You are on page 1of 24

Available online at www.sciencedirect.

com

Progress in Crystal Growth and Characterization of Materials


55 (2009) 22e45
www.elsevier.com/locate/pcrysgrow

Synthesis, properties, and applications of magnetic


iron oxide nanoparticles
Amyn S. Teja*, Pei-Yoong Koh
School of Chemical & Biomolecular Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0100, USA

Abstract

Magnetic nanoparticles exhibit many interesting properties that can be exploited in a variety of
applications such as catalysis and in biomedicine. This review discusses the properties, applications, and
syntheses of three magnetic iron oxides e hematite, magnetite, and maghemite e and outlines methods of
preparation that allow control over the size, morphology, surface treatment and magnetic properties of
their nanoparticles. Some challenges to further development of these materials and methods are also
presented.
Ó 2008 Elsevier Ltd. All rights reserved.

PACS: 61.46.þw; 82.Rx; 81.40.z; 81.16.c; 75

Keywords: B1. Oxides; B1. Nanomaterials; B2. Magnetic materials; A2. Supercritical fluid technology; A2. Hydro-
thermal technology

1. Introduction

Iron oxides exist in many forms in nature, with magnetite (Fe3O4), maghemite (g-Fe2O3),
and hematite (a-Fe2O3) being probably the most common [1]. These three oxides are also very
important technologically, and they are therefore the subject of this review. Some of their
physical and magnetic properties are summarized in Table 1.
Hematite is the oldest known of the iron oxides and is widespread in rocks and soils. It is
also known as ferric oxide, iron sesquioxide, red ochre, specularite, specular iron ore, kidney
ore, or martite. Hematite is blood-red in color if finely divided, and black or grey if coarsely

* Corresponding author.
E-mail address: amyn.teja@chbe.gatech.edu (A.S. Teja).

0960-8974/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pcrysgrow.2008.08.003
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 23
55 (2009) 22e45
Table 1
Physical and magnetic properties of iron oxides [1].
Property Oxide
Hematite Magnetite Maghemite
Molecular formula a-Fe2O3 Fe3O4 g-Fe2O3
Density (g/cm3) 5.26 5.18 4.87
Melting point ( C) 1350 1583e1597 e
Hardness 6.5 5.5 5
Type of magnetism Weakly ferromagnetic Ferromagnetic Ferrimagnetic
or antiferromagnetic
Curie temperature (K) 956 850 820e986
MS at 300 K (A-m2/kg) 0.3 92e100 60e80
Standard free 742.7 1012.6 711.1
energy of formation DGf
(kJ/mol)
Crystallographic system Rhombohedral, hexagonal Cubic Cubic or tetrahedral
Structural type Corundum Inverse spinel Defect spinel
Space group R3c (hexagonal) Fd3m P4332 (cubic); P41212
(tetragonal)
Lattice parameter (nm) a ¼ 0.5034, c ¼ 1.375 (hexagonal) a ¼ 0.8396 a ¼ 0.83474 (cubic);
aRh ¼ 0.5427, a ¼ 55.3 a ¼ 0.8347, c ¼ 2.501
(rhombohedral) (tetragonal)

crystalline. It is extremely stable at ambient conditions, and often is the end product of the
transformation of other iron oxides. Magnetite is also known as black iron oxide, magnetic iron
ore, loadstone, ferrous ferrite, or Hercules stone. It exhibits the strongest magnetism of any
transition metal oxide [1,2]. Maghemite occurs in soils as a weathering product of magnetite, or
as a product of heating of other iron oxides. It is metastable with respect to hematite, and forms
continuous solid solutions with magnetite [3].
The crystal structure of the three iron oxides can be described in terms of close-packed
planes of oxygen anions with iron cations in octahedral or tetrahedral interstitial sites. In
hematite, oxygen ions are in a hexagonal close-packed arrangement, with Fe(III) ions occu-
pying octahedral sites (Fig. 1a). In magnetite and maghemite, the oxygen ions are in a cubic

Fig. 1. Crystal structures of (a) hematite and (b) magnetite.


24 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
close-packed arrangement (Fig. 1b). Magnetite has an inverse spinel structure with Fe(III) ions
distributed randomly between octahedral and tetrahedral sites, and Fe(II) ions in octahedral
sites [4]. Maghemite has a spinel structure that is similar to that of magnetite but with vacancies
in the cation sublattice. Two-thirds of the sites are filled with Fe(III) ions arranged regularly,
with two filled sites being followed by one vacant site [1]. Haneda and Morrish [5] found that
the degree of vacancy ordering decreases with decreasing particle size, with no vacancy
ordering in maghemite smaller than about 20 nm.

2. Magnetic behavior of iron oxides

The iron atom has a strong magnetic moment due to four unpaired electrons in its 3d
orbitals. When crystals are formed from iron atoms, different magnetic states can arise as
shown in Fig. 2. In the paramagnetic state, the individual atomic magnetic moments are
randomly aligned with respect to each other, and the crystal has a zero net magnetic moment. If
this crystal is subjected to an external magnetic field, some of these moments will align, and the
crystal will attain a small net magnetic moment. In a ferromagnetic crystal, all the individual
moments are aligned even without an external field. A ferrimagnetic crystal, on the other hand,
has a net magnetic moment from two types of atoms with moments of different strengths that
are arranged in an antiparallel fashion (see Fig. 2). If the antiparallel magnetic moments are of
the same magnitude, then the crystal is antiferromagnetic and possesses no net magnetic
moment.
In a bulk ferromagnetic material, the magnetization M is the vector sum of all the magnetic
moments of the atoms in the material per unit volume of the material. The magnitude of M is
generally less than its value when all atomic moments are perfectly aligned, because the bulk
material consists of domains (Fig. 3) with each domain having its own magnetization vector
arising from an alignment of atomic magnetic moments within the domain. The magnetization
vectors of all the domains in the material may not be aligned, leading to a decrease in the
overall magnetization (Fig. 3). When the length scale of the material becomes small, however,
the number of domains decreases until there is a single domain when the characteristic size of
the material is below some critical size dC.

Fig. 2. Alignment of individual atomic magnetic moments in different types of materials.


A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 25
55 (2009) 22e45

Fig. 3. Magnetic domains in a bulk material.

If an external magnetic field of strength H is applied to a ferromagnet of magnetic strength


M, the magnetization curve of Fig. 4 is obtained showing that M increases with H until
a saturation value MS is reached. The magnetization curve displays a hysteresis loop, because
all domains do not return to their original orientations when H is decreased after the saturation
magnetization value is attained. Thus, when H returns to zero, there is a remnant magnetization
MR which can only be removed by applying a coercive field HC in the opposite direction to the

Fig. 4. Magnetization M as a function of an applied magnetic field H.


26 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
initially applied field. A single domain magnetic material has no hysteresis loop and is said to
be superparamagnetic. Iron oxide nanoparticles smaller than about 20 nm often display
superparamagnetic behavior at room temperature [1].
The ordered arrangement of magnetic moments decreases with increasing temperature due
to thermal fluctuations of the individual moments. Beyond the Néel or Curie temperature, the
material becomes disordered and loses its magnetization. The transition temperature is termed
the Curie temperature TC for ferromagnetic and ferrimagnetic substances, and the Néel
temperature TN for antiferromagnetic substances. Superparamagnetic particles are usually
ordered below a blocking temperature, TB.
Hematite is paramagnetic at temperatures above its Curie temperature of 956 K. At room
temperature, it is weakly ferromagnetic and undergoes a phase transition at 260 K (the Morin
temperature, TM) to an antiferromagnetic state [1,6e8]. The magnetic behavior of hematite
depends on crystallinity, particle size and on the extent of cation substitution [1,9,10]. The
Morin temperature of hematite decreases as the particle size decreases (Fig. 5) and tends to
vanish for particles smaller than 8e20 nm [7,11,12]. Poor crystallinity and substitution of
cations tend to lower TC and TM (and may even completely suppress the Morin transition) at all
temperatures [13]. Among the cations that lower TM of hematite are Al, Ga, Cr, In, Mn, Sn and
Ti [6,14e17], whereas Rh raises TM instead [18,19].
Magnetite is ferrimagnetic at room temperature and has a Curie temperature of 850 K [1,13].
Magnetite particles smaller than 6 nm are superparamagnetic at room temperature, although
their magnetic properties depend strongly on the methods used in their synthesis [20e25].
Margulies et al. reported that the magnetic properties of nanosized magnetite appeared to
depend strongly on changes in the crystal morphology [21,26]. The crystal morphology affects
coercivity in the order: spheres < cubes < octahedra in line with the increase in the number of
magnetic axes along this series of shapes. Magnetite particles with coercivities ranging from
2.4 to 20 kA/m have been produced by controlling their synthesis conditions [1,27].

Fig. 5. Morin temperature of hematite as a function of the diameter d of the particle (adapted from ref. [7]).
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 27
55 (2009) 22e45
Maghemite is ferrimagnetic at room temperature, unstable at high temperatures, and loses its
susceptibility with time [28]. However, it can be stabilized by doping with other metal ions
[29]. For example, doping by tetravalent ions such as tin can be used to stabilize its trans-
formation into hematite at elevated temperatures [30]. Its Curie temperature is difficult to
determine experimentally because maghemite undergoes an irreversible crystallographic
change to hematite at w400  C, with a consequent dramatic loss of magnetization. However,
the Curie temperature is believed to be between 820 K and 986 K [1]. Maghemite particles
smaller than 10 nm are superparamagnetic at room temperature [31]. Aggregation of ultrafine
maghemite particles sometimes leads to magnetic coupling between particles and ordering of
the magnetic moment, which is termed superferromagnetism [1].
Surface effects have a strong influence on the magnetic properties of iron oxide nano-
particles [32]. As a result, their net magnetization decreases at a faster rate with increasing
temperature than that of the corresponding bulk material because a large fraction of the atoms
are near the surface where the exchange field is lower [33,34]. Modifications of the surface by
chemical treatments have also been found to affect the coercivity of oxide nanoparticles [35e
38]. As a result of this dependence on size and surface treatment, nanostructuring of magnetic
materials via the method of preparation can be used to tune magnetic properties as discussed
below.

3. Applications

The magnetic properties of iron oxides have been exploited in a broad range of applications
including magnetic seals and inks, magnetic recording media, catalysts, and ferrofluids [39e
49], as well as in contrast agents for magnetic resonance imaging and therapeutic agents for
cancer treatment [39,42,48,50e55]. These applications demand nanomaterials of specific sizes,
shapes, surface characteristics, and magnetic properties.
In data storage applications, the particles must have a stable, switchable magnetic state that
is not be affected by temperature fluctuations [45]. For optimum performance in recording, the
particles should exhibit both high coercivity and high remanence, and they should be uniformly
small, and resistant to corrosion, friction, and temperature changes [56]. Maghemite is useful in
recording and data storage applications because of its chemical and physical stability [57]. It is
often doped or coated with 1e5% cobalt in order to improve its coercivity and storage capacity.
Coated nanoparticles have greater thermal stability than their doped counterparts and display
uniaxial magnetic anisotropy [1,57]. As a result, cobalt-modified maghemite particles are the
predominant material for use in video tapes, high bias audio tapes and magnetic discs [1,57].
Maghemite nanoparticles embedded in a nonmagnetic matrix also display giant magnetore-
sistance, which is a decrease in the resistance due to an applied magnetic field, and is useful in
magnetic recording heads and sensing elements in magnetometers [58].
The use of magnetite in ferrofluids was originally proposed for high-performance seals in
space applications. Ferrofluids contain nanometer sized superparamagnetic particles dispersed
in aqueous or organic media [1]. A ferrofluid has no net magnetic moment except when it is
under the influence of an applied field. An external magnet is therefore able to trap the fluid in
a specific location to act as a seal. Ferrofluids have interesting properties such as magnetic field
dependent optical anisotropy that could prove useful in optical switches and tunable diffraction
gratings. They are currently employed in sealing computer disk units, and in vibrating envi-
ronments in place of conventional seals [46]. Other uses include NMR probes for oil pro-
specting [1,59] and they have also been suggested for use in eye surgery to repair damaged
28 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
retinas [46]. Ferrofluids exhibit a high degree of colloidal stability in a magnetic field gradient
[1]. One of the keys to improving their performance in these applications is to make the
particles smaller and more uniform.
The use of magnetite nanoparticles has received considerable attention in the development
of immunoassays, magnetic resonance imaging contrast agents, and targeted drug delivery
vehicles, as well as in magnetic hyperthermia [45,46,50e52,60e66]. These applications
require particles that exhibit superparamagnetic behavior at room temperature. Remnant
magnetization could lead to agglomeration of these particles, and this must obviously be
avoided within the body to prevent blockage of blood vessels. In addition, applications in
biology and medical diagnosis require stable magnetic particles in water at neutral pH and
physiological conditions [2,45,46]. The colloidal stability of magnetic fluids depends on the
dimensions of the particles, which should be sufficiently small to minimize precipitation due to
gravitation forces, and on the charge and surface chemistry. Magnetite or maghemite are by far
the most commonly employed materials for biomedical applications [2,52]. Functionalized
superparamagnetic magnetite and maghemite nanoparticles, in combination with an external
magnetic field, allow delivery of drugs to the desired target where the medication can be
released locally [31,67]. This allows the dosage of the medication to be reduced and any
adverse effect of the drugs to be kept to a minimum [67]. The surfaces of nanoparticles used in
drug delivery are generally functionalized with drugs, proteins, and genetic materials to achieve
localized delivery of these therapeutic agents [31,68].
Magnetite and maghemite have attracted attention in biomedical applications because of
their biocompatibility and low toxicity in the human body [2,45,69,70]. A major area of
application has been the field of bio-assays where the magnetic properties have been exploited
in vitro to manipulate magnetite nanoparticles with an external magnetic field [2,71]. Wang and
co-workers [72] have developed extremely sensitive magnetic microarrays using ferromagnetic
sensors to detect binding of target DNA and proteins.
Magnetic nanoparticles have been used in vivo as magnetic resonance imaging (MRI)
contrast agents for molecular and cell imaging [2,45,62,65,66,70,73e82]. Superparamagnetic
magnetite is used as the core in these agents which are used to differentiate between healthy
and diseased tissue. The superparamagnetic particles are generally coated with a poly-
saccharidic layer for colloidal stability [83]. In vivo MRI cell tracking has been successfully
performed by Song and co-workers [84]. Magnetic particles with a polymer coating have been
used in cell separation [50,51,68,85,86], protein purification [87], environment and food
analyses [88], organic and biochemical syntheses [89,90], industrial water treatment [91] and
biosciences [75,77,92,93]. Encapsulation of magnetic nanoparticles with organic polymers is
used to enhance their chemical stability, dispersability and functionality [70].
Another application of magnetic nanoparticles is hyperthermia in cancer therapy. Super-
paramagnetic magnetic nanoparticles when exposed to an alternating magnetic field can be
used to heat tumor cells to 41e45  C, where tissue damage for normal tissue is reversible while
the tumor cells are irreversibly damaged [31].
Magnetite and hematite have been used as catalysts for a number of industrially important
reactions [1,39,55,94e99], including the synthesis of NH3 (the Haber process), the high
temperature water gas shift reaction, and the desulfurization of natural gas. Other reactions
include the dehydrogenation of ethyl benzene to styrene, the FishereTropsch synthesis for
hydrocarbons, the oxidation of alcohols, and the large scale manufacture of butadiene.
Magnetite and hematite are semiconductors and can catalyze oxidation/reduction reactions
[97,100e104]. Hematite has also been used as a support material for gold in catalysts for the
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 29
55 (2009) 22e45
oxidation of carbon monoxide at low temperature [105e107]. Iron oxides can be used as acid/
base catalysts [94] and to catalyze the degradation of acrylnitrileebutadieneestyrene copol-
ymer into fuel oil. Hematite has been used as photocatalyst for the degradation of chlorophenol
and azo dyes [108], whereas maghemite and magnetite/carbon composites have been found
useful for reducing the amount of undesirable N2 in fuel oil [109].
All three forms of magnetic iron oxide are commonly used in synthetic pigments in paints,
ceramics, and porcelain [1]. They possess a number of desirable attributes for these applications
because they display a range of colors with pure hues and high tinting strength. They are also
extremely stable and highly resistant to acids and alkalis. Pigments based on hematite are red,
those based on maghemite are brown, and magnetite-based pigments are black [110]. The
transparent yellow pigments based on goethite can be transformed into the transparent red
pigments of hematite by calcination at 400e500  C [111]. These pigments are widely used in
water-repellent stains for wood as they enable the wood grain to be seen while still providing
protection against the damaging effects of sunlight. Pigments made from magnetite are also
used in magnetic ink character recognition devices, and superparamagnetic magnetite particles
are used in metallography for detecting flaws in engines [1].
As noted above, many of the useful attributes of iron oxides depend on the preparation
method for the nanomaterials. The preparation method plays a key role in determining the
particle size and shape, size distribution, surface chemistry and therefore the applications of the
material [112]. In addition, the preparation method also determines the degree of structural
defects or impurities present in the particles, and the distribution of such defects [45]. Many
synthesis routes have been developed to achieve proper control of particle size, polydispersity,
shape, crystallinity, and the magnetic properties [43,45,46,48,53,61,113e115]. Some of these
methods are described in the following sections.

4. Methods of preparation

4.1. Gas phase methods

Gas phase methods for preparing nanomaterials depend on thermal decomposition (pyrol-
ysis), reduction, hydrolysis, disproportionation, oxidation, or other reactions to precipitate solid
products from the gas phase [116]. In the chemical vapor deposition (CVD) process, a carrier
gas stream with precursors is delivered continuously by a gas delivery system to a reaction
chamber maintained under vacuum at high temperature (>900  C) [114,117]. The CVD
reactions take place in the heated reaction chamber and the products combine to form clusters
or nanoparticles. Growth and agglomeration of the particles are mitigated via rapid expansion
of the two-phase gas stream at the outlet of the reaction chamber (Fig. 6). Subsequent heat
treatment of the synthesized nanopowders in various high-purity gas streams allows compo-
sitional and structural modifications, including particle purification and crystallization, as well
as transformation to a desirable size, composition, and morphology [114,116]. The CVD
process has been employed to deposit iron oxide by the reaction of a halide, such as iron
trichloride, with water at 800e1000  C [118]. The success of this method depends on low
concentrations of precursor in the carrier gas, as well as rapid expansion and quenching of the
nucleated clusters or nanoparticles as they exit from the reactor [114,117,119].
The use of metallo-organics as precursors (in the MOCVD process) allows reactions to take
place at somewhat lower temperatures (300e800  C) and at pressures varying from less than
1 Torr to ambient [116]. Iron oxide thin films have been obtained via decomposition of
30 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45

Fig. 6. Schematic diagram of a CVD apparatus (adapted from ref. [114]).

acetylacetonate at 400e500  C, and iron trifluoro-acetylacetonate at 300  C in oxygen


[120,121]. Other precursors include tris(2,2,6,6-tetramethyl-3,5-heptadionato) Fe(III) and tris(t-
butyl-3-oxo-butanoato) Fe(III) [122]. Recently, Park et al. [123] deposited magnetite thin films
using Fe(II) dihydride complexes H2Fe[P(CH3)3]4 at 300  C in oxygen. Direct growth of
magnetite has been achieved by a low-pressure CVD using metal-organic ferric dipivaloyl-
methanate as a precursor. Upon oxidation, these films were converted to maghemite [124].
Another MOCVD reaction uses a microwave plasma to decompose either iron cyclopentadienyl
or iron acetylacetonate in an oxygen atmosphere at 300e500  C and 1e20 Torr [116,125].
Laser pyrolysis of organometallic precursors [114,126e128] is based on the resonant
interaction between laser photons and at least one gaseous species, reactant or sensitizer.
A sensitizer is an energy transfer agent that is excited by absorption of CO2 laser radiation and
transfers the absorbed energy to the reactants by collision [129]. The method involves heating
a flowing mixture of gases with a continuous wave CO2 laser to initiate and sustain a chemical
reaction until a critical concentration of nuclei is reached in the reaction zone, and homoge-
neous nucleation of particles occurs [70]. A schematic representation of a CO2 laser pyrolysis
device is shown in Fig. 7 [70]. The nucleated particles formed during the reaction are entrained
by the gas stream and are collected at the exit [45].
Well-crystallized and uniform iron oxide nanoparticles, including nanoparticles of
hematite and maghemite, have been obtained in one step using laser pyrolysis
[126,128,130e134]. Iron pentacarbonyl is commonly used as a precursor in this method
[132] and ethylene is used as the carrier gas to transport the carbonyl vapor to the reaction
zone, as ethylene does not absorb radiation at the laser wavelength [45]. The iron penta-
carbonyl decomposes into iron and carbon monoxide [45,132] which are then oxidized using
air, introduced into the system with the iron pentacarbonyl vapor or mixed with argon [45].
Iron particles with 14 nm mean diameter and about 4 nm oxide shell thickness have been
produced by laser pyrolysis of iron pentacarbonyl and ethylene mixtures followed by
a controlled step-by-step passivation process.
The effect of process conditions on the structural and magnetic properties of maghemite
nanoparticles produced by laser pyrolysis have been studied by Veintemillas-Verdaguer and
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 31
55 (2009) 22e45

Fig. 7. Schematic representation of laser pyrolysis device (adapted from ref. [129]).

co-workers [135]. They reported that the particle size depends on the oxygen content of the gas
phase and is independent of the laser power, and suggested that the amount of oxygen be used
to tune the size and crystallinity of the particles [135].
Although gas phase methods are able to deliver high quality products, the yields are usually
low and scale-up of the equipment is challenging. Variables such as oxygen concentration, gas
phase impurities, and the heating time must be controlled precisely to obtain pure products. The
equipment used in these methods also tends to be expensive.

4.2. Liquid phase methods

Liquid phase methods are generally inexpensive and offer better yields of products as well as
ease of surface treatment [2]. Most nanoparticles prepared to date by such methods have been
via coprecipitation from aqueous solutions, although other liquid solvents can also be used. It
has been shown that spherical magnetite particles with mean diameters ranging from 30 to
100 nm can be obtained by the reaction of a Fe(II) salt, a base and a mild oxidant (nitrate ions)
in aqueous solutions [136]. Stoichiometric mixtures of ferrous and ferric hydroxides can also be
reacted in aqueous media to yield homogeneous spherical particles of either magnetite or
maghemite [45,70]. The phase and size of the particles depend on the concentration of cations,
the counterions present, and the pH of the solution [137e141]. In addition, it is possible to
control the mean size of the particles by adjusting the pH and the ionic strength
[45,70,142,143].
Due to the large surface-area to volume ratio, nanoparticles formed by liquid phase
coprecipitation tend to aggregate in solution in order to reduce their surface energy [143]. The
suspension of nanoparticles can be stabilized by adding anionic surfactants as dispersing agents
[144,145]. The nature of the counterions, pH, and ionic strength can then be used to stabilize
32 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
the charged particles via interactions between electrical double layers [144,145]. Increasing the
concentration of inert electrolyte in the system compresses the double layer and promotes
coagulation [1]. However, suspensions of iron oxide that are stabilized entirely by electrostatic
repulsion are too sensitive to external conditions such as pH and ionic strength to offer any
flexibility in engineering the surface properties of the particles [145].
Stabilization can also be achieved by coating the particle surfaces with proteins [52,90,146],
starches [75], non-ionic detergents [147], or polyelectrolytes [144,148]. Adsorption of such
substances stabilizes the particles at electrolyte concentrations that would otherwise be high
enough for coagulation to occur [1]. Shimoiizaka prepared water-based magnetic fluids by
dispersing magnetite particles in water containing oleic acid [149]. Following this work,
Khalafalla and Reimers [149] produced stable aqueous magnetic fluids using dodecanoic acid
as a dispersing agent. The use of saturated and unsaturated fatty acids to stabilize magnetic
fluids has been investigated by Wooding et al. [150].
High quality monodisperse and monocrystalline iron oxide nanoparticles can be produced by
the thermal decomposition of organometallic precursors in organic solvents containing stabi-
lizing surfactants such as oleylamine [151,152], oleic acid [152e157], and steric acid [2].
Precursors that have been investigated include iron acetylacetonate [153e155], iron carbox-
ylate [158e161], iron cupferronates [162] and iron carbonyls [2,163]. High quality magnetite
nanoparticles with diameters ranging from 3 to 20 nm were synthesized via thermal decom-
position of Fe(III) acetylacetonate in phenyl/benzyl ether and 2-pyrrolidone [152,164,165].
More recently, Stefanescu et al. investigated the synthesis of maghemite by thermal decom-
position of some complex combinations of Fe(III) with carboxylate type ligands obtained from
the redox reaction between polyols and ferric nitrate. Maghemite was obtained at 250e300  C
and hematite at 400e500  C.
Sun and co-workers have reported the preparation of monodisperse iron oxide nanoparticles
by the thermal decomposition of iron acetylacetonate [152]. They also proposed a simple
method to transform hydrophobic nanoparticles into hydrophilic ones by adding bipolar
surfactants and produced 4 nm magnetite nanoparticles by decomposition of Fe(III) acetyla-
cetonate in a mixture of phenyl ether, 1,2-hexadecanediol, oleic acid, and oleylamine [164]. It
has also been reported that 6e7 nm maghemite nanoparticles can be produced via the reaction
of iron cupferronates with trioctylamine at 300  C [162]. Maghemite nanoparticles with sizes
ranging from 4 to 16 nm have been produced by decomposition of iron pentacarbonyl in octyl
ether and oleic acid or lauric acid [166]. These results show the effectiveness of the thermal
decomposition method for the synthesis of iron oxide nanoparticles. However, the presence of
residual surfactants may hamper the efficiency of subsequent surface modification of the
synthesized nanoparticles [2,167]. In addition, the use of toxic solvents and surfactants may be
detrimental to the biocompatibility of the product [2,167].

4.3. Two-phase methods

Water-in-oil (w/o) microemulsions consisting of nanosized water droplets dispersed in an oil


phase and stabilized by surfactant molecules at the water/oil interface have been widely used to
obtain iron oxide nanoparticles [114,168e171]. The surfactant-covered water pools offer
a unique microenvironment for the formation of nanoparticles and for limiting their growth.
The size of the microemulsion droplets is determined by the water to surfactant ratio, although
the eventual size of the nanoparticles may also be influenced by factors such as concentration of
reactants (especially surfactant) and flexibility of the surfactant film [171].
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 33
55 (2009) 22e45
There are several ways to utilize microemulsions to produce nanoparticles. In one method,
reactants A and B are dissolved in the aqueous phases of two identical w/o microemulsions and
form an AB precipitate upon mixing (see Fig. 8a). The precipitate is confined to the interior of
the droplets, thereby limiting the size and shape of the particle formed [169]. In another
method, nanoparticles are produced by the addition of a reducing or precipitating agent to
a microemulsion containing the primary reactant dissolved in an aqueous phase (Fig. 8b). The
reducing or precipitating agent can be a liquid such as hydrazine or a gas such as hydrogen.
Fig. 8c depicts another method for the formation of oxide, hydroxide or carbonate precipitates
by bubbling gases like O2, NH3, or CO2 through a microemulsion containing soluble salts of the
cations [169].
Water-in-oil microemulsions have been used to synthesize iron oxide [169,171e176],
metallic iron nanoparticles [169,171], magnetic polymeric iron oxide nanoparticles [177e180],
and silica-coated iron oxide nanoparticles [181e183]. A variety of surfactants have been used
when preparing these materials, including bis(2-ethylhexyl) sulfosuccinate (AOT), sodium
dodecyl sulfate (SDS), cetyltrimethylammonium bromide (CTAB), polyvinylpyrrolidone

Fig. 8. Schematic representation of nanoparticle synthesis in microemulsions (a) by mixing two microemulsions, (b) by
adding a reducing agent, and (c) by bubbling gas through the microemulsion.
34 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
(PVP), diethyl sulfosuccinate (DES) [171,172,174,175,177,181e185]. Iron-manganese mixed
oxides have also been produced by Herranz and co-workers [186] using microemulsions.
Larpent et al. accomplished the catalytic oxidation of alkanes in water microdroplets con-
taining an iron salt and hydrogen peroxide dispersed in an alkane continuous phase [187]. They
showed that the catalytic process was favored by the low interfacial waterehydrocarbon tension
and the very high interfacial area in liquideliquid dispersions. The use of microemulsions for
the preparation of silica-encapsulated magnetic nanocomposites has also been investigated
[70,181,188]. Yang et al. [71] have synthesized ultrafine magnetite-containing spherical silica
nanoparticles doped with proteins via reverse-micelle microemulsions.
Significant disadvantages of the microemulsion methods are the difficulty in their scale-up,
and the adverse effects of residual surfactants on the properties of the particles.

4.4. Sol-gel methods

Sol-gel methods generally refer to the hydrolysis and condensation of metal alkoxides or
alkoxide precursors, leading to dispersions of oxide particles in a ‘‘sol’’. The ‘‘sol’’ is then dried
or ‘‘gelled’’ by solvent removal or by chemical reaction. The solvent used is generally water,
but the precursors can also be hydrolyzed by an acid or base. Basic catalysis induces the
formation of a colloidal gel, whereas acid catalysis yields a polymeric form of the gel [110].
The rates of hydrolysis and condensation are important parameters that affect the properties of
the final products. Smaller particle sizes are obtained at slower and more controlled hydrolysis
rates. The particle size also depends on the solution composition, pH, and temperature [114].
Magnetic ordering in the sol-gel system depends on the phases formed and the particle volume
fraction, and is very sensitive to the size distribution and dispersion of the particles [114]. In the
case of nanocomposites derived from gels, structural parameters and material porosity are
determined by the rate of hydrolysis and condensation of the gel precursors and also by other
oxidation-reduction reactions that occur during the gelling and subsequent heat treatment stages
[114].
Iron oxideesilica aerogel composites have been prepared by the sol-gel method [189e191]
and found to be 2e3 orders of magnitude more reactive than conventional iron oxide [110]. The
increase in reactivity was attributed to the large surface area of iron oxide nanoparticles sup-
ported on the silica aerogel [96,192]. Commercial precursors (TEOS and Fe(III) solutions) were
dissolved in an alcoholic aqueous medium, and the gels formed after a few days were heated to
produce the final materials [114,193,194]. Ferric nitrate, ferric acetylacetonate and ferric
chloride were used as metal oxide precursors [189,191e196], although the use of the metallic
complex FeNa(EDTA) and a mixture of this metallic complex with ferric nitrate has also been
reported [197]. In experiments with only pure metallic complex, iron oxide nanoparticles in the
range 20e160 nm were obtained. Also, the low solubility of EDTA salt in the solvent prevented
the synthesis of high iron content aerogels when this complex is used as a precursor [197].
Aerogels generated from both pure ferric nitrate and its mixture with the metallic complex
exhibit low saturation magnetization values of 14 and 8.5 emu/g, respectively. This compares
with bulk maghemite values of 74e76 emu/g [110]. The low magnetization values suggest that
the aerogels may not be suitable for magnetic applications, although they may still be useful in
catalysis and other applications.
In most methods of preparing iron oxide-silica composites, iron oxide precursors were
initially mixed with silica precursors in a solvent to form a ‘‘sol’’ [189,191e203]. Recently,
Popovici and co-workers produced iron oxideesilica composites using a novel synthesis route
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 35
55 (2009) 22e45
which consists of impregnating silica wet gels with anhydrous Fe(II) precursors followed by
supercritical drying of the gels by ethanol [204]. Single-phase maghemite dispersed nano-
particles embedded in silica aerogels were obtained without post-annealing. The nano-
composites exhibited a high saturation magnetization value and were superparamagnetic at
room temperature. The overall success was believed to be due to the impregnations of
precursors after gelation, the careful exchange of water with ethanol before impregnation, and
use of the anhydrous ferric salt instead of the hydrated salt [110,197,204]. The sol-gel method
has also been used to synthesize magnetite and maghemite thin films [205e207], transparent
iron-doped titanium oxide thin films [208], ferroelectromagnetic bismuth iron oxide films
[209], mixed iron oxides [95,210e214], and iron oxide-alumina nanocomposites [215].
Disadvantages of the sol-gel methods include contamination from byproducts of reactions,
as well as the need for post-treatment of the products.

4.5. High pressure hydrothermal methods

High pressure hydrothermal methods rely on the ability of water at elevated pressures and
temperatures to hydrolyze and dehydrate metal salts, and the very low solubility of the resulting
metal oxides in water at these conditions to generate supersaturation [114]. The elevated
temperatures favor high dehydration rates, as does the high diffusivity of reactants in water at
these conditions [225,226]. Very high supersaturations can be achieved in this process because
of the very low solubility of metal hydroxides and oxides, so that very fine crystals are obtained
[110,224,225,227]. Parameters such as pressure, temperature, reaction time, and the precursore
product system can be tuned to maintain high nucleation rates and to control growth [216,217].
The process is environmentally benign and versatile, since it does not involve any organic
solvents or post-treatments such as calcination [218]. As a result, high pressure hydrothermal
processes have been widely investigated for the synthesis of metal oxides [43,219e222] as
powders, nanoparticles and single crystals [43,114,219e233].
The hydrothermal synthesis of iron oxides can also be performed in situ within porous
structures. Xu and Teja [48,234] have successfully deposited hematite in the pores of activated
carbon pellets using supercritical water. The hematite nanoparticles were 16e36 nm in
diameter and were uniformly distributed throughout the pellets (see Fig. 9). The resulting
activated carboneiron oxide nanocomposites were used as catalysts in the oxidation of
propanal at rates that were an order of magnitude greater than those for activated carbon
catalysts [235].

Fig. 9. Distribution of hematite nanoparticles deposited (a) on the surface and (b) in the interior of activated carbon
pellets by supercritical water.
36 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
The hydrothermal method has been successfully employed to synthesize fine iron oxide
nanoparticles in a ‘‘continuous flow’’ reactor by Matson and co-workers [236e238]. Their
method involves rapid heating of flowing solutions by contact with supercritical water and
residence times of 5e30 s in the reactor to control growth [110]. A schematic of their exper-
imental setup is presented in Fig. 10. Hematite nanoparticles <10 nm in size were produced
from ferric nitrate and ferric ammonium sulfate solutions using this method [237]. Magnetite
nanoparticles were also produced using ferrous sulfate, with the addition of urea [238]. It was
noted that the particle size and morphology were strongly dependent on operating temperature
and processing time. The particle size increased as the operating temperature increased and
different forms of iron oxide were produced from the same precursor by changing the
temperature. For instance, mixtures of 6-line ferrihydrite (a weakly crystalline form of iron
oxide) and hematite were formed at 250e350  C, while pure hematite was obtained at
temperatures >350  C.
Another flow technique was developed by Adschiri and Arai [239,240] and employed rapid
and intimate mixing of the water stream with the precursor solution in a T-mixer. High
supersaturation rates were achieved in the T-mixer because of the low solubility of metal
hydroxides in supercritical water [230]. They were able to produce hematite, magnetite, and
other oxides in the size range of 10 nm to 1 mm at 400e490  C and 30e40 MPa [228,229,241].
The residence time of the precursor solutions was about 2 min in most of their experiments.
Hydrolysis of metal acetates in supercritical water was studied by Poliakoff et al. [224,233,242]
who produced many oxides, including magnetite, in the size range 3e105 nm at 200e400  C
and 25 MPa. Uniform particles of hematite were produced by Teja et al. [230,231] using two
variations of the continuous hydrothermal technique. Adschiri et al. [240,243] successfully
synthesized 50 nm hematite and magnetite nanoparticles without the addition of a strong base
using the continuous technique. Ferric ammonium sulfate, ferric nitrate, ferric sulfate, and
ferrous chloride were used for the synthesis of hematite, while ferric ammonium citrate was
used as a precursor for magnetite. In the magnetite experiments, Fe(III) was reduced to Fe(II)
by carbon monoxide produced via the thermal decomposition of ammonium citrate. Since CO

Fig. 10. Schematic diagram of the apparatus of Matson et al. [236].


A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 37
55 (2009) 22e45
gas is miscible in supercritical water, it provided a uniform reducing atmosphere throughout the
reactor. This suggests that the final product form can be controlled by the introduction of an
oxidizing or reducing gas during continuous hydrothermal synthesis at supercritical conditions
[240].
Hao and Teja [225] conducted a detailed investigation of the effects of precursor
concentration, temperature, and residence time on particle size and morphology in this
method. The particle size and size distribution increased with precursor concentration, as
shown in (TEM) images in Fig. 11. However, the residence time had a more significant impact
on the average particle size than feed concentration. Monodisperse particles were produced at
short residence times. Teja et al. [244] have also employed the continuous hydrothermal
method to produce polyvinyl alcohol (PVA) coated iron oxide nanoparticles. PVA was chosen
as the coating material because it has the desired solution properties in water and it contains
many isolated hydroxyl functional groups, which can adsorb and complex with metal ions.
The results of their experiment showed that particles with uniform shape and narrow particle
size distribution were obtained in the presence of PVA. The average particle size decreased
with increasing PVA concentration when the residence time was of the order of 2 s, and
became nearly independent of PVA concentration when the residence time was 10 s or
higher [245].
Recently, Lester et al. [246] reported a new design that employs a nozzle mixer to exploit
differences in the densities of supercritical water and precursor solutions to improve mixing
inside the reactor. A diagram of their reactor is shown in Fig. 12. Supercritical water is
introduced into the reactor from the top and the precursor metal salt stream from the bottom.
Mixing of the two streams is almost instantaneous, and makes use of buoyancy induced eddies
to produce ‘‘ideal’’ mixing conditions. This leads to very low residence times and limits
subsequent particle growth. Lester et al. synthesized many metal oxides in the size range of
6e64 nm in their system, thereby demonstrating the effective separation of nucleation and
growth steps in continuous hydrothermal processing.
The continuous hydrothermal technique offers many opportunities for controlling particle
size and morphology by keeping residence times low and mixing processes efficient. It is also
easy to scale up. However, engineering of particle surfaces cannot be accomplished in situ and
requires additional post-processing steps.

Fig. 11. TEM images of hematite nanoparticles obtained via continuous processing at 300  C in 12 s, with a precursor
concentration of 0.03 M (right) and 0.50 M (left).
38 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45

Fig. 12. Schematic of the nozzle mixer (adapted from ref. [246]).

5. Summary

Substantial progress has been made in the synthesis of monodisperse magnetic nano-
particles for applications in nanotechnology and biotechnology. Methods have been devel-
oped that offer control over the size, size distribution, shape, crystal structure, defect
distribution and surface structure of nanoparticles and their magnetic properties. Among the
methods reviewed, continuous supercritical hydrothermal synthesis probably offers the
most promise for process control and scalability. Furthermore, this method is environ-
mentally benign and does not involve the use of toxic solvents or surfactants. Surface
treatment of nanoparticles, however, will require additional steps in this method. A major
challenge for all methods is the design of magnetic nanoparticles with effective surface
coatings that provide optimum performance in vitro and in vivo biological applications.
Additional challenges include scale-up, toxicity, and safety of large-scale particle production
processes.
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 39
55 (2009) 22e45
References

[1] R.M. Cornell, U. Schwertmann, The Iron Oxides: Structure, Properties, Reactions, Occurrences and Uses, second
ed. Wiley-VCH, Weinheim, 2003.
[2] P. Majewski, B. Thierry, Critical Reviews in Solid State and Material Sciences 32 (3-4) (2007) 203e215.
[3] P. Majewski, Maghemite, Encyclopaedia Britannica Online, http://search.eb.com/eb/article-9049986, 2008.
[4] S. Klotz, G. Steinle-Neumann, T. Strassle, J. Philippe, T. Hansen, M.J. Wenzel, Physical Review B 77 (1) (2008).
[5] K. Haneda, A.H. Morrish, Solid State Communications 22 (12) (1977) 779e782.
[6] F.J. Morin, Physical Review 78 (6) (1950) 819e820.
[7] R.D. Zysler, D. Fiorani, A.M. Testa, Journal of Magnetism and Magnetic Materials 224 (1) (2001) 5e11.
[8] P.J. Besser, A.H. Morrish, C.W. Searle, Physical Review 153 (2) (1967) 632e640.
[9] T.P. Raming, A.J.A. Winnubst, C.M. van Kats, A.P. Philipse, Journal of Colloid and Interface Science 249 (2)
(2002) 346e350.
[10] F. Bødker, S. Mørup, Europhysics Letters 52 (2) (2000) 217e223.
[11] N. Amin, S. Arajs, Physical Review B 35 (10) (1987) 4810.
[12] R.C. Nininger, D. Schroeer, Journal of Physics and Chemistry of Solids 39 (2) (1978) 137e144.
[13] E. Murad, in: J.W. Stucki, B.A. Goodman, U. Schwertmann (Eds.), Series C: Mathematical and Physical
Sciences. D. Reidel, 1985.
[14] E. Svab, E. Kren, Journal of Magnetism and Magnetic Materials 14 (2-3) (1979) 184e186.
[15] E. De Grave, L.H. Bowen, S.B. Weed, Journal of Magnetism and Magnetic Materials 27 (1) (1982) 98e108.
[16] R. Vandenberghe, A. Verbeeck, E. De Grave, W. Stiers, Hyperfine Interactions 29 (1) (1986) 1157e1160.
[17] A.H. Morrish, G.B. Johnston, N.A. Curry, Physics Letters 7 (3) (1963) 177e178.
[18] A.H. Morrish, J.A. Eaton, Journal of Applied Physics 42 (4) (1971) 1495e1496.
[19] J.M.D. Coey, G.A. Sawatzky, Journal of Physics C: Solid State Physics 4 (15) (1971) 2386e2407.
[20] S.P. Sena, R.A. Lindley, H.J. Blythe, C. Sauer, M. Al-Kafarji, G.A. Gehring, Journal of Magnetism and Magnetic
Materials 176 (2-3) (1997) 111e126.
[21] D.T. Margulies, F.T. Parker, F.E. Spada, R.S. Goldman, J. Li, R. Sinclair, A.E. Berkowitz, Physical Review B 53
(14) (1996) 9175e9187.
[22] F.C. Voogt, T.T.M. Palstra, L. Niesen, O.C. Rogojanu, M.A. James, T. Hibma, Physical Review B 57 (14) (1998)
R8107eR8110.
[23] T. Kado, Journal of Applied Physics 103 (4) (2008) 043902-1e043902-4.
[24] M.G. Chapline, S.X. Wang, Journal of Applied Physics 97 (12) (2005) 123901-1e123901-3.
[25] H. Qiu, L. Pan, L. Li, H. Zhu, X. Zhao, M. Xu, L. Qin, J.Q. Xiao, Journal of Applied Physics 102 (11) (2007)
113913-1e113913-5.
[26] D.T. Margulies, F.T. Parker, M.L. Rudee, F.E. Spada, J.N. Chapman, P.R. Aitchison, A.E. Berkowitz, Physical
Review Letters 79 (25) (1997) 5162e5165.
[27] U. Meisen, H. Kathrein, Journal of Imaging Science and Technology 44 (6) (2000) 508e513.
[28] R. Dronskowski, Advanced Functional Materials 11 (1) (2001) 27e29.
[29] O. Helgason, J.-M. Greneche, F.J. Berry, F. Mosselmans, Journal of Physics: Condensed Matter 15 (17) (2003)
2907e2915.
[30] F.J. Berry, C. Greaves, O. Helgason, J. McManus, Journal of Materials Chemistry 9 (1) (1999)
223e226.
[31] T. Neuberger, B. Schopf, H. Hofmann, M. Hofmann, B. von Rechenberg, Journal of Magnetism and Magnetic
Materials 293 (1) (2005) 483e496.
[32] E. Tronc, A. Ezzir, R. Cherkaoui, C. Chaneac, M. Nogues, H. Kachkachi, D. Fiorani, A.M. Testa, J.M. Greneche,
J.P. Jolivet, Journal of Magnetism and Magnetic Materials 221 (1e2) (2000) 63e79.
[33] P.V. Hendriksen, S. Linderoth, P.A. Lindgård, Physical Review B 48 (10) (1993) 7259.
[34] M.P. Morales, C.J. Serna, F. Bødker, S. Mørup, Journal of Physics: Condensed Matter 9 (25) (1997)
5461e5467.
[35] F.E. Spada, F.T. Parker, C.Y. Nakakura, A.E. Berkowitz, Journal of Magnetism and Magnetic Materials 120
(1e3) (1993) 129e135.
[36] F. Itoh, M. Satou, Y. Yamazaki, IEEE Transactions on Magnetics 13 (5) (1977) 1385e1387.
[37] F.E. Spada, A.E. Berkowitz, N.T. Prokey, Journal of Applied Physics 69 (8) (1991) 4475e4477.
[38] F. Itoh, M. Satou, Japanese Journal of Applied Physics 14 (12) (1975) 2091e2092.
[39] M. Azhar Uddin, H. Tsuda, S. Wu, E. Sasaoka, Fuel 87 (4e5) (2008) 451e459.
40 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
[40] X.J. Cui, M. Antonietti, S.H. Yu, Small 2006;2(6):756e759.
[41] T.J. Daou, G. Pourroy, S. Begin-Colin, J.M. Greneche, C. Ulhaq-Bouillet, P. Legare, P. Bernhardt, C. Leuvrey,
G. Rogez, Chemistry of Materials 18 (18) (2006) 4399e4404.
[42] F. dos Santos Coelho, J.D. Ardisson, F.C.C. Moura, R.M. Lago, E. Murad, J.D. Fabris, Chemosphere 71 (1)
(2008) 90e96.
[43] S.Y. Lian, E. Wang, Z.H. Kang, Y.P. Bai, L. Gao, M. Jiang, C.W. Hu, L. Xu, Solid State Communications 129 (8)
(2004) 485e490.
[44] S. Takami, T. Sato, T. Mousavand, S. Ohara, M. Umetsu, T. Adschiri, Materials Letters 61 (26) (2007)
4769e4772.
[45] P. Tartaj, M.D. Morales, S. Veintemillas-Verdaguer, T. Gonzalez-Carreno, C.J. Serna, Journal of Physics D:
Applied Physics 36 (13) (2003) R182eR197.
[46] A.S. Teja, L.J. Holm, in: Y.-P. Sun (Ed.), Supercritical Fluid Technology in Materials Science and Engineering:
Synthesis, Properties, and Applications, Elsevier, 2002, pp. 327e349.
[47] S.B. Wang, Y.L. Min, S.H. Yu, Journal of Physical Chemistry C 111 (9) (2007) 3551e3554.
[48] C. Xu, A.S. Teja, Journal of Supercritical Fluids 39 (1) (2006) 135e141.
[49] M. Hermanek, R. Zboril, N. Medrik, J. Pechousek, C. Gregor, Journal of the American Chemical Society 129
(35) (2007) 10929e10936.
[50] R. Jurgons, C. Seliger, A. Hilpert, L. Trahms, S. Odenbach, C. Alexiou, Journal of Physics: Condensed Matter 18
(38) (2006) S2893eS2902.
[51] Y. Qiang, J. Antony, A. Sharma, J. Nutting, D. Sikes, D. Meyer, Journal of Nanoparticle Research 8 (3e4) (2006)
489e496.
[52] Z.P. Xu, Q.H. Zeng, G.Q. Lu, A.B. Yu, Chemical Engineering Science 61 (3) (2006) 1027e1040.
[53] T. Hyeon, Chemical Communications (8) (2003) 927e934.
[54] L. Cabrera, S. Gutierrez, N. Menendez, M.P. Morales, P. Herrasti, Electrochimica Acta 53 (8) (2008) 3436e3441.
[55] C. Li, Y. Shen, M. Jia, S. Sheng, M.O. Adebajo, H. Zhu, Catalysis Communications 9 (3) (2008) 355e361.
[56] H. Hibst, E. Schwab, in: R.W. Cahn (Ed.), Electronic and Magnetic Properties of Metals and Ceramics, VCH,
Weinheim, 1994.
[57] M.P. Sharrock, R.E. Bodnar, Journal of Applied Physics 57 (8) (1985) 3919e3924.
[58] R.E. Camley, R.L. Stamps, Journal of Physics: Condensed Matter 5 (23) (1993) 3727e3786.
[59] K. Raj, R. Moskowitz, Journal of Magnetism and Magnetic Materials 85 (1e3) (1990) 233e245.
[60] A.S. Lubbe, C. Bergemann, J. Brock, D.G. McClure, Journal of Magnetism and Magnetic Materials 194 (1e3)
(1999) 149e155.
[61] Y.H. Zheng, Y. Cheng, F. Bao, Y.S. Wang, Materials Research Bulletin 41 (3) (2006) 525e529.
[62] H. Ai, C. Flask, B. Weinberg, X. Shuai, M.D. Pagel, D. Farrell, J. Duerk, J.M. Gao, Advanced Materials 17 (16)
(2005) 1949e1952.
[63] A.D. Arelaro, A.L. Brandl, E. Lima, L.F. Gamarra, G.E.S. Brito, W.M. Pontuschka, G.F. Goya, Journal of
Applied Physics 97 (10) (2005) 10J316-1e10J316-3.
[64] M. Gonzales, K.M. Krishnan, Journal of Magnetism and Magnetic Materials 293 (1) (2005) 265e270.
[65] H. Lee, H.P. Shao, Y.Q. Huang, B. Kwak, IEEE Transactions on Magnetics 41 (10) (2005) 4102e4104.
[66] N. Sadeghiani, L.S. Barbosa, M.H.A. Guedes, S.B. Chaves, J.G. Santos, O. Silva, F. Pelegrini, R.B. Azevedo,
P.C. Morais, Z.G.M. Lacava, IEEE Transactions on Magnetics 41 (10) (2005) 4108e4110.
[67] A.S. Lubbe, C. Bergemann, H. Riess, F. Schriever, P. Reichardt, K. Possinger, M. Matthias, B. Dorken,
F. Herrmann, R. Gurtler, P. Hohenberger, N. Haas, R. Sohr, B. Sander, A.-J. Lemke, D. Ohlendorf, W. Huhnt,
D. Huhn, Cancer Research 56 (20) (1996) 4686e4693.
[68] N. Kohler, C. Sun, J. Wang, M. Zhang, Langmuir 21 (19) (2005) 8858e8864.
[69] J.S. Kim, T.J. Yoon, B.G. Kim, S.J. Park, H.W. Kim, K.H. Lee, S.B. Park, J.K. Lee, M.H. Cho, Toxicological
Sciences 89 (1) (2006) 338e347.
[70] P. Tartaj, M.P. Morales, T. Gonzalez-Carreno, S. Veintemillas-Verdaguer, C.J. Serna, Journal of Magnetism and
Magnetic Materials 290 (2005) 28e34.
[71] H.H. Yang, S.Q. Zhang, X.L. Chen, Z.X. Zhuang, J.G. Xu, X.R. Wang, Analytical Chemistry 76 (5) (2004)
1316e1321.
[72] S.X. Wang, S.-Y. Bae, G. Li, S. Sun, R.L. White, J.T. Kemp, C.D. Webb, Journal of Magnetism and Magnetic
Materials 293 (1) (2005) 731e736.
[73] C. Duanmu, I. Saha, Y. Zheng, B.M. Goodson, Y. Gao, Chemistry of Materials 18 (25) (2006) 5973e5981.
[74] L. Kaufner, R. Cartier, R. Wustneck, I. Fichtner, S. Pietschmann, H. Bruhn, D. Schutt, A.F. Thunemann, U. Pison,
Nanotechnology 18 (11) (2007) 115710.
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 41
55 (2009) 22e45
[75] D.K. Kim, M. Mikhaylova, F.H. Wang, J. Kehr, B. Bjelke, Y. Zhang, T. Tsakalakos, M. Muhammed, Chemistry
of Materials 15 (23) (2003) 4343e4351.
[76] H.Y. Lee, N.H. Lim, J.A. Seo, S.H. Yuk, B.K. Kwak, G. Khang, H.B. Lee, S.H. Cho, Journal of Biomedical
Materials Research Part B: Applied Biomaterials 79B (1) (2006) 142e150.
[77] A.F. Thunemann, D. Schutt, L. Kaufner, U. Pison, H. Mohwald, Langmuir 22 (5) (2006) 2351e2357.
[78] J. Giri, P. Pradhan, V. Somani, H. Chelawat, S. Chhatre, R. Banerjee, D. Bahadur, Journal of Magnetism and
Magnetic Materials 320 (5) (2008) 724e730.
[79] Y.W. Jun, J.S. Choi, J. Cheon, Chemical Communications (12) (2007) 1203e1214.
[80] D. Pouliquen, R. Perdrisot, A. Ermias, S. Akoka, P. Jallet, J.J. Le Jeune, Magnetic Resonance Imaging 7 (6)
(1989) 619e627.
[81] D. Pouliquen, J.J. Le Jeune, R. Perdrisot, A. Ermias, P. Jallet, Magnetic Resonance Imaging 9 (3) (1991)
275e283.
[82] M. Taupitz, J. Schnorr, C. Abramjuk, S. Wagner, H. Pilgrimm, H. Hünigen, B. Hamm, Journal of Magnetic
Resonance Imaging 12 (6) (2000) 905e911.
[83] L. Babes, B. Denizot, G. Tanguy, J.J. Le Jeune, P. Jallet, Journal of Colloid and Interface Science 212 (2) (1999)
474e482.
[84] H.T. Song, J.S. Choi, Y.M. Huh, S. Kim, Y.w Jun, J.S. Suh, J. Cheon, Journal of the American Chemical Society
127 (28) (2005) 9992e9993.
[85] S. Sieben, C. Bergemann, A. Lubbe, B. Brockmann, D. Rescheleit, Journal of Magnetism and Magnetic Materials
225 (1e2) (2001) 175e179.
[86] A. Petri-Fink, M. Chastellain, L. Juillerat-Jeanneret, A. Ferrari, H. Hofmann, Biomaterials 26 (15) (2005)
2685e2694.
[87] D. Tanyolaç, A.R. Özdural, Journal of Applied Polymer Science 80 (5) (2001) 707e715.
[88] B.H. Pyle, S.C. Broadaway, G.A. McFeters, Applied and Environmental Microbiology 65 (5) (1999) 1966e1972.
[89] I. Avital, D. Inderbitzin, T. Aoki, D.B. Tyan, A.H. Cohen, C. Ferraresso, J. Rozga, W.S. Arnaout,
A.A. Demetriou, Biochemical and Biophysical Research Communications 288 (1) (2001) 156e164.
[90] H. Lee, E. Lee, D.K. Kim, N.K. Jang, Y.Y. Jeong, S. Jon, Journal of the American Chemical Society 128 (22)
(2006) 7383e7389.
[91] L. Cumbal, J. Greenleaf, D. Leun, A.K. SenGupta, Reactive and Functional Polymers 54 (1e3) (2003) 167e180.
[92] J.L. Zhang, R.S. Srivastava, R.D.K. Misra, Langmuir 23 (11) (2007) 6342e6351.
[93] M. Kumagai, Y. Imai, T. Nakamura, Y. Yamasaki, M. Sekino, S. Ueno, K. Hanaoka, K. Kikuchi, T. Nagano,
E. Kaneko, K. Shimokado, K. Kataoka, Colloids and Surfaces B: Biointerfaces 56 (1e2) (2007) 174e181.
[94] F. Shi, M.K. Tse, M.M. Pohl, A. Bruckner, S.M. Zhang, M. Beller, Angewandte Chemie International Edition 46
(46) (2007) 8866e8868.
[95] R.J. Zhang, J.J. Huang, H.T. Zhao, Z.Q. Sun, Y. Wang, Energy & Fuels 21 (5) (2007) 2682e2687.
[96] C.T. Wang, R.J. Willey, Journal of Non-Crystalline Solids 225 (1) (1998) 173e177.
[97] S. Al-Sayari, A.F. Carley, S.H. Taylor, G.J. Hutchings, Topics in Catalysis 44 (1e2) (2007) 123e128.
[98] Y. Wang, B.H. Davis, Applied Catalysis A: General 180 (1e2) (1999) 277e285.
[99] F.M. Bautista, J.M. Campelo, D. Luna, J.M. Marinas, R.A. Quiros, A.A. Romero, Applied Catalysis B: Envi-
ronmental 70 (1e4) (2007) 611e620.
[100] Z. Zhong, J. Ho, J. Teo, S. Shen, A. Gedanken, Chemistry of Materials 19 (19) (2007) 4776e4782.
[101] A.K. Kandalam, B. Chatterjee, S.N. Khanna, B.K. Rao, P. Jena, B.V. Reddy, Surface Science 601 (21) (2007)
4873e4880.
[102] Q. Liu, Z.M. Cui, Z. Ma, S.W. Bian, W.G. Song, L.J. Wan, Nanotechnology 18 (38) (2007).
[103] T.L. Jorgensen, H. Livbjerg, P. Glarborg, Chemical Engineering Science 62 (16) (2007) 4496e4499.
[104] M.L. Peterson, J.G.E. Brown, G.A. Parks, C.L. Stein, Geochimica et Cosmochimica Acta 61 (16) (1997)
3399e3412.
[105] Z. Zhong, J. Lin, S.P. Teh, J. Teo, F.M. Dautzenberg, Advanced Functional Materials 17 (8) (2007) 1402e1408.
[106] A.P. Kozlova, S. Sugiyama, A.I. Kozlov, K. Asakura, Y. Iwasawa, Journal of Catalysis 176 (2) (1998) 426e438.
[107] G.J. Hutchings, M.S. Hall, A.F. Carley, P. Landon, B.E. Solsona, C.J. Kiely, A. Herzing, M. Makkee,
J.A. Moulijn, A. Overweg, J.C. Fierro-Gonzalez, J. Guzman, B.C. Gates, Journal of Catalysis 242 (1) (2006)
71e81.
[108] J. Bandara, U. Klehm, J. Kiwi, Applied Catalysis B: Environmental 76 (1e2) (2007) 73e81.
[109] M. Brebu, M.A. Uddin, A. Muto, Y. Sakata, C. Vasile, Energy & Fuels 15 (3) (2001) 559e564.
[110] U.T. Lam, R. Mammucari, K. Suzuki, N.R. Foster, Industrial & Engineering Chemistry Research 47 (3) (2008)
599e614.
42 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
[111] H. Gaedcke, in: G. Buxbaum (Ed.), Industrial Inorganic Pigments, VCH, Weinheim, 1998, pp. 224e228.
[112] U. Jeong, X.W. Teng, Y. Wang, H. Yang, Y.N. Xia, Advanced Materials 19 (1) (2007) 33e60.
[113] M. Pregelj, P. Umek, B. Drolc, B. Jancar, Z. Jaglicic, R. Dominko, D. Arcon, Journal of Materials Research 21
(11) (2006) 2955e2962.
[114] A. Tavakoli, M. Sohrabi, A. Kargari, Chemical Papers 61 (3) (2007) 151e170.
[115] H.L. Zhu, D.R. Yang, L.M. Zhu, Surface & Coatings Technology 201 (12) (2007) 5870e5874.
[116] H.O. Pierson, Handbook of Chemical Vapor Deposition: Principles, Technology, and Applications, William
Andrew Inc, 1999.
[117] W. Chang, G. Skandan, S.C. Danforth, B.H. Kear, H. Hahn, Nanostructured Materials 4 (5) (1994) 507e520.
[118] C. Powell, J. Oxley, J. Blocher, Vapor Deposition, John Wiley & Sons, New York, 1966.
[119] W. Chang, G. Skandan, H. Hahn, S.C. Danforth, B.H. Kear, Nanostructured Materials 4 (3) (1994) 345e351.
[120] W. Kern, V. Bam, in: J. Vossen, W. Kern (Eds.), Thin Film Processes, Academic Press, New York, 1978.
[121] T. Maruyama, T. Kanagawa, Journal of the Electrochemical Society 143 (5) (1996) 1675e1677.
[122] K. Shalini, G.N. Subbanna, S. Chandrasekaran, S.A. Shivashankar, Thin Solid Films 424 (1) (2003) 56e60.
[123] S. Park, S. Lim, H. Choi, Chemistry of Materials 18 (22) (2006) 5150e5152.
[124] S. Dhara, B.R. Awasthy, A.C. Rastogi, B.K. Das, N.V. Gelfond, N.E. Fedotova, A.F. Bykov, I.K. Igumenov,
Journal of Magnetism and Magnetic Materials 134 (1) (1994) 29e33.
[125] E. Fujii, H. Torii, A. Tomozawa, R. Takayama, T. Hirao, Journal of Crystal Growth 151 (1-2) (1995) 134e139.
[126] S. Martelli, A. Mancini, R. Giorgi, R. Alexandrescu, S. Cojocaru, A. Crunteanu, I. Voicu, M. Balu, I. Morjan,
Applied Surface Science 154-155 (2000) 353e359.
[127] R. Alexandrescu, I. Morjann, A. Crunteanu, S. Cojocaru, S. Petcu, V. Teodorescu, F. Huisken, B. Koh,
M. Ehbrecht, Materials Chemistry and Physics 55 (2) (1998) 115e121.
[128] I. Morjan, R. Alexandrescu, I. Soare, F. Dumitrache, I. Sandu, I. Voicu, A. Crunteanu, E. Vasile, V. Ciupina,
S. Martelli, Materials Science and Engineering: C 23 (1e2) (2003) 211e216.
[129] F. Dumitrache, I. Morjan, R. Alexandrescu, V. Ciupina, G. Prodan, I. Voicu, C. Fleaca, L. Albu, M. Savoiu,
I. Sandu, E. Popovici, I. Soare, Applied Surface Science 247 (1e4) (2005) 25e31.
[130] S. Veintemillas-Verdaguer, M.P. Morales, C.J. Serna, Materials Letters 35 (3e4) (1998) 227e231.
[131] O. Bomati-Miguel, L. Mazeina, A. Navrotsky, S. Veintemillas-Verdaguer, Chemistry of Materials 20 (2) (2008)
591e598.
[132] R. Alexandrescu, I. Morjan, I. Voicu, F. Dumitrache, L. Albu, I. Soare, G. Prodan, Applied Surface Science 248
(1e4) (2005) 138e146.
[133] M.P. Morales, O. Bomati-Miguel, R. Perez de Alejo, J. Ruiz-Cabello, S. Veintemillas-Verdaguer, K. O’Grady,
Journal of Magnetism and Magnetic Materials 266 (1e2) (2003) 102e109.
[134] M.P. Morales, S. Veintemillas-Verdaguer, C.J. Serna, Journal of Materials Research 14 (7) (1999) 3066e3072.
[135] S. Veintemillas-Verdaguer, O. Bomati-Miguel, M.P. Morales, Scripta Materialia 47 (9) (2002) 589e593.
[136] T. Sugimoto, E. Matijevic, Journal of Colloid and Interface Science 74 (1) (1980) 227e243.
[137] A.A. Khaleel, Chemistry A European Journal 10 (4) (2004) 925e932.
[138] A. Chastellain, A. Petri, H. Hofmann, Journal of Colloid and Interface Science 278 (2) (2004) 353e360.
[139] Y. Tamaura, K. Ito, T. Katsura, Journal of the Chemical Society Dalton Transactions (2) (1983) 189e194.
[140] E. Tronc, P. Belleville, J.P. Jolivet, J. Livage, Langmuir 8 (1) (1992) 313e319.
[141] J.-P. Jolivet, E. Tronc, Journal of Colloid and Interface Science 25 (2) (1988) 688e701.
[142] Z.L. Liu, H.B. Wang, Q.H. Lu, G.H. Du, L. Peng, Y.Q. Du, S.M. Zhang, K.L. Yao, Journal of Magnetism and
Magnetic Materials 283 (2e3) (2004) 258e262.
[143] D.K. Kim, Y. Zhang, W. Voit, K.V. Rao, M. Muhammed, Journal of Magnetism and Magnetic Materials 225 (1e
2) (2001) 30e36.
[144] D.K. Kim, M. Mikhaylova, Y. Zhang, M. Muhammed, Chemistry of Materials 15 (8) (2003) 1617e1627.
[145] C.L. Lin, C.F. Lee, W.Y. Chiu, Journal of Colloid and Interface Science 291 (2) (2005) 411e420.
[146] S. Mohapatra, N. Pramanik, S. Mukherjee, S.K. Ghosh, P. Pramanik, Journal of Materials Science 42 (17) (2007)
7566e7574.
[147] R.F. Soares, C.A.P. Leite, W. Botter, F. Galembeck, Journal of Applied Polymer Science 60 (11) (1996)
2001e2006.
[148] M. Mikhaylova, D.K. Kim, N. Bobrysheva, M. Osmolowsky, V. Semenov, T. Tsakalakos, M. Muhammed,
Langmuir 20 (6) (2004) 2472e2477.
[149] S.E. Khalafalla, G.W. Reimers, IEEE Transactions on Magnetics 16 (2) (1980) 178e183.
[150] A. Wooding, M. Kilner, D.B. Lambrick, Journal of Colloid and Interface Science 144 (1) (1991) 236e242.
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 43
55 (2009) 22e45
[151] S.G. Kwon, Y. Piao, J. Park, S. Angappane, Y. Jo, N.M. Hwang, J.G. Park, T. Hyeon, Journal of the American
Chemical Society 129 (41) (2007) 12571e12584.
[152] S. Sun, H. Zeng, D.B. Robinson, S. Raoux, P.M. Rice, S.X. Wang, G. Li, Journal of the American Chemical
Society 126 (1) (2004) 273e279.
[153] A.G. Roca, M.P. Morales, K. O’Grady, C.J. Serna, Nanotechnology 17 (11) (2006) 2783e2788.
[154] A.G. Roca, M.P. Morales, C.J. Serna, IEEE Transactions on Magnetics 42 (10) (2006) 3025e3029.
[155] L. Zhang, R. He, H.C. Gu, Applied Surface Science 253 (5) (2006) 2611e2617.
[156] L.M. Bronstein, X.L. Huang, J. Retrum, A. Schmucker, M. Pink, B.D. Stein, B. Dragnea, Chemistry of Materials
19 (15) (2007) 3624e3632.
[157] H. Jung, H. Park, J. Kim, J.H. Lee, H.G. Hur, N.V. Myung, H. Choi, Environmental Science &Technology 41 (13)
(2007) 4741e4747.
[158] K. Muraishi, T. Takano, K. Nagase, N. Tanaka, Journal of Inorganic and Nuclear Chemistry 43 (10) (1981)
2293e2297.
[159] M.M. Rahman, V.A. Mukhedkar, A. Venkataraman, A.K. Nikumbh, S.B. Kulkarni, A.J. Mukhedkar, Thermo-
chimica Acta 125 (1988) 173e190.
[160] A.K. Nikumbh, A.A. Latkar, M.M. Phadke, Thermochimica Acta 219 (1993) 269e282.
[161] S. Music, M. Gotic, S. Popovic, I. Czako-Nagy, Materials Letters 20 (3-4) (1994) 143e148.
[162] J. Rockenberger, E.C. Scher, A.P. Alivisatos, Journal of the American Chemical Society 121 (49) (1999)
11595e11596.
[163] C. Amiens, B. Chaudret, Modern Physics Letters B 21 (18) (2007) 1133e1141.
[164] S.H. Sun, H. Zeng, Journal of the American Chemical Society 124 (28) (2002) 8204e8205.
[165] Z. Li, L. Wei, M.Y. Gao, H. Lei, Advanced Materials 17 (8) (2005) 1001e1005.
[166] T. Hyeon, S.S. Lee, J. Park, Y. Chung, H.B. Na, Journal of the American Chemical Society 123 (51) (2001)
12798e12801.
[167] N. Pinna, S. Grancharov, P. Beato, P. Bonville, M. Antonietti, M. Niederberger, Chemistry of Materials 17 (11)
(2005) 3044e3049.
[168] Y.X. Pang, X.J. Bao, Journal of Materials Chemistry 12 (12) (2002) 3699e3704.
[169] V. Pillai, P. Kumar, M.J. Hou, P. Ayyub, D.O. Shah, Advances in Colloid and Interface Science 55 (1995)
241e269.
[170] P. Tartaj, L.C. De Jonghe, Journal of Materials Chemistry 10 (12) (2000) 2786e2790.
[171] I. Capek, Advances in Colloid and Interface Science 110 (1e2) (2004) 49e74.
[172] V. Chhabra, P. Ayyub, S. Chattopadhyay, A.N. Maitra, Materials Letters 26 (1e2) (1996) 21e26.
[173] A.B. Chin, I.I. Yaacob, Journal of Materials Processing Technology 191 (1e3) (2007) 235e237.
[174] N. Nassar, M. Husein, Physica Status Solidi A: Applications and Materials Science 203 (6) (2006) 1324e1328.
[175] J. Vidal-Vidal, J. Rivas, M.A. Lopez-Quintela, Colloids and Surfaces A: Physicochemical and Engineering
Aspects 288 (1e3) (2006) 44e51.
[176] J.A.L. Perez, M.A.L. Quintela, J. Mira, J. Rivas, S.W. Charles, Journal of Physical Chemistry B 101 (41) (1997)
8045e8047.
[177] Y. Deng, L. Wang, W. Yang, S. Fu, A. Elaissari, Journal of Magnetism and Magnetic Materials 257 (1) (2003)
69e78.
[178] P.A. Dresco, V.S. Zaitsev, R.J. Gambino, B. Chu, Langmuir 15 (6) (1999) 1945e1951.
[179] J. Zhi, Y.J. Wang, Y.C. Lu, J.Y. Ma, G.S. Luo, Reactive & Functional Polymers 66 (12) (2006) 1552e1558.
[180] Z.Z. Xu, C.C. Wang, W.L. Yang, Y.H. Deng, S.K. Fu, Journal of Magnetism and Magnetic Materials 277 (1e2)
(2004) 136e143.
[181] P. Tartaj, C.J. Serna, Chemistry of Materials 14 (10) (2002) 4396e4402.
[182] S. Santra, R. Tapec, N. Theodoropoulou, J. Dobson, A. Hebard, W.H. Tan, Langmuir 17 (10) (2001) 2900e2906.
[183] M. Zhang, B.L. Cushing, C.J. O’Connor, Nanotechnology 19 (8) (2008) 085601 (5pp).
[184] A.K. Ganguli, T. Ahmad, Journal of Nanoscience and Nanotechnology 7 (6) (2007) 2029e2035.
[185] J. Esquivel, I.A. Facundo, M.E. Trevino, R.G. Lopez, Journal of Materials Science 42 (21) (2007) 9015e9020.
[186] T. Herranz, S. Rojas, M. Ojeda, F.J. Perez-Alonso, P. Tefferos, K. Pirota, J.L.G. Fierro, Chemistry of Materials 18
(9) (2006) 2364e2375.
[187] C. Larpent, H. Patin, Journal of Molecular Catalysis 72 (3) (1992) 315e329.
[188] P. Tartaj, C.J. Serna, Journal of the American Chemical Society 125 (51) (2003) 15754e15755.
[189] M. Tadic, D. Markovic, V. Spasojevic, V. Kusigerski, M. Remskar, J. Pirnat, Z. Jaglicic, Journal of Alloys and
Compounds 441 (1e2) (2007) 291e296.
[190] Z.Z. Xu, C.C. Wang, W.L. Yang, S.K. Fu, Journal of Materials Science 40 (17) (2005) 4667e4669.
44 A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials
55 (2009) 22e45
[191] Y.H. Deng, C.C. Wang, J.H. Hu, W.L. Yang, S.K. Fu, Colloids and Surfaces A: Physicochemical and Engineering
Aspects 262 (1e3) (2005) 87e93.
[192] C.-T. Wang, S.-H. Ro, Applied Catalysis A: General 285 (1e2) (2005) 196e204.
[193] G. Ennas, A. Musinu, G. Piccaluga, D. Zedda, D. Gatteschi, C. Sangregorio, J.L. Stanger, G. Concas, G. Spano,
Chemistry of Materials 10 (2) (1998) 495e502.
[194] S. Bruni, F. Cariati, M. Casu, A. Lai, A. Musinu, G. Piccaluga, S. Solinas, Nanostructured Materials 11 (5) (1999)
573e586.
[195] B. Heinrichs, L. Rebbouh, J.W. Geus, S. Lambert, H.C.L. Abbenhuis, F. Grandjean, G.J. Long, J.P. Pirard,
R.A. van Santen, Journal of Non-Crystalline Solids 354 (2e9) (2008) 665e672.
[196] M. Soufyani, D. Bourret, A. Sivade, R. Sempere, Journal of Non-Crystalline Solids 145 (1992) 60e62.
[197] L. Casas, A. Roig, E. Rodriguez, E. Molins, J. Tejada, J. Sort, Journal of Non-Crystalline Solids 285 (1e3)
(2001) 37e43.
[198] A. Braileanu, M. Raileanu, M. Crisan, D. Crisan, R. Birjega, V.E. Marinescu, J. Madarasz, G. Pokol, Journal of
Thermal Analysis and Calorimetry 88 (1) (2007) 163e169.
[199] C. Caizer, C. Savii, M. Popovici, Materials Science and Engineering B: Solid State Materials for Advanced
Technology 97 (2) (2003) 129e134.
[200] S.A. Corr, Y.K. Gun’ko, A.P. Douvalis, M. Venkatesan, R.D. Gunning, P.D. Nellist, Journal of Physical
Chemistry C 112 (4) (2008) 1008e1018.
[201] D. Predoi, O. Crisan, A. Jitianu, M.C. Valsangiacom, M. Raileanu, M. Crisan, M. Zaharescu, Thin Solid Films
515 (16) (2007) 6319e6323.
[202] P.P.C. Sartoratto, K.L. Caiado, R.C. Pedroza, S.W. da Silva, P.C. Morais, Journal of Alloys and Compounds 434
(2007) 650e654.
[203] C. Savii, M. Popovici, C. Enache, J. Subrt, D. Niznansky, S. Bakardzieva, C. Caizer, I. Hrianca, Solid State Ionics
151 (1e4) (2002) 219e227.
[204] M. Popovici, M. Gich, A. Roig, L. Casas, E. Molins, C. Savii, D. Becherescu, J. Sort, S. Surinach, J.S. Munoz,
M.D. Baro, J. Nogues, Langmuir 20 (4) (2004) 1425e1429.
[205] H.S.W. Chang, C.C. Chiou, Y.W. Chen, S.R. Sheen, Journal of Solid State Chemistry 128 (1) (1997) 87e92.
[206] B. Orel, M. Macek, F. Svegl, K. Kalcher, Thin Solid Films 246 (1e2) (1994) 131e142.
[207] M. Sedlar, Ceramics International 20 (1) (1994) 73e78.
[208] K.S. Hwang, Y.S. Jeon, K.O. Jeon, B.H. Kim, Optica Applicata 35 (2) (2005) 191e199.
[209] H.R. Liu, Z.L. Liu, Q. Liu, K.L. Yao, Thin Solid Films 500 (1e2) (2006) 105e109.
[210] A.A. Ismail, Applied Catalysis B: Environmental 58 (1e2) (2005) 115e121.
[211] C.D.E. Lakeman, D.A. Payne, Materials Chemistry and Physics 38 (4) (1994) 305e324.
[212] S.Y. An, I.B. Shim, C.S. Kim, Journal of Magnetism and Magnetic Materials 290 (2005) 1551e1554.
[213] R.J. Willey, S.A. Oliver, G. Oliveri, G. Busca, Journal of Materials Research 8 (6) (1993) 1418e1427.
[214] K. Yamaguchi, T. Fujii, S. Kuranouchi, Y. Yamanobe, A. Ueno, IEEE Transactions on Magnetics 25 (5) (1989)
3321e3323.
[215] M. Liu, H. Li, L. Xiao, W. Yu, Y. Lu, Z. Zhao, Journal of Magnetism and Magnetic Materials 294 (3) (2005)
294e297.
[216] C. Burda, X. Chen, R. Narayanan, M.A. El-Sayed, Chemical Reviews 105 (4) (2005) 1025e1102.
[217] R.W. Shaw, T.B. Brill, A.A. Clifford, C.A. Eckert, E.U. Franck, Chemical & Engineering News 69 (51) (1991)
26e39.
[218] K. Sue, K. Kimura, K. Arai, Materials Letters 58 (25) (2004) 3229e3231.
[219] Q.S. Dou, H. Zhang, J.B. Wu, D.R. Yang, Journal of Inorganic Materials 22 (2) (2007) 213e218.
[220] S. Giri, S. Samanta, S. Maji, S. Ganguli, A. Bhaumik, Journal of Magnetism and Magnetic Materials 285 (1e2)
(2005) 296e302.
[221] M. Sorescu, L. Diamandescu, D. Tarabasanu-Mihaila, Journal of Physics and Chemistry of Solids 65 (10) (2004)
1719e1725.
[222] X. Wang, X.Y. Chen, L.S. Gao, H.G. Zheng, M.R. Ji, C.M. Tang, T. Shen, Z.D. Zhang, Journal of Materials
Chemistry 14 (5) (2004) 905e907.
[223] M. Yoshimura, S. Somiya, Materials Chemistry and Physics 61 (1) (1999) 1e8.
[224] A. Cabanas, J.A. Darr, E. Lester, M. Poliakoff, Journal of Materials Chemistry 11 (2) (2001) 561e568.
[225] Y.L. Hao, A.S. Teja, Journal of Materials Research 18 (2) (2003) 415e422.
[226] C.A. Eckert, B.L. Knutson, P.G. Debenedetti, Nature 383 (6598) (1996) 313e318.
[227] K. Sue, M. Suzuki, K. Arai, T. Ohashi, H. Ura, K. Matsui, Y. Hakuta, H. Hayashi, M. Watanabe, T. Hiaki, Green
Chemistry 8 (7) (2006) 634e638.
A.S. Teja, P.-Y. Koh / Progress in Crystal Growth and Characterization of Materials 45
55 (2009) 22e45
[228] T. Adschiri, Y. Hakuta, K. Arai, Industrial & Engineering Chemistry Research 39 (12) (2000) 4901e4907.
[229] T. Adschiri, Y. Hakuta, K. Sue, K. Arai, Journal of Nanoparticle Research 3 (2e3) (2001) 227e235.
[230] L.J. Cote, A.S. Teja, A.P. Wilkinson, Z.J. Zhang, Journal of Materials Research 17 (9) (2002) 2410e2416.
[231] L.J. Cote, A.S. Teja, A.P. Wilkinson, Z.J. Zhang, Fluid Phase Equilibria 210 (2) (2003) 307e317.
[232] K. Sue, Y. Hakuta, R.L. Smith, T. Adschiri, K. Arai, Journal of Chemical and Engineering Data 44 (6) (1999)
1422e1426.
[233] A. Cabanas, M. Poliakoff, Journal of Materials Chemistry 11 (5) (2001) 1408e1416.
[234] C. Xu, Continuous and batch hydrothermal synthesis of metal oxide nanoparticles and metal oxide-activated
carbon nanocomposites, PhD thesis, Georgia Institute of Technolog, GA, 2006.
[235] J.R. Kastner, R. Ganagavaram, P. Kolar, A. Teja, C.B. Xu, Environmental Science & Technology 42 (2) (2008)
556e562.
[236] J. Darab, D. Matson, Journal of Electronic Materials 27 (10) (1998) 1068e1072.
[237] D.W. Matson, J.C. Linehan, R.M. Bean, Materials Letters 14 (4) (1992) 222e226.
[238] D.W. Matson, J.C. Linehan, J.G. Darab, M.F. Buehler, Energy & Fuels 8 (1) (1994) 10e18.
[239] T. Adschiri, K. Kanazawa, K. Arai, Journal of the American Ceramic Society 75 (9) (1992) 2615e2618.
[240] T. Adschiri, K. Kanazawa, K. Arai, Journal of the American Ceramic Society 75 (4) (1992) 1019e1022.
[241] T. Adschiri, Y. Hakuta, K. Kanamura, K. Arai, High Pressure Research 20 (1e6) (2001) 373e384.
[242] A. Cabanas, J.A. Darr, E. Lester, M. Poliakoff, Chemical Communications(11) (2000) 901e902.
[243] T. Adschiri, K. Arai, in: Y.-P. Sun (Ed.), Supercritical Fluid Technology in Materials Science and Engineering:
Synthesis, Properties, and Applications, Elsevier, 2002, pp. 311e325.
[244] C. Xu, A.S. Teja, Journal of Supercritical Fluids 44 (1) (2008) 85e91.
[245] C. Xu, J. Lee, A.S. Teja, Journal of Supercritical Fluids 44 (1) (2008) 92e97.
[246] E. Lester, P. Blood, J. Denyer, D. Giddings, B. Azzopardi, M. Poliakoff, Journal of Supercritical Fluids 37 (2)
(2006) 209e214.

You might also like