detergency, whereby
fats and oils
are removed by incorporation into the hydrocarbon core of the micelle. There
are four
general possible ways for the incorporation of the solubilization: (1) in the
hydrocarbon
core of the micelle; (2) orientation in the micelle which could be deep or
shallow; (3) in
the hydrophilic portion of the surfactant (e.g., ethylene oxide of non-ionic
surfactants);
and (4) adsorption on the surface of the micelle (Fig. 1.6).
Figure 1.6 Schematic representation of four ways of solubilization of micelles.
8
1.2.3.2 Microemulsions
The microemulsion systems were first reported by Hoar and Schulman
(1943), who
described transparent or translucent systems, formed spontaneously when oil
and water
were mixed with a relatively large amount of an ionic surfactant combined
with a
cosurfactant, e.g., a medium size alcohol. Later, in 1959, Schulman and co-
workers
(Schulman, et al., 1959) introduced the concept of microemulsions as
transparent or
translucent systems with a spherical or cylindrical size range of 8–100 nm.
This is the
right size for preparing spherical and rod-like nanometer particles.
The solubilization theories of microemulsions have been proposed by Shinoda
(Shinoda,
1974), who considered microemulsions as solubilized systems extended from
the threecomponent
phase diagrams of water-surfactant and co-surfactant (Fig. 1.7). It is clear
that in the phase diagrams there are two isotropic regions: one in the top
corner, the so
called L2 phase or inverse micelles, and one in the left corner, i.e., L1 phase or
normal
micelles. The L2 phase is capable of dissolving a large amount of water,
thereby forming
a W/O microemulsion. Similarly, the L1 phase can solubilize oil to form an O/W
microemulsion. Thus, O/W microemulsions can be considered as an extension
of the L2
phase, whereas W/O microemulsions can be considered as an extension of
the L1 phase.
Figure 1.7 Schematic representation of a tree-component phase diagram for
water-surfactant and cosurfactant (After Overbeek et al., 1983).
The advantages of microemulsions in many industrial processes are distinct:
from their
spontaneous formation, thermodynamic stability to lack of aging.
Applications are based
on the low interfacial tension (as in tertiary oil recovery), the possibility of
preparing
both hydrophilic and hydrophobic nearly homogeneous nanoparticles, the
small droplet
size produced and their isodisperse nature.
9
1.3 Synthesis of Nanoparticles from W/O Microemulsions (Reversed Micelles)
O/W microemulsions (reversed micelles) can be formed by ionic surfactants
with double
long alkyl chains alone, such as, AOT (Aerosol OT) by or a mixture of ionic
and
nonionic surfactants with a short oxyethylene chain dissolved in organic
solvents.
Reversed micelles are usually thermodynamically stable mixtures of four
components:
surfactant, co-surfactant, organic solvent and water. AOT, SDS (sodium
dodecyl sulfate),
CTAB (cetyltrimethy lammonium bromide) and Triton-X are the usual
surfactants. Cosurfactants
are often aliphatic alcohols with a chain length of C6–C8. Organic solvents
used for reversed micelle formation are usually alkane or cycloalkane with 6
to 8
carbons.
Reversed micelles can solubilize relatively large amounts of water. It is this
water pool
that makes the reversed micelles particularly favorable for the synthesis of
nanoparticles
because the water pool is in the range of nanometer size which can be
controlled by
adjusting the water content. Solubilization of water in the reverse micelles
can be
expressed by w, the ratio of water to surfactant concentrations (w = [H2O]/
[surfactant]).
w is an important parameter in determining the size of the reversed micelles
and the
structure of water. For a typical spherical AOT reversed micelle, there is a
linear
relationship between the diameter of the water pool (D) and w. D = 0.3 w
when w is
larger than 15 (Pileni, et al., 1985). In addition, w is related to the structure of
water. For
an AOT reverse micelle, when w increases, the structure of the water changes
from
bound water to free water.
Due to the controllable water pool, reversed micelles are particularly
favorable for the
preparation of monodisperse nanoparticles with various particle sizes. The
nanoparticles
can be fabricated using the reversed micelles having the following two
features: (1) the
nanoparticles are harder to aggregate because the surface of the
nanoparticles is covered
with surfactants; (2) the surface of the particles can be modified further.
Preparation of nanoparticles using reverse micelles can be dated back to the
pioneer
work of Boutonnet et al. (Boutonnet, et al., 1982). In 1982 they first
synthesized
monodispersed Pt, Rh, Pd, Ir nanoparticles with diameters of 3–6 nm. After
that, many
nanoparticles were synthesized and the method of preparing nanoparticles
using reverse
micelles became a world wide interest in nanoscience and nanotechnology. In
the
following sections we will review the synthesis of various nanoparticles using
the
technique of reversed micelles.
The general method to synthesizing nanoparticles using reverse micelles is
schematically illustrated in Fig. 1.8. This can be divided into three cases. The
first one is
the mixing of two reverse micelles. Due to the coalescence of the reverse
micelles,
exchange of the materials in the water droplets occurs, which causes a
reaction between
the cores. Since the diameter of the water droplet is constant, nuclei in the
different
water cores can not exchange with each other. As a result, nanoparticles are
formed in
the reversed micelles. The second case is that one reactant (A) is solubilized
in the
reversed micelles while another reactant (B) is dissolved in water. After
mixing the two
10
reverse micelles containing different reactants (A and B), the reaction can
take place by
coalescence or aqueous phase exchange between the two reverse micelles.
Figure 1.8 Schematic illustration of various stages in the growth of
nanoparticles in microemulsions (After Leung, at al., 1988).
There are essentially three procedures to form nanoparticles by reversed
micelles:
precipitation, reduction and hydrolysis. Precipitation is usually applied in the
synthesis
of metal sulfate (Qi, et al., 1996), metal oxide (Ayyub, et al., 1990; 1988),
metal
carbonate (Kandori, et al., 1988; Pillai, et al., 1993) and silver halide
(Dvolaitzky, et al.,
1983; Hou and Shah, 1988; Chew, et al., 1990) nanoparticles. In this method
two reverse
micelles containing the anionic and cationic surfactants are mixed. Because
every
reaction takes place in a nanometer-sized water pool, water-insoluble
nanoparticles are
formed.
The reduction procedure is one of the most common ways to prepare metal
nanoparticles
using W/O microemulsions. By dissolving the metal salts in the reversed
micelles, the
salts undergo a dissociation step inside the aqueous domain. Following a
reduction step
(Men+ → Me0), a subsequent precipitation of particles can take place inside the
water
pools. Strong reduction agents such as N2H4, NaBH4 and sometimes hydrogen
gas can
be used.
The hydrolysis procedure is usually used in the preparation of metal oxide
nanoparticles.
It utilizes the hydrolysis properties of metal alkoxide dissolved in oil and
reacting with
water inside the droplets.
1.3.1 Preparation of Nanoparticles of Metals
Since metals display surface catalytic properties, the synthesis of size-
controllable and
monodisperse metal nanoparticles is of considerable importance. The
reduction method
is one of the most common ways to prepare metal nanoparticles through
reverse micelles.
11
Boutonnet et al. have prepared platinum, palladium, rhodium and iridium
nanoparticles
using reverse micelles (Boutonnet, et al., 1982; 1989). H2PtCl6 was dissolved
in
CTAB/water/octanol reverse micelle. Subsequent reduction with hydrazine
produced
nanoparticles. Pd particles were formed by reducing Pd(NH2)4Cl2 or K2PdCl4
with N2H4.
Rhodium particles were formed by reducing RhCl2 with bubbling hydrogen,
whereas
iridium particles could be obtained by bubbling active hydrogen through 2%
Pt-Al2O3 at
70°C. Ag and Au colloidial nanoparticles were successfully prepared by
reducing the
AgNO3 and HAuCl4 in water/cyclohexane/PEGDE or PEGDE/water/n-hexane
reverse
micelles (Barnickel and Wokaun, 1990), where NaBH4 was used as the
reduction
reagent. Silver and copper salts of Aerosol OT can be used for the preparation
of Ag and
Cu nanoparticles (Lisiecki and Pileni, 1993; Pileni, et al., 1993a; 1993b; Petit,
et al.,
1993; Lisiecki and Pileni, 1995). Copper nanosized particles have been
synthesized in
the reverse micelles using hydrazine as a reducing reagent. The size of Cu
nanoparticles
can be controlled by the water content in the reversed micelles (Lisiecki and
Pileni,
1995). Gold and silver nanoparticles were also produced by reducing gold
chloride
tetrahydrate HAuCl4 with citric acid at 80°C for half an hour (Chen, et al.,
1996; Frens,
1973; Enustun and Turkevich, 1963). Nanoparticles of other metals such as
Co (Chen, et
al., 1994; Eastoe, et al., 1996), Ni (Lopez-Quintela and Rivas, 1993) and metal
alloys
FeNi (Lopez-Quintela and Rivas, 1993), Cu3Au (Sangregorio, et al., 1996) and
Co-Ni
(Nagy, 1989) have also been synthesized using the reversed micelles.
1.3.2 Preparation of Nanoparticles of Metal Sulfides
Colloidal semiconductors are attracting much interest due to their
applications as
enhancement of photoreactivity and photocatalysis and non-linear optical
properties.
The key to synthetic investigation of this kind of nanoparticles must be the
careful
control of semiconductor size and size distribution. The precipitation method
is usually
applied in the preparation of metal sulfide particles (Motte, et al., 1992; Hirai,
et al.,
1994; Ward, et al., 1993; Boalkye, et al., 1994; Modes and Lianos, 1989). CdS
particles
have been synthesized in AOT and Triton reversed micelles with functional
surfactant
such as cadmium lauryl sulfate and cadmium AOT (Petit, et al., 1990; Petit
and Pileni,
1988). The average diameters of the particles were found to depend on the
relative
amount of Cd2+ and S2-. The particles obtained from AOT were smaller and
more
monodisperse than those from the Triton reverse micelle. Colloidal CdS was
prepared in
the mixed sodium AOT/cadmium AOT/isooctane reverse micelle (Motte, et al.,
1992).
PbS nanoparticles can be prepared by mixing one polyoxyethylene dodecyl
ether-nhexane
reverse micelle, which supplies Pb2+ from electrolytes such as Pb(NO3)2 or
Pb(ClO4)2, and another reverse micelle that contains S2- from Na2S or H2S
(Ward, et al.,
1993). A number of nanoparticle semiconductors such as CdS (Lianos and
Thomas,
1987; Petit, et al., 1990; Pileni, et al., 1992; Karayigitoglu, et al., 1994), PbS
(Ward, et
al., 1993; Eastoe, et al., 1996), CuS (Lianos and Thomas, 1987), Cu2S (Haram,
et al.,
1996), Ag2S (Motte, et al., 1996), MoS3 (Boalkye, et al., 1994), CdSe
(Steigerwald, et al.,
1988) have also been synthesized using this method.
In recent years apart from the synthesis of nanoparticles, surface
modification of the
metal sulfide particles has attracted much interest. The modification of the
semiconductor surface is also very important either from the point of view of
enhancing
12
the stability of the nanoparticles or for providing unique physical and
chemical
properties. An additional profit from this treatment is that it allows the
particles to be
separated from the micellar solution and redispersed in another solvent.
Some surfacecapped
semiconductor nanoparticles have been synthesized with the cap agents
such as
sodium hexamephosphlate (Meyer, et al., 1984; Petit and Pileni, 1988) of the
surfacecapping
agents such as thiophenol and phenyl (trimethyl) selenium (Steigerwald, et
al.,
1988; Herron, et al., 1990; Dance, et al., 1984).
1.3.3 Preparation of Nanoparticles of Metal Salts
Many metal salts such as silver halide, metal sulfate and metal carbonate
possess unique
properties. Precipitation methods are usually used to prepare the
nanoparticles of these
materials. Silver halide nanoparticles were synthesized by reacting AgNO3
with sodium
halides in Aerosol OT W/O microemulsions (Dvolaitzky, et al., 1983; Hou and
Shah,
1988; Chew, et al., 1990).
However, metal carbonate nanoparticles such as BaCO3, CaCO3 and SrCO3
were
prepared by bubbling CO2 through the reversed micelle solutions containing
the
corresponding aqueous metal hydroxides. Kandori et al. (1987) used the
hexaethylene
glycol dodecyl ether (HEGDE)/water/cyclohexane and calcium AOT based
reverse
micellar system to synthesize CaCO3 nanoparticles with diameters of 5.4 nm.
The
nanoparticle diameter from the CaAOT system was 48–130 nm (Kandori, et
al., 1987;
1988). Nanoparticles of metal sulfate can also be synthesized by the
precipitation
method. Nanoparticles of AgCl (Bagwe and Khilar, 1997) and AgBr (Chew, et
al., 1990;
Monnoyer, et al., 1996) have been synthesized using reverse micelles.
1.3.4 Preparation of Nanoparticles of Metal Oxides
Nanoparticles of metal oxides are usually produced by the hydrolysis method
in which
the metal alkoxides react with water droplets in the reverse micelles.
Nanoparticles of
metal oxides such as ZrO2 (Kawai, et al., 1996), TiO2 (Chang, et al., 1994;
Joselevich
and Willner, 1994; Chhabra, et al., 1995), SiO2 (Osseo-Asare and Arriagada,
1990;
Wang, et al., 1993; Arriagada and Osseo-Asare, 1995; Gan, et al., 1996;
Chang and
Fogler, 1997; Esquena, et al., 1997), GeO2 (Kon-no, 1996), g-Fe2O3 (Lopez-
Perez, et al.,
1997) and F2O3 (Liz, et al., 1994) have been synthesized. GeO2 nanoparticles
can
directly be obtained from AOT-cyclohexane W/O microemulsions by adding
anhydrous
cyclohexane solutions of Ge(OC2H4)4 into the microemulsions. And SiO2
nanoparticles
could be formed by adding Si(OC2H4)4 to the solubilized ammonia aqueous
solution in
AOT and polyoxyethylated nonylphenyl ether W/O microemulsions. Similarly,
ZrO2
nanoparticles can be obtained by hydrolyzing Zr(OC4H9)4 with sulfuric acid in
polyoxyethylene nonylphenyl ether-cyclohexane systems and then washed
with
ammonia aqueous solution. TiO2 nanoparticles can be prepared by adding
benzene
solution of TiCl4 to cetylbenzyldimethylammonium chloride-benzene W/O
microemulsions.
13
1.3.5 Preparation of Nanoparticles of Other Compounds
YBa3CuO7-x particles were synthesized by co-precipitation of the oxalate salts
of Y, Ba
and Cu nitrates in CTAB/n-butanol/n-octane reversed micelles (Ayyub, et al.,
1988;
1990). BaFe12O19 particles were synthesized by the calcination of barium-iron
carbonate
particles made by mixing the two reverse micelles containing the (NH4)2CO3
and a
mixture of aqueous Ba(NO3)2 and Fe(NO3)3 (Pillai, V., et al., 1993).
1.3.6 Synthesis of Nanowires Using Reversed Micelles
The nanoparticles fabricated in the reversed micelles are spherical particles
in most
cases. However, since the optical, electric, and other properties of
nanoparticles are
affected by the shape of nanoparticles, various shapes have been
synthesized. For
example, cubic Pt nanoparticles have been synthesized and they showed
extremely good
catalysis selectivity and activity (Ahmadi, et al., 1996a; 1996b). Addition of
CdS
nanowire into the porous aluminum oxide film will be of potential use in
photoelectronics
(Routkevitch, et al., 1996). Qi et al., using reversed micelles of TX-
100/hexanol.cyclohexane/water, have successfully synthesized cubic BaSO4
nanoparticles (Qi, et al., 1997). They have found that the water content in the
reversed
micelles greatly affected the shape of the nanoparticles. Cubic nanoparticles
of BaSO4
were obtained in the higher content of water. On the other hand, in the non-
ionic reverse
micelle C12E4/cyclohexane, adding 0.1 M BaCl2 and Na2CO3 aqueous solution
to 0.2 M
C12E4/cyclohexane solution, and mixing the two reversed micelles, BaCO3
nanowires
were obtained (Fig. 1.9). Hopwood and Mann have also synthesized BaSO4
nanowire
using reversed micelles (Hopwood and Mann, 1997).
Figure 1.9 TEM micrographs and electron diffraction pattern of BaCO3
anowires synthesized in reversed micells (Qi, et al., 1997).
14
1.3.7 Synthesis of Composite Nanoparticles Using Reversed Micelles
Composite nanoparticles are composed of two kinds of nanoparticles, not
only modifing
the properties of single semiconductor nanoparticles, but also producing
some new
electric and optical properties. The composite semiconductor nanoparticles
can be
divided into sandwich type and shell-core type. Sandwich type CdS-TiO2
(Spanhel, et al.,
1987; Gopidas, et al., 1990; Lawless, et al., 1995 and CdS-SnO2 (Nasr, et al.,
1997) have
been prepared and show prospects in solar cell application. On the other
hand, shell-core
type composite nanoparticles such as CdS-ZnS (Hirai, et al., 1994), CdS/PbS
(Zhou, et
al., 1993; 1994), CdS/HgS (Hasselbarth, et al., 1993; Mews, et al., 1994;
Schooss, et al.,
1994; Kamalov, et al., 1996; Mews, et al., 1996), CdS/Ag2S (Han, et al., 1998),
CdS/CdSe (Tian, et al., 1996; Peng, et al., 1997), CdSe/ZnS (Kortan, et al.,
1990; Hines
and Guyot-Sinnest, 1996; Dabbousi, et al., 1997), CdSe/ZnSe (Hoener, et al.,
1992;
Danek, et al., 1996) have been synthesized using different methods. They
showed
enhancement of photocatalytic efficiency and strong enhancement of
emission.
Reversed micelle is also an important method for synthesizing the composite
nanoparticles. So far reversed micelles have been successfully used to
synthesize
composite nanoparticles such as CdS-ZnS (Hirai, et al., 1994), CdS-Ag2S (Han,
et al.,
1998) and CdSe-ZnS (Kortan, et al., 1990), CdSe-ZnSe (Hoener, et al., 1992).
For shell-core type nanoparticles the synthesis contains two steps: the first
step is the
formation of core nanoparticles in the reverse micelles and the second step is
the growth
of the shell particles on the core. CdS/ZnS (where CdS is the core and the ZnS
is the
shell) is a typical shell-core type composite nanoparticles and can be
synthesized as
follows. Mixing the reverse micelles containing Cd2+ and S2- in a 1 : 2 ratio,
one can
obtain the core CdS reverse micelle solution. In this reversed micelle S2- is
excess. After
several minutes, Zn2+ containing reverse micelle was added. ZnS precipitated
in the core
CdS nanoparticles, and a shell-core type CdS/ZnS composite nanoparticle was
obtained.
For the ZnS/CdS composite, the same method can be used, only changing the
order of
synthesis. Using this method, Ma et al. have prepared composite
nanoparticles of
CdS/ZnS and ZnS/CdS. Another type of composite nanoparticle contains two
metals not
in the 1 : 1 ratio. They can be synthesized as follows. In the synthesis of
coated
Ag2S/CdS nanoparticles, after mixing the two reverse micelles containing the
equal
molar Cd2+ and S2-, AgNO3 was added to the mixed reversed micelles. The
reaction of
2Ag+ + CdS(s) → Cd2+ + Ag2S. Coated Ag2S/CdS small particles with a
diameter of ~10
nm was obtained. The nanoparticles showed large nonlinear absorption.

Вам также может понравиться