You are on page 1of 12

Review

pubs.acs.org/cm

Oleylamine in Nanoparticle Synthesis


Stefanos Mourdikoudis*, and Luis M. Liz-Marzan*,,,

Departamento de Qumica Fsica, Universidade de Vigo, 36310 Vigo, Spain


BioNanoPlasmonics Laboratory, CIC biomaGUNE, Paseo de Miramon 182, 20009 Donostia - San Sebastian, Spain

Ikerbasque, Basque Foundation for Science, 48011 Bilbao, Spain

ABSTRACT: Wet chemistry in organic solvents has proven highly ecient for the
preparation of several types of metallic, metal-oxide, and semiconductor nanostructures.
This Short Review focuses on the use of oleylamine (OAm) as a versatile reagent for the
synthesis of various nanoparticle systems. We describe the ability of OAm to act as a
surfactant, solvent, and reducing agent, as a function of other synthesis parameters. We also
discuss the specic role of OAm either alone or in combination with other reactants, to
form nanostructures using a variety of organic or inorganic compounds as precursors. In
certain cases OAm can form complex compounds with the metal ions of the corresponding
precursor, leading to metastable compounds that can act as secondary precursors and thus
be decomposed in a controlled way to yield nanoparticles. We also point out that OAmstabilized particles can often be dispersed in dierent organic solvents yielding solutions
with enhanced colloidal stability over long times and the potential to nd applications in a
number of dierent elds.
KEYWORDS: alkylamines, precursor complex, nonaqueous, ligand, capping agent, shape control, oleic acid

1. INTRODUCTION
Nanoscale materials can display physical and chemical properties that are dierent from those of their bulk counterparts. For
this reason, design and optimization of synthetic methods for
the production of nanoparticles have been intensively
investigated during the last few decades. As a result, important
progress has been achieved on the synthesis of nanomaterials
with tailored composition, size, shape, and crystalline structure.
Such nanostructures can nd applications in a wide range of
domains. Nevertheless, the scale-up of the successful
laboratory-adapted protocols to a more industrial level is
often not straightforward, and there is still a strong need for the
development of simple and inexpensive protocols for large-scale
synthesis of high quality nanoparticles.
Examples of wet-chemical routes for the synthesis of
nanoparticles are the seeded-growth method, the polyol
process, and the DMF-mediated reduction. In the two latter
approaches, diethyleneglycol (DEG) and dimethylformamide
(DMF) are used as solvents that can also reduce selected salts
to produce nanoparticles in the presence of surfactants or
polymers.1 Very often, the compound used as metal source
cannot be decomposed at room temperature, even in the
presence of strong reducing agents. However, thermolytic
reduction at higher temperatures can be achieved, in principle,
in two ways: (i) using a heating mantle or an oil bath to
increase the temperature of the reaction mixture (heat up
approach) and (ii) through the so-called hot-injection
technique, where a cold solution of precursor molecules is
rapidly injected into a hot coordinating alkyl solvent. Wang and
Li have recently reported an improved hot-injection method
which involves the simultaneous use of octadecylamine (ODA)
2013 American Chemical Society

as a solvent, surfactant, and reducing agent, for the controlled


preparation of a wide range of noble metal, metal oxide, and
bimetallic nanocrystals.2
Oleylamine (OAm) is another long-chain primary alkylamine, such as ODA or hexadecylamine (HDA), which can act
as electron donor at elevated temperatures. Interestingly,
commercial OAm has a much lower cost than commonly
used pure alkylamines, though some concerns regarding purity
and reproducibility have also been raised. Moreover, OAm is
liquid at room temperature, which may simplify the washing
procedures that follow the chemical synthesis of nanoparticles.
However, care should be taken during handling, since OAm can
be corrosive to the skin, as indicated in the corresponding
material safety data sheet. The molecular structure of OAm is
depicted in Scheme 1.3 The double bond (CC) in the middle
of the molecule is another special feature of the compound
under discussion. In particular, although ODA and OAm
exhibit similar basicity and anity to metals through their NH2
functional groups, the resulting morphology and crystallinity of
the produced nanoparticles can be signicantly dierent. An
example is the formation of Au decahedra in the presence of
either OAm or ODA, where the lack of the CC bond in
ODA was suggested to limit its coordination with AuCl, thus
yielding modied shapes.4 In the context of understanding the
bonding mode of OAm with nanoparticle surfaces, FTIR
spectroscopy has become a primary characterization tool.
Figure 1 shows the FTIR spectrum of pure oleylamine,5 while
Received: January 7, 2013
Revised: February 11, 2013
Published: February 12, 2013
1465

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

synthesis of heterostructures comprising a metal and a metal


oxide with combined properties. The preparation of certain
kinds of nanostructures containing rare-earth elements is also
presented. Finally, Section 7 presents several examples of largescale NP synthesis.

Scheme 1. Chemical Structure of Oleylamine


(CH3(CH2)7CHCH(CH2)8NH2)a

2. OAM IN THE SYNTHESIS OF MAGNETIC NPS


Oleylamine is widely used in the synthesis of nanostructures
comprising at least one magnetic element. The high boiling
point of OAm (350 C) allows the possibility to employ
strong heating conditions if necessary. OAm can not only act as
a solvent for many organic and inorganic compounds, but it can
also become a surfactant and even a mild reducing agent. Such
properties are certainly correlated to the specic nature of the
target nanomaterial and to the reaction conditions. For
example, in the presence of stronger reducing agents, the role
of OAm is limited to act as a surfactant and/or solvent. As a
rst example of the use of OAm in the synthesis of
monometallic magnetic nanoparticles, we refer to the work
by Nam et al.,7 where fcc-Co hollow nanoparallelepipeds were
prepared by thermolysis of the fcc-CoO solid counterparts in
neat OAm at 290 C. Plausible reaction pathways for such
transformation are proposed in Scheme 2. Pure OAm was also

Reprinted with permission from ref 3. Copyright 2011 Elsevier B.V.

Table 1 summarizes the assignment of its characteristic FTIR


absorption bands.5,6

Scheme 2. Proposed Reaction Pathwaysa,b

Figure 1. FTIR spectrum of pure oleylamine in the 10003500 cm1


region. Reprinted with permission from ref 5. Copyright 2003 Elsevier
B.V.

Table 1. Infrared Vibrational Assignments for the OAm


Molecule5,6
vibrational modesa

frequency (cm1)

as(NH2) and s(NH2)


(CH)
as(CH) and s(CH)
(CC)
(NH2)
(CH3)
(CN)
(CC)

3376, 3300
3006
2922, 2854
1647
1593, 795
1465
1071
722

Reprinted with permission from ref 7. Copyright 2008 WILEY-VCH.


(a) Reduction of fcc CoO to fcc Co by oleylamine with formation of
heptadecene and heptadecadiene; (b) conversion of oleylamine into
octadecene and octadecadiene by fcc Co.

s = symmetric stretching vibration; as = asymmetric stretching


vibration; = bending vibration.

In the following sections we review the role of oleylamine on


the synthesis of various types of nanoparticle systems. Section 2
is devoted to nanostructures that contain at least one magnetic
element (Fe, Co, Ni, etc.). Noble nonplasmonic nanomaterials (e.g., Pd- and Pt-based) are discussed in Section 3, while
plasmonic nanostructures (Ag, Au, Cu) are presented in
Section 4. Section 5 discusses the use of OAm for the synthesis
of semiconductor nanomaterials (for example, sulfur-based
systems). We show in Section 6 some examples the OAm-based

used to produce polypod-like structures using Co(OAc)2 as a


metal source.8 Chaudret and co-workers used H2-mediated
reduction to obtain hcp-Co nanorods using a complex of Co
with cyclooctadiene and cyclooctatetraene (Co(COD)(COT))
as precursor in anisole, with an oleic acid (OAc)alkylamine
surfactant mixture. They reported that OAm produced
nanorods with lower regularity and less opportunities for
long-range organization compared to the use of ODA, which
was related to the cis conguration and buckled molecular

1466

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

chain of the OAm molecule.9 In a more comprehensive study,


the same group managed to produce hcp-Co nanowires (NWs)
with micrometer scale length and 79 nm diameter, using a
selected oleic acid/OAm concentration ratio.10
Iron is another monometallic magnetic system of large
interest. OAm has been reported to complex iron pentacarbonyl, forming Fe(CO)x-OAm (x < 5), which can be used as
nanoparticle precursor.11 The thermal decomposition of this
complex precursor resulted in the formation of spherical iron
nanoparticles with high saturation magnetization and sizes
ranging between 2 and 10 nm. Farrell et al. have described the
synthesis of iron/iron oxide nanoparticles using OAm/OAc
surfactants via both heterogeneous and homogeneous
nucleation strategies.12 Slight modications of the above
protocols resulted in the production of Fe@FexOy@FePt
and Fe@FexOy@Pt coated particles.13 Sun and colleagues
employed OAm as the only surfactant in the preparation of
controllably oxidized Fe/Fe3O4 core shell nanoparticles.14 Such
particles could be transferred into water upon surface
modication, by replacing the OAm capping ligand with a
dopamine-based surfactant.
The third room-temperature ferromagnetic element, Ni, has
also been synthesized in nanoscale sizes with the help of OAm.
Highly disordered Ni nanoparticles with a size range of 816
nm were produced by decomposition of nickel acetylacetonate
(Ni(acac)2), in the presence of OAm as ligand, together with
OAc and trioctylphosphine (TOP).15 Carenco et al. investigated the role of the binary ligand system comprising OAm
and TOP for the synthesis of monodisperse, size tunable (ca. 2
to 30 nm) Ni nanoparticles, by the thermal decomposition of
Ni(acac)2.16 OAm served as the main reducing agent while
TOP provided a tunable surface stabilization through
coordination on the Ni(0) surface. The simple route of
combining Ni(acac)2 with OAm at high temperatures was also
studied by Zhang et al. through three dierent independent
processes: direct thermolysis, seed-assisted growth, and hot
injection. The product contained single-crystalline fcc Ni
particles (2060 nm) with narrow size distributions.17 On
the other hand, applying a strong reducing agent such as borane
tributylamine (BTB) resulted in the production of small (3
nm) Ni nanoparticles using Ni(acac)2 as precursor and OAm as
solvent and cosurfactant, together with OAc.18 Those particles
showed good catalytic behavior for hydrogen release from the
hydrolysis of ammoniaborane at ambient conditions.
The thermolytic reduction of Ni(acac)2 in alkylamines
(including OAm) illustrated the possibility of tuning the
crystalline structure of the nanoparticles. In principle, higher
reaction temperatures (e.g., >240 C) favored the formation of
an hcp phase, while lower temperatures yielded fcc Ni
particles.19 The particle size distribution and morphology
were further controlled by introducing additional surfactants
(OAc and TOP). The same group improved their approach by
using a TOPO/TOP surfactant mixture for better size control.
The control over the crystalline phase was achieved by
adjusting not only the reaction temperature but also the
amine concentration and heating rate.20 A similar fcc-hcp
structure tuning was reported when Ni acetate was used as
nickel source, while 1-adamantanecarboxylic acid (ACA) and
TOPO facilitated size control in the range of 5120 nm.21 This
interesting ability to tailor the crystal structure of Ni
nanoparticles has continued to draw scientic attention for
further investigation, and signicantly dierent magnetic

properties have been reported for fcc- and hcp-structured Ni


particles.22
The introduction of atoms as C or H in the Ni crystal lattice
has also been achieved using OAm as the reaction medium.
More specically, the addition of hydrazine helped toward the
formation of NiHx nanoparticles.23 These particles possess a
larger lattice constant compared to pure Ni, and their magnetic
characterization revealed both ferromagnetic and paramagnetic
features, indicating the existence of two compositional phases.
Besides, Schaefer et al. showed that heating a mixture of
Ni(acac) 2 with OAm and octadecene allowed Ni 3 C 1x
nanoparticles to be acquired with increasing carbon content
as the reaction time was prolonged. The insights from this work
helped to experimentally rationalize the discrepancies in lattice
constants and magnetic properties that were previously
reported for hcp-Ni.24 A relevant work by Goto et al. on the
thermolysis of nickel acetylacetonate in OAm is in agreement
with the existence of hexagonal nickel carbide. These authors
proposed a formation mechanism for such hcp Ni3C nanoparticles:25 metallic Ni NPs are rst formed by the thermal
decomposition of the acetylacetonate and reduction of Ni(II)
by OAm, and then carbidization into hexagonal nickel carbide
occurs, with formation of cubic nickel carbide as an
intermediate product. The carbidization of the fcc-Ni
progresses by exposure to the CO gas emitted during acac
decomposition above 240 C.
Hyeons group reported the synthesis of highly monodisperse Ni and NiO nanoparticles through the thermal
decomposition of NiOAm complexes. The authors explained
that the NiOAm complex was prepared by reacting Ni(acac)2
with OAm at moderate temperature. The resulting solution was
later injected into TPP (triphenylphosphine), followed by
stronger heating conditions. Finally the particles could be
readily dispersed in nonpolar solvents such as hexane and
toluene and presented the ability to self-assemble into
superlattices with long-range order via controlled solvent
evaporation. In addition, these particles were active catalysts
for a Suzuki coupling reaction.26 The Ni(acac)2OAm system
has been used for the synthesis and thorough characterization
of Ni/NiO coreshell nanoparticles,27,28 while the combination
of the traditional chemical synthesis with microwave irradiation
for the control of size, shape, and shell thickness has also been
reported.29 OAm has also served as a reaction medium for the
synthesis of Ni2PNi coreshell nanoparticles.30 In this case,
Ni(0) NPs were synthesized in a rst step, followed by the
addition of P4 and further in situ thermal treatment. A detailed
study on the role of OAm concentration during the synthesis of
nickel phosphides was reported by Muthuswamy et al. These
researchers prepared hollow Ni12P5 and Ni2P particles as well as
solid Ni 2 P NPs. It was shown that increasing OAm
concentration favored the formation of the Ni12P5 phase rather
than the Ni2P one.31 Moreover, it was shown that an increase in
OAm concentration was systematically related to a tunable
modication of void size in hollow Ni12P5 particles formed at
low P:Ni rations.
Metal Oxide Magnetic Nanostructures. OAm has been
employed successfully with a triple role (solvent, surfactant,
reductant) in the synthesis of hexagonal and cubic CoO
nanocrystals, using Co(acac)3 as precursor.32 Taking into
account that the reactions were carried out under inert
atmosphere, the oxygen in CoO is assumed to originate from
the acac ligand of the precursor.
1467

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

Size- and shape-controlled FeO nanoparticles (Figure 2)


were prepared by reductive decomposition of Fe(acac)3 in an

Figure 3. Self-assembled multilayer pattern of octahedral Fe3O4


nanoparticles. TEM (a) and SEM (bd) images of a self-assembled
Fe3O4 NPs multilayer. Reprinted with permission from ref 39.
Copyright 2009 Royal Society of Chemistry.

Figure 2. TEM images of FeO nanoparticles: (a) 14-nm spheres and


(b) 32-nm and (c) 53-nm truncated octahedra. (d) SEM image of 100nm truncated octahedra. The sizes refer to average lengths of one side
in the projected images. (reproduced from ref 33 with permission from
Wiley-VCH).

Such nanomaterials presented size-dependent but rather


unusual magnetic properties.40 Moreover, MnO nanospheres
and nanorods were produced by tuning the temperature in a
reaction involving the preparation of a MnOAm complex,
which was subsequently injected into preheated TOP.41
Bimetallic Magnetic Nanoparticles. Magnetic nanostructures comprising at least two metals, of which at least one is a
magnetic element, have been successfully synthesized using
OAm as a reaction medium. Chaubey et al. employed OAm as a
cosurfactant, together with OAc, in the synthesis of 20 nm
FeCo nanoparticles with high saturation magnetization, using
Fe(acac)3 and Co(acac)2 as precursors.42 Polycrystalline FeCo
nanoparticles with a random orientation were obtained using
the above surfactant mixture with Co(COD)(COT) and
Fe(CO)5 as metal sources.43 OAm was also used as surfactant
for growing Fe on preformed Co NPs by Fe(CO) 5
decomposition at 180 C. The further thermal treatment of
the as-synthesized Co/Fe coreshells at 250 C yielded FeCo
alloy NPs.44 Ternary superparamagnetic alloys with tunable
composition (FexCoyPt100xy) were also prepared using the
OAc/OAm surfactant mixture.45 In addition, anisotropic
cobaltiron phosphide nanocrystals with controlled composition and magnetic response were produced using the Co and
Fe oleates, in the presence of OAm and TOP.46
Superparamagnetic Co50Ni50 and ferromagnetic Co80Ni20
nanoparticles were produced using OAm in a triple ligand
mixture together with OAc and TOP. It was suggested that the
metalacetylacetonate precursors were rst decomposed at 200
C to form their TOP complexes, which were later
decomposed at 240 C to produce Co and Ni nuclei.47 Ni
Co nanoparticles with a Ni-rich core and a Co-rich shell were
produced by mixing and heating NiOAm and CoOAm
complex precursors under microwave irradiation (Figure 4).
Those precursors were initially prepared by stirring and heating
at intermediate temperatures mixtures of Co(II) formate
dihydrate or Ni(II) acetate tetrahydrate with OAm.48

OAc/OAm mixture.33 The binding dierence between oleate


and OAm on the crystal planes was claimed to determine the
nal morphology of the particles. The same argument was used
to explain the growth of -Fe2O3 tetrapods using a ternary
surfactant mixture (OAc, OAm, and hexadecanediol).34 Fe3O4
(magnetite) nanocrystals in the range of 710 nm were also
produced using OAm as reducing agent, stabilizer ,and
cosolvent with benzyl ether, in a facile heating protocol using
iron(III) acetylacetonate as metal source.35 Suitable ratios
between OAm, OAc, and hexadecanediol resulted in ca. 4 nm
Fe3O4 nanoparticles, which readily formed self-assembled
monolayers and multilayers upon evaporation of hexane.36
This approach was extended to the synthesis of other ferrite
nanocrystals with the general formula MFe2O4 (M = Fe, Co,
Mn).37 It was shown that the simultaneous use of OAc and
OAm was necessary for the desired particle formation, whereas
OAc alone would lead to a viscous red-brown product which
was dicult to purify and characterize. On the other hand,
OAm alone did produce iron oxide nanoparticles but in a very
low yield. Regarding more anisotropic shapes, star-like cubes
and ower-like magnetite nanoparticles were synthesized by the
pyrolysis method using surfactants with dierent hydrocarbon
structures such as OAm on the one hand and adamantaneamine, adamantanecarboxylic acid, and trioctylamine on the
other hand.38 It was considered that the linear structure of
OAm should attack the electropositive carbonyl carbon of an
intermediate iron complex more easily than the bulky
adamantyl groups of adamantaneamine (or the corresponding
bulky groups of trioctylamine), thus inuencing the particle
growth mode. Additionally, the molar ratio of OAm to iron
oleate has proven to be crucial for the formation of octahedral
ferromagnetic Fe3O4 NPs (Figure 3).39
Other magnetic oxide NPs such as MnO and Mn3O4 have
been prepared by prolonged heating of slurries of OAm with
Mn(acac)2 in the absence or presence of water, respectively.
1468

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

also resulted in the production of spherical FePd nanoparticles


when starting from iron pentacarbonyl and Pd(acac)2.51
FePt nanowires and nanorods were obtained in a simple
protocol reported by Sun and co-workers, where OAm was
used alone or diluted with octadecene in a heated solution
containing Pt(acac)2 and Fe(CO)5. It was proposed that
possibly OAm self-organized into an elongated reverse-micellelike structure within which the FePt nuclei were formed.52 The
same researcher was involved in a pioneering, highly inuential
work on the formation of monodisperse FePt nanoparticles and
ferromagnetic FePt nanocrystal superlattices using OAm and
OAc as stabilizers.53 Modications of the latter protocol
(absence of hexadecanediol, sequential addition of surfactants)
allowed the preparation of more faceted FePt NPs with an
increased size up to 9 nm and the possibility to coat with a
Fe3O4 shell if an excess of iron precursor was used.54 For
example, it has been reported that if OAm was added rst,
sphere-like FePt nanoparticles were obtained, possibly through
the formation of a stable Pt-NH2 complex. On the other hand,
if OAc was added in a rst stage before the addition of OAm,
7 nm FePt nanocubes were obtained.55 In the OAm/OAc
surfactant pair, Fe tends to bind to OOC while Pt binds to
the NH2 group. This dual ligand system has been introduced
as the optimum for the preparation of extended ordered layers
of monodisperse FePt nanoparticles, in comparison, for
instance, to ligand pairs such as octanoic acid/octylamine,
oleic acid/ODA, or octadecanoic acid/ODA.56 Advanced realtime TEM imaging has provided important knowledge toward
the understanding of the mechanism of one-dimensional
colloidal nanocrystal growth during the formation of FePt3
nanorods in a pentadecane/OAm system in the presence or
absence of OAc.57 It has been shown that the OAm/OAc
binary surfactant system can be replaced by PEI (polyethyleneimine) via ligand exchange to form a PEI/FePt NPs
multilayer assembly. Such an assembly might be suitable for
ultrahigh-density data storage media, as its thermal annealing
transformed the initial chemically disordered fcc phase into the
chemically ordered tetragonal fct phase, rendering FePt
nanoparticles with the desirable high magnetocrystalline
anisotropy and room temperature ferromagnetism.58,59 Such
slightly reductive annealing treatment (e.g., under a ow of
Ar(95%)/H2(5%)) has also been described to result in the
formation of relatively large (17 nm) fct-FePt NPs, using Pt@
Fe2O3 coreshell NPs prepared in the presence of OAm/
OAc.60 The synthesis of biocompatible FePt@Au nanoparticles
also required a ligand exchange approach. In this case the
OAm/OAc pair was successfully replaced by mercaptoundecanoic acid (MUA), rendering the particles water-dispersible
while the gold coating would allow further functionalization
with dierent biomolecules and provide heating resistance,
which is necessary for hyperthermia applications.61
The synthesis of Ni-based bimetallic magnetic NPs was also
carried out in the presence of OAm. Irregular-shaped NiFe NPs
were formed using OAm in the triple (solvent, reductant,
surfactant) role, during the decomposition of Fe(acac)3 and
Ni(acac) 2 . The morphology was better controlled by
introducing cosurfactants such as dodecanediol or TOP.62
Nickel and platinumacetylacetonates were reduced into
spherical NiPt NPs using the OAm/OAc binary surfactant in
benzyl ether.63 NiPt3 nanopolyhedra were produced via the
tungsten-mediated reduction of Ni(acac)2 and Pt(acac)2 in the
presence of OAm/OAc. The authors explained that OAm
played the role of a reducing agent but it could also stabilize

Figure 4. HAADF-STEM image of Ni50Co50 NPs (a). Elemental maps


of Co (b), Ni (c), and O (d) in the area shown in (a). Reprinted with
permission from ref 48. Copyright 2011 American Chemical Society.

Interesting ferromagnetic CoPt polypod-like nanostructures


were obtained by direct thermolytic reduction of Pt(acac)2 and
Co(OAc)2 in hot OAm (Figure 5). In this case oleylamine

Figure 5. Bright eld TEM micrographs of CoPt nanopolypods


prepared in oleylamine. Inset: Electron diraction pattern. Reproduced
with permission from ref 49. Copyright 2005 American Chemical
Society.

played the role of high boiling point coordinating solvent, as


well as reducing and capping agent. It was proposed that OAm
might act as a ligand to form stable complexes with Pt2+ while
heating induced the thermolytic reduction of such Pt complexes
to the metallic state, and the obtained small Pt seeds could
reduce Co2+ to Co(0), thereby generating CoPt nanoparticles.49 A similar process, but replacing Coacetate with
Co2(CO)8, resulted in the formation of either nanowires or
ower-like CoPt structures, depending on the specic temperature and concentrations of the employed reactants.50 Another
cobalt precursor, Co(CO)3NO, was applied for the formation
of spherical CoPt nanoparticles using OAm as a cosurfactant
with OAc and 1,2-hexadecanediol as reductant. This approach
1469

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

used the H2-mediated reduction approach for the preparation


of Pt nanostructures with complex shapes such as cubic
dendrites and vefold stars (Scheme 4), by reacting H2PtCl6

Pt3Ni {111} by lowering the surface energy of Pt3Ni {111}


facets, thus inducing an octahedral morphology.64 NiCu alloy
nanoplates with either hexagonal or triangular shape were
produced using OAm as a solvent, reductant, and cosurfactant
with TOP.65 The nal shape of those particles was determined
by the overall reducing environment of the reaction (temperature, absence or presence of benzyl ether as cosolvent
Scheme 3). Moreover, the OAm/OAc ligand pair facilitated the

Scheme 4. General Overview of the Versatile Synthesis of Pt0


NPs in OAma,b

Scheme 3. Schematic Illustration for the Formation of


Triangular and Hexagonal NiCu Alloy Nanoplatesa

Reprinted with permission from ref 75. Copyright 2012 Wiley-VCH.


The shape could be controlled by the platinum concentration and the
nature of the seeds.

with OAm at selected temperatures, Pt concentrations, and


dihydrogen pressures.75 In another approach, in situ formed Au
nanoparticles during the reaction of PtCl2 in OAm in the
presence of AuCl3 triggered the nucleation and anisotropic
growth of Pt nanoowers, nanopolyhedra, and other similar
structures.76 However, gold nanocrystals were rather inactive
toward the seeded growth of Pt nanocrystals in dichlorobenzene, in the presence of OAm and hexadecanediol. Cobalt
traces were used as surfactant-independent means of shape
control, leading to the formation of various anisotropic Pt
nanocrystal morphologies, as reported by Puntes and
colleagues.77 Besides, Tilley and co-workers published several
works on the reduction of platinum acetylacetonate in OAm
under hydrogen atmosphere. They managed to produce shapes
such as octapods and cubes, showing, for example, that the
initial Pt concentration can aect the reaction kinetics and the
nal shape.7880
The platinum precursortungsten system formed by
introduction of W(CO)6 in the reaction mixture served as a
buer facilitating also the growth of Pt and Pt3Co nanocubes,
in the presence of OAm/OAc.81 The binary surfactant system
was in a 4:1 volume ratio, and both compounds acted as
solvents and capping agents. The resulting nanocubes showed
good catalytic activity for the electrooxidation of methanol.
Besides, OAm served as a surfactant during the preparation of
Pt and PtCo nanocrystals, where the inuence of competing
reducing agents toward size and shape control was examined.82
Pt-on-Pd bimetallic nanostructures were also prepared in the
presence of OAm. This protocol involved a two-step procedure,
and the nal material was tested for electrocatalytic oxygen
reduction.83 In addition, various Pt alloys (PtM, M = Pd, Ni,
Fe, Co) with controlled shape and composition were prepared
using the CO gas reduction approach and OAm as one of the
surfactants.84 PtAu nanoparticles are also interesting catalysts. It
has been suggested that, in the presence of butyllithium, a
strong reductant, OAm may be deprotonated to form an amide,
and the amine-to-amide ratio might be important for PtAu
NWs formation and subsequent branching.85 In a rather simple
protocol, the sole use of OAm was reported to reduce a mixture
of HAuCl4 and H2PtCl6 at 160 C, resulting in the formation of
PtAu NPs.86 The seed-mediated method was employed for the
preparation of heterogeneous PtAu nanostructures, where the
role of OAm had to do mainly with the initial reduction of
HAuCl4 to form the gold seeds.87,88

Reprinted with permission from ref 65. Copyright 2012 Royal Society
of Chemistry.

synthesis of MnPt alloy NPs with spherical or cubic shape in


ether solvents, which displayed interesting magnetic and
electrocatalytic properties, respectively.66,67

3. NOBLE (NONPLASMONIC) METAL


NANOPARTICLES
Pt-Based Nanostructures. Nanostructures based on
metals such as palladium and platinum have attracted interest
for various applications, such as catalysis and electrocatalysis.
OAm was used as a surfactant together with OAc for the
synthesis of faceted Pt nanoparticles in the 37 nm size range.
These particles were active for the catalytic reduction of O2.68
The OAm/OAc combination helped also toward the
production of 8 nm Pt nanocubes, in the presence of trace
amounts of Fe(CO)5 in octadecene.69 Unlike the preparation of
FePt nanocubes, in which the addition sequence of OAc and
OAm was important,55 the synthesis of these Pt nanocubes did
not require such sequential surfactant addition. Wu et al.
demonstrated the crucial role of carbon monoxide (CO) as a
reductant and stabilizer of the Pt(100) surface for the formation
of Pt nanocubes in OAm, even in the absence of oleic acid.70
The ability of CO to inuence the reaction kinetics was
extended also for the production of Pt nanocubes, Pd NPs, and
Au NWs in the presence of OAm.71 The thermolytic reduction
of Pt(NH3)2Cl2 in dierent reaction media as OAm, HDA, and
oleyl alcohol resulted in dierent anisotropic structures. In that
case, the nature of the selected solvent (including its capping
and reducing capabilities) seemed to play a major role in
directing the morphology of the nal structure.72 The use of a
suitable environment to burst nucleation, using morpholine
borane and N-methyl-pyrrolidone, yielded 1012 nm Pt
nanocubes within one minute of reaction, where OAm served
as solvent and surfactant.73 Control of the reaction kinetics in a
relatively simple protocol involving the heating of Pt(acac)2 in
OAm was shown to yield either spherical or branched Pt
nanostructures, depending on the reaction temperature (higher
temperatures favored more isotropic shapes).74 Lacroix et al.
1470

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

The role of the OAm/OAc mixture on the compositiondependent formation of PtAg NWs via the oriented attachment
mechanism on the {111} surface was proposed by Peng et al.89
OAm was also necessary for the formation of PtCu nanorod
catalysts, presumably due to its tendency to stabilize the (100)
facets during growth.90 Furthermore, high-quality PtCu nanocube catalysts for formic acid oxidation were synthesized thanks
to a synergistic stabilizing eect oered by a mixture of OAm
and tetraoctylammonium bromide.91
Other Noble Metal (Nonplasmonic) Nanostructures.
OAm acted as a solvent, surfactant, and coreductant of
Pd(acac)2 (together with BTB) for the production of 4.5 nm
Pd NPs, which were evaluated for catalytic formic acid
oxidation.92 An excess of OAm/OAc with respect to the
Pd(OAc)2 concentration led to the synthesis of 3.0 nm Pd NPs
via the reducing ability of tert-butylamine borane (TBAB).93
On the other hand, OAm was found to form intermediate
complexes expressed as [Pd(acac)x(OAm)y] in the presence of
formaldehyde, and the formation of these intermediates
allowed a good kinetic control of the synthesis with subsequent
shape control over the nal products.94 The role of OAm was
also demonstrated upon the study of the eect of the local
ligand environment on the nucleation and growth of NPs in a
two-component reagent mixture (Pd precursorsystematically
variedand OAm).95 The combination of OAm with
alkylammonium alkylcarbamate (AAAC) allowed the manipulation of the nal Pd morphology from spheres to tetrahedra
and polypods.96 Metin et al. reported that OAm not only
stabilizes Pd(0) nanoclusters but in addition does not have any
deleterious eect on their catalytic activity for the dehydrogenation of ammoniaborane.97 Tilley and colleagues employed
the H2-mediated reduction for the preparation of Pd
nanostructures, when the introduction of OAc as a cosurfactant
with OAm favored the formation of highly branched
morphologies, while OAm alone yielded rather isotropic
shapes.98,99 On the other hand, the simultaneous use of OAm
with TOP has also been reported to provide monodisperse Pd
NPs, both using Pd(acac)2100 and Na2PdCl4 as precursors.101
Rh and Ir NPs were also prepared in OAm as a reaction
medium. OAm-capped rhodium nanotetrahedra showed
excellent catalytic activity for the hydrogenation of anthracene.102 On the contrary, the insertion of OAc (together with
OAm) for the synthesis of iridium particles rather quenched
their catalytic activity for the hydrogenation of 1-decene.103

Figure 6. TEM images at dierent magnications of Au NWs prepared


by OAm reduction (a, b) and HRTEM image (c) showing their singlecrystalline structure. Reproduced with permission from ref 104.
Copyright 2008 American Chemical Society.

had previously reported the synthesis of single-crystalline thin


gold wires, but the method was rather complicated as it
involved a variety of reagents, high temperature, and
intermediate steps.111 By lowering the precursor concentration
in the latter protocol, spherical Au particles were obtained.112
These NPs displayed strong broad-band optical-limiting
properties for both femtosecond and nanosecond laser pulses.
On the other hand, thicker Au NWs (d 9 nm) were produced
by adding oleic acid in the HAuCl4/OAm mixture in a 1:1
volume ratio with respect to OAm.113 If chloroauric acid was
dissolved in water prior to addition of excess OAm, 1113 nm
NPs were nally isolated, as shown by Polavarapu and Xu.114
Size tuning in the 110 nm range was achieved by using the
burst nucleation method with TBAB as reductant in the
precursor/OAm mixture and tetralin as solvent. These particles
were highly active for CO oxidation.115 Moreover, the addition
of OAc in the presence of a dierent precursor, Au(ac)3, in
OAm led to the formation of 6.7 nm Au NPs.116 Although
OAm-capped Au NPs are hydrophobic, they can be made
hydrophilic through a ligand-exchange procedure that involved
replacing OAm with 3-mercaptopropionic acid.117
Silver nanoparticles were also obtained by using OAm as a
reducing agent for AgNO3 and stabilizer in the presence of a
high-boiling-point solvent, liquid paran.118 Size- and shapecontrolled Cu nanoparticles were prepared with the OAm/OAc
surfactant mixture in the presence of hexadecanediol as
reductant and copper acetylacetonate as metal precursor.119
Besides, Hyeon and colleagues have shown that Cu(acac)2
could be thermally decomposed in neat OAm to yield uniform
15 nm Cu particles which readily formed a Cu2O oxide shell
upon air exposure. These Cu@Cu2O coreshell NPs were
good catalysts for Ullmann type coupling reactions of aryl
chlorides.120 AuxCuy alloy nanoparticles were also produced by
using excess OAm in the presence of hexadecanediol at 160 C,
employing HAuCl4 and Cu(acac)2 as metal sources. These
particles were envisaged not only for optical but also for
catalytic applications (e.g., electrochemical reduction of
CO2).121 Wang et al. used OAm as reductant/surfactant, in
the presence of a small amount of octadecene and chloroauric
acid/silver nitrate metal precursors at moderate temperature,
120 C. The weak bonding of OAm allowed its easy removal

4. PLASMONIC NANOSTRUCTURES
Nanoparticle systems based on gold, silver, and copper have
been successfully prepared by using oleylamine. A relatively
simple general protocol that involves the use of only two
reagents (HAuCl4 and OAm) has been simultaneously reported
by several groups104106 for the synthesis of long, ultrathin Au
NWs (Figure 6). The main idea behind the formation
mechanism of the Au NWs was that mesostructures of Au+
OAm complex were initially formed, serving as growth
templates that govern one-dimensional growth in the nanoscale. In certain cases, Ag or Fe NPs were inserted in the
reaction mixture to boost the reduction stage or act as growth
templates.107109 The addition of triisopropylsilane (TIPS) as
highly eective reductant in the binary reactant mixture
accelerated the reaction rate and improved the yield of ultrathin
gold nanowires at room temperature within a few hours. These
NWs demonstrated an intriguing application in surfaceenhanced Raman scattering (SERS).110 Halder and Ravishankar
1471

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

from the surface of the particles, rendering clean enough


AuAg NPs that are catalytically active for CO oxidation, as
shown by the corresponding evaluation tests.122

5. NANOSCALE SEMICONDUCTORS PREPARED WITH


OLEYLAMINE
Colloidal semiconductor nanocrystals can also be synthesized
in the presence of OAm. For example, uniformly sized Ag2S
nanocrystals were prepared by pyrolysis using AgNO3 and S
powder as precursors and OAm as solvent, surfactant, and
reductant. These particles were proposed for application as
SERS substrates.123 A similar heat-up protocol using OAm was
employed for the synthesis of CdS nanorods and CdSe
nanoparticles.124 Highly luminescent CdSe NPs were produced
using OAm as a coligand whose concentration aected the
reaction kinetics and the nal particle size.125 Elemental sulfur
and copper acetylacetonate were thermally treated in OAm to
obtain monodisperse hexagonal Cu2S nanoplates. These
materials were considered as potential solutions for application
in elds such as solar cells, photoelectric devices, and colloidal
photonic crystals.126 The combination of Pb(OAc)2, dodecanethiol, and OAm has been recently reported to yield PbS
nanocrystals with polyhedral or truncated-cubic shape depending on the reaction temperature.127 PbS NPs were also
prepared by mixing and heating initially prepared PbCl2
OAm and sulfurOAm solutions.128 PbTe nanocubes were
prepared by mixing OAm/OAc surfactants with Pdacetate,
followed by hot-injection of trioctylphosphine telluride at 200
C.129 ZnS nanostructures with dierent morphologies were
also obtained by heating a mixture of OAm, zinc stearate, and
sulfur precursor at 280 C. The kind of precursor used (sulfur
powder, thiourea, dodecanethiol) determined the nal particles
morphology (nanorods, dot-shaped, and quasi-cubic-shaped,
respectively).130 Reagents such as TOP, octenoic acid, and
dichlorobenzene were added for the production of high-quality
Ni3S4 and CuS nanocrystals with OAm as the reaction
medium.131 Hyeons group reported a generalized strategy for
the synthesis of semiconducting metal sulde NPs (PbS, ZnS,
CdS, and MnS) with various sizes and shapes. This procedure
initially involved the formation of a metaloleylamine complex,
by dissolution of metal chloride with OAm at an intermediate
temperature, and the nal nanomaterials were obtained after
injecting elemental sulfur and further heating.132 Thomson et
al. used NMR techniques to study the sulfuroleylamine
interactions during the synthesis of sulde NPs. They showed
that thioamides can be also used as S precursors due to their
rapid kinetics.133
Copper-based quaternary chalcogenides have recently drawn
interest as low-cost alternatives to conventional absorber
materials in photovoltaics. Shavel et al.134 prepared Cu2ZnSnS4
nanoparticles with controlled composition by reacting metal
amino complexes with S in a continuous ow reactor at 300
330 C (Figure 7). The complexes were formed by dissolving
metal precursors in an OAm/octadecene mixture. OAm was
also employed for the production of Cu2ZnSn(SxSe1x)4,135
CuInS2,136 and CuInSe2137 and other Cu-based selenide
NPs.138
Germanium NWs139 and NPs140 were produced in the
presence of OAm. In the latter case oleylamine initially served
for the formation of a secondary complex precursor after
mixing with GeCl4. Additionally, In2O3 nanoparticles were
produced by thermal decomposition of In(acac)3 in OAm
under inert atmosphere. The oxygen in the nanoparticles

Figure 7. (A) Scheme of the ow reactor setup and image of a 1 g


pellet made of CZTS nanoparticles. (B) TEM micrograph of cleaned
CZTS nanoparticles prepared inside the ow reactor at 300 C at a
ow rate of 2.0 mL/min. The inset shows a HRTEM image of a CZTS
nanocrystal and the corresponding SAED pattern. Reproduced with
permission from ref 134. Copyright 2012 American Chemical Society.

composition was claimed to originate from the acetylacetonate


ligand of the precursor.141 The combination of OAm, OAc, and
trimethylamine N-oxide in hexadecane was considered essential
for the formation of monodisperse indium oxide NPs upon
decomposition of indium acetate.142 Selishcheva et al. recently
reported that the addition of copper ions in the reaction
solution can help toward the shape tuning of the obtained
In2O3 nanostructures.143 NMR data indicate that OAm is
converted to oleylamide during nanoparticle formation:
2In(CH3COO)3 + 6R NH 2
= 6R NH CO CH3 + In2O3 + 3H 2O

On the other hand, ZnO nanostructures with several shapes


such as nanorods, nanotetrahedrons, and nanosquamas were
produced via an eective aminolytic reaction of zinc
carboxylates with OAm in coordinating or noncoordinating
solvents. These nanostructures displayed interesting optical
properties such as sharp band-edge emission or broad deep-trap
emission, depending on the shape and the state of their
structural defects.144

6. HETEROSTRUCTURES AND RARE EARTH-BASED


NPS
A variety of heterostructured nanomaterials composed of a
metal (Au, Ag, Pt, or Ni) and Fe3O4 or MnO were synthesized
by thermal decomposition of mixtures of metaloleate
complexes (for the oxide component) and metalOAm
complexes (for the metallic component).145 AgFexOy hybrid
nanoparticles were obtained by reducing Ag+ ions by OAm in
the presence of Fe/FexOy nanoparticles. These structures
exhibited both plasmonic and superparamagnetic properties,
thus being candidate materials, for example, to be used as both
SERS substrates and contrast agents for magnetic resonance
imaging (MRI) and optical imaging.146 OAm also served as a
surfactant and reducing agent of HAuCl4 to coat Fe3O4 NPs
with Au shells.147 In the approach by Yu et al.,148 Au NPs were
1472

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

rst formed, followed by decomposition of Fe(CO)5 on their


surface and subsequent oxidation, to yield AuFe3O4 dumbbell-like nanoparticles. In this case OAm/OAc were used as
surfactants and octadecene was the high-boiling-point solvent.
Platinum was attached on the surface of iron oxide NPs with
the help of OAm as surfactant and dodecanediol as the
reducing agent of K2PtCl4.149 In addition, FePt@MnO
nanoheterostructures were prepared by heating manganese
oleate with FePt NPs in benzyl ether containing OAm/OAc
surfactants.150
Surface-modied rare-earth compound nanocrystals have
attracted interest due to their unique optical properties and
promising applications in, for example, UV shielding,
luminescent displays, optical communication, biochemical
probes, and medical diagnostics. The role of the OAm/OAc
surfactants on the synthesis of anisotropic La2O3, Pr2O3,
Nd2O3, Eu2O3, and other rare-earth oxide nanocrystals has been
discussed by Si et al.151 Ceria (CeO2) nanoparticles were
prepared by decomposing cerium(III) nitrate hexahydrate using
OAm surfactant. Adding oleic acid as a cosurfactant with OAm
led to the production of anisotropic wire- and tadpole-shaped
nanocrystals.152 The OAm/OAc surfactant pair was also
employed for the formation of gadolinium oxide (Gd2O3)
NPs in octadecene.153 Moreover, the eect of the concentration
of OAm on the size of NaYF4:2% Er3+,20%Yb3+ particles has
been studied by Ostrowski et al.154 The cooperative eect of
the OAm/OAc surfactant pair toward the synthesis of
monodisperse NaLa(MoO4)2 bipyramids has been discussed
in detail by Bu et al.155 In fact, Cao and colleagues have shown
that these two commonly used surfactants can be condensed
together to form N-(cis-9-octadecenyl)oleamide (OOA) during
the synthesis of uranium dioxide (UO2) NPs. Their results
claim to provide unambiguous evidence that it is the OOA
not the simple mixture of OAc and OAmthat plays the major
role in controlling the formation of UO2 nanocrystals.156 The
above binary ligand pair was also used for the formation of
ferromagnetic SmCo5 NPs,157 as well as one-dimensional
W18O49 and V2O5 nanostructures.158

h in air,161 whereas Bi2Te3 nanoplates were prepared by


reaction between bismuth thiolate and tri-n-octylphosphine
telluride in OAm. The subsequent fabrication of n-type
nanostructured thermoelectric materials was implemented
through the sintering of surfactant-removed Bi2Te3 nanoplates
using a spark plasma sintering process.162
Hiramatsu and Osterloh reported almost a decade ago a
simple large-scale synthesis of nearly monodisperse Au and Ag
nanoparticles using only three reagents (tetrachloroauric acid
or silver acetate, OAm, and a solvent).163 The synthesis of
about 10 g of Ag3PO4 NPs with precise size control ranging
from 8 to 16 nm has been achieved by a room-temperature
reaction of silver nitrate, oleylamine, and H3PO4 in a toluene/
ethanol solvent mixture. These NPs exhibited high visible light
photocatalytic activity.164 Water-dispersible cubic ceria nanocrystals were produced by a simple large-scale solgel reaction
of cerium salt in the presence of OAm.165 The use of
microwave heating was proposed as a scalable approach for the
synthesis of rare earth oxides (M2O3, M = Pr, Nd, Sm, Eu, Gd,
Tb, Dy) with various shapes, as it provides the advantage of
avoiding thermal gradient eects. Typical reaction conditions
included the combination of metal acetate or acetylacetonate
precursors with the OAm/OAc surfactant mixture.166
Fe/Fe3O4 core/shell NPs prepared following the protocol
described by Peng et al.14 were employed by Jaeger and coworkers for the fabrication of large-scale freestanding nanoparticle monolayer membranes via a drying-mediated selfassembly process. These freely suspended layers showed
remarkable mechanical properties with Youngs moduli of the
order of several GPa, regardless of membrane size.167
Moreover, Shi et al. reported that monodisperse magnetite
nanocrystals can be successfully prepared in large quantities via
a facile synthetic procedure based on the pyrolysis of Fe(acac)3
in the presence of OAm/OAc surfactants.168 The presence of
an organic OAm layer was reported as the key point to serve as
a carbon source for the large-scale preparation of graphitic
carbon-coated FeCo NPs via the reductive thermal conversion
of nanometric Prussian blue analogues (PBAs).169
Finally, it has to be noted that OAm is reported, in many
cases, to improve the colloidal stability of various types of
nanostructures, for example, including dispersions of nickelcoated single walled carbon nanotubes (SWNTs) in THF or
toluene. In the above example, the electrons in the SWNTs
interact with the hydrophobic part of the OAm molecule, and
the amine groups interact with the nickel atoms. Amineamine
interactions between OAm molecules lead to formation of
micelles, which help to form a network in the solvent. Such NiSWNT materials are aimed to be used as llers in composite
materials, for the enhancement of electrical conductivity and
optimized performance in lightning-strike protection of aircrafts.170

7. OLEYLAMINE FOR LARGE-SCALE NANOPARTICLE


SYNTHESIS
The signicance of OAm as reagent of choice in nanoparticle
synthesis, not only from the academic point of view but also
regarding technological exploitation, is better illustrated by the
separate presentation of some characteristic examples of its use
in large-scale NP synthesis. Large-scale preparation is usually
achieved at the expense of compromising NPs quality.125
However, some of the synthetic protocols presented in the
former sections are readily adaptable for large-scale production,
which may even reach gram-scale quantities.93,110,114,126,134,152
Du et al. have recently reported a general method for the largescale synthesis of metal sulde nanocrystals (CuS nanosheets,
ZnS NWs, Bi2S3 NWs, and Sb2S3 NWs). For example, ultrathin
CuS nanosheets were prepared from an intermediate lamellar
complex formed via the reaction of CuCl with OAm and
octylamine, followed by the addition of S powder at moderate
temperatures.159 Large-scale superlattices of wurtzite CuInS2
nanocrystals were obtained by solvent evaporation using copper
nitrate, indium nitrate, dodecanethiol, OAc, and OAm. The
band gap of these CuInS2 NPs was measured to be 1.63 eV,
which is optimal for solar cell applications.160 High-quality
Bi2S3 nanorods were successfully synthesized in large scale by
reacting Bi[S2P(OC8H17)2]3 with OAm at 140160 C for 5

8. SUMMARY AND OUTLOOK


This Short Review illustrated all the important aspects of the
suitability and the versatility of oleylamine as the reaction
medium for the chemical synthesis of a broad range of
nanoscale materials. We showed the dierent and combined
roles of OAm in such syntheses, which demonstrate its
suitability as solvent, surfactant, and reducing agent. Other
advantages of the use of oleylamine, such as its liquid state at
room temperature, easy removal via centrifugation, high boiling
point, low cost, and tendency to form metalOAm complexes
at intermediate temperatures, so that it can be controllably
1473

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

(15) Winnischofer, H.; Rocha, T. C. R.; Nunes, W. C.; Socolovsky, L.


M.; Knobel, M.; Zanchet, D. ACS Nano 2008, 2, 1313.
(16) Carenco, S.; Boissiere, C.; Nicole, L.; Sanchez, C.; Le Floch, P.;
Mezailles, N. Chem. Mater. 2010, 22, 1340.
(17) Zhang, H. T.; Wu, G.; Chen, X. H.; Qiu, X. G. Mater. Res. Bull.
2006, 41, 495.
(18) Metin, O.; Mazumder, V.; Ozkar, S.; Sun, S. J. Am. Chem. Soc.
2010, 132, 1468.
(19) Chen, Y.; Peng, D. L.; Lin, D.; Luo, X. Nanotechnology 2010, 18,
505703.
(20) Chen, Y.; Luo, X.; She, H.; Yue, G. H.; Peng, D. L. J. Nanosci.
Nanotechnol. 2009, 9, 5157.
(21) Mourdikoudis, S.; Simeonidis, K.; Vilalta-Clemente, A.; Tuna,
F.; Tsiaoussis, I.; Angelakeris, M.; Dendrinou-Samara, C.; Kalogirou,
O. J. Magn. Magn. Mater. 2009, 321, 2723.
(22) Luo, X.; Chen, Y.; Yue, G. H.; Peng, D. L.; Luo, X. J. Alloys
Compd. 2009, 476, 864.
(23) Jeon, Y.; Lee, G. H.; Park, J.; Kim, B.; Chang, Y. J. Phys. Chem. B
2005, 109, 12257.
(24) Schaefer, Z. L.; Weeber, K. M.; Misra, R.; Schiffer, P.; Schaak, R.
E. Chem. Mater. 2011, 23, 2475.
(25) Goto, Y.; Taniguchi, K.; Omata, T.; Otsuka-Yao-Matsuo, S.;
Ohashi, N.; Ueda, S.; Yoshikawa, H.; Yamashita, Y.; Oohashi, H.;
Kobayashi, K. Chem. Mater. 2008, 20, 4156.
(26) Park, J.; Kang, E.; Son, S. U.; Park, H. M.; Lee, M. K.; Kim, J.;
Kim, K. W.; Noh, H. J.; Park, J. H.; Bae, C. J.; Park, J. G.; Hyeon, T.
Adv. Mater. 2005, 17, 429.
(27) Johnston-Peck, A. C.; Wang, J.; Tracy, J. B. ACS Nano 2009, 3,
1077.
(28) Railsback, J. G.; Johnston-Peck, A. C.; Wang, J.; Tracy, J. B. ACS
Nano 2010, 4, 1913.
(29) Chopra, N.; Claypoole, L.; Bachas, L. G. J. Nanopart. Res. 2010,
12, 2883.
(30) Carenco, S.; Le Goff, X. F.; Shi, J.; Roiban, L.; Ersen, O.;
Boissiere, C.; Sanchez, C.; Mezailles, N. Chem. Mater. 2011, 23, 2270.
(31) Muthuswamy, E.; Savithra, G. H. L.; Brock, S. L. ACS Nano
2011, 5, 2402.
(32) Seo, W. S.; Shim, J. H.; Oh, S. J.; Lee, E. K.; Hur, N. H.; Park, J.
T. J. Am. Chem. Soc. 2005, 127, 6188.
(33) Hou, Y.; Xu, Z.; Sun, S. Angew. Chem., Int. Ed. 2007, 46, 6329.
(34) Cozzoli, P. D.; Snoeck, E.; Garcia, M. A.; Giannini, C.;
Guagliardi, A.; Cervellino, A.; Gozzo, F.; Hernando, A.; Achterhold,
K.; Ciobanu, N.; Parak, F. G.; Cingolani, R.; Manna, L. Nano Lett.
2006, 6, 1966.
(35) Xu, Z.; Shen, C.; Hou, Y.; Gao, H.; Sun, S. Chem. Mater. 2009,
21, 1778.
(36) Sun, S.; Zeng, H. J. Am. Chem. Soc. 2002, 124, 8204.
(37) Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.;
Wang, S. X.; Li, G. J. Am. Chem. Soc. 2004, 126, 273.
(38) Zhang, L.; Dou, Y. H.; Gu, H. C. J. Cryst. Growth 2006, 296,
221.
(39) Zhang, L.; Wu, J.; Liao, H.; Hou, Y.; Gao, S. Chem. Commun.
2009, 4378.
(40) Seo, W. S.; Jo, H. H.; Lee, K.; Kim, B.; Oh, S. J.; Park, J. T.
Angew. Chem., Int. Ed. 2004, 43, 1115.
(41) Park, J.; Kang, E.; Bae, C. J.; Park, J. G.; Noh, H. J.; Kim, J. Y.;
Park, J. H.; Park, H. M.; Hyeon, T. J. Phys. Chem. B 2004, 108, 13594.
(42) Chaubey, G. S.; Barcena, C.; Poudyal, N.; Rong, C.; Gao, J.;
Sun, S.; Liu, J. P. J. Am. Chem. Soc. 2007, 129, 7214.
(43) Desvaux, C.; Dumestre, F.; Amiens, C.; Respaud, M.; Lecante,
P.; Snoeck, E.; Fejes, P.; Renaud, P.; Chaudret, B. J. Mater. Chem.
2009, 19, 3268.
(44) Wang, C.; Peng, S.; Lacroix, L. M.; Sun, S. Nano Res. 2009, 2,
380.
(45) Chen, M.; Nikles, D. E. Nano Lett. 2002, 2, 211.
(46) Ye, E.; Zhang, S. Y.; Lim, S. H.; Bosman, M.; Zhang, Z.; Win, K.
Y.; Han, M. Y. Chem.Eur. J. 2011, 17, 5982.
(47) Sharma, S.; Gajbhiye, N. S.; Ningthoujam, R. S. J. Colloid
Interface Sci. 2010, 351, 323.

decomposed to produce nanoparticles, were clearly explained.


The composition, size, and shape of the obtained nanocrystals
can be tuned by careful choice of additional reaction
parameters, depending on the system under study. Concerning
commercially available OAm, a future goal might be the
improvement of its purity, as up to now OAm is not available at
purities higher than 8590% (or 70%, for the technical grade
reagent). A decrease in the level of impurities could possibly
ensure an even better reproducibility and facilitate scaling the
synthetic reaction protocols where OAm is involved up to
industrial scale, provided that its cost would remain reasonably
low. In this way, the number of applications for the OAmcapped nanomaterials, though already remarkable, might be
greatly increased.

AUTHOR INFORMATION

Corresponding Author

*E-mail: mourdik@uvigo.es (S.M.); llizmarzan@cicbiomagune.


es (L.M.L.-M.).
Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
This work was implemented within the framework of the
Action Supporting Postdoctoral Researchers of the Operational Program Education and Lifelong Learning (Actions
Beneciary: General Secretariat for Research and Technology
of Greece) and is conanced by the European Social Fund
(ESF) and the Greek State [Project code PE4(1546)]. L.M.L.M. acknowledges funding from the European Research Council
(ERC Advanced Grant 267867, PLASMAQUO). S.M. thanks
L. Polavarapu for suggesting the idea to write this Short Review
for a compound he is also familiar with.

REFERENCES

(1) Pastoriza-Santos, I.; Liz-Marzan, L. M. Adv. Funct. Mater. 2009,


19, 679.
(2) Wang, D.; Li, Y. Inorg. Chem. 2011, 50, 5196.
(3) Borges, J.; Ribeiro, J. A.; Pereira, E. M.; Carreira, C. A.; Pereira, C.
M.; Silva, F. J. Colloid Interface Sci. 2011, 358, 626.
(4) Ma, Y.; Zeng, J.; Li, W.; McKiernan, M.; Xe, Z.; Xia, Y. Adv.
Mater. 2010, 22, 1930.
(5) Shukla, N.; Liu, C.; Jones, P. M.; Weller, D. J. Magn. Magn. Mater.
2003, 266, 178.
(6) Zhang, J. L.; Srivastava, R. S.; Misra, R. D. K. Langmuir 2007, 23,
6342.
(7) Nam, K. M.; Shim, J. H.; Ki, H.; Choi, S. I.; Lee, G.; Jang, J. K.;
Jo, Y.; Jung, M. H.; Song, H.; Park, J. T. Angew. Chem., Int. Ed. 2008,
47, 9504.
(8) Mourdikoudis, S.; Simeonidis, K.; Tsiaoussis, I.; DendrinouSamara, C.; Angelakeris, M.; Kalogirou, O. J. Nanopart. Res. 2009, 11,
1477.
(9) Dumestre, F.; Chaudret, B.; Amiens, C.; Fromen, M. C.;
Casanove, M. J.; Renaud, P.; Zurcher, P. Angew. Chem., Int. Ed. 2002,
41, 4286.
(10) Ciuculescu, D.; Dumestre, F.; Comesana-Hermo, M.; Chaudret,
B.; Spasova, M.; Farle, M.; Amiens, C. Chem. Mater. 2009, 21, 3987.
(11) Kura, H.; Takahashi, M.; Ogawa, T. J. Phys. Chem. C 2010, 114,
5835.
(12) Farrell, D.; Majetich, S. A.; Wilcoxon, J. P. J. Phys. Chem. B 2003,
107, 11022.
(13) Somaskandan, K.; Veres, T.; Niewczas, M.; Simard, B. New J.
Chem. 2008, 32, 201.
(14) Peng, S.; Wang, C.; Sun, S. J. Am. Chem. Soc. 2006, 128, 10676.
1474

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

(48) Yamauchi, T.; Tsukuhara, Y.; Yamada, K.; Sakata, T.; Wada, Y.
Chem. Mater. 2011, 23, 75.
(49) Tzitzios, V.; Niarchos, D.; Gjoka, M.; Boukos, N.; Petridis, D. J.
Am. Chem. Soc. 2005, 127, 13756.
(50) Mourdikoudis, S.; Simeonidis, K.; Gloystein, K.; Angelakeris,
M.; Dendrinou-Samara, C.; Tsiaoussis, I.; Kalogirou, O. J. Magn. Magn.
Mater. 2009, 321, 3120.
(51) Chen, M.; Nikles, D. E. J. Appl. Phys. 2002, 91, 8477.
(52) Wang, C.; Hou, Y.; Kim, J.; Sun, S. Angew. Chem., Int. Ed. 2007,
46, 6333.
(53) Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Science
2000, 287, 1989.
(54) Chen, M.; Liu, J. P.; Sun, S. J. Am. Chem. Soc. 2004, 126, 8394.
(55) Chen, M.; Kim, J.; Liu, J. P.; Fan, H.; Sun, S. J. Am. Chem. Soc.
2006, 128, 7132.
(56) Acet, M.; Mayer, C.; Muth, O.; Terheiden, A.; Dyker, G. J. Cryst.
Growth 2005, 285, 365.
(57) Liao, H. G.; Cui, L.; Whitelam, S.; Zheng, H. Science 2012, 336,
1011.
(58) Sun, S.; Anders, S.; Hamann, H. F.; Thiele, J. U.; Baglin, J. E. E.;
Thomson, T.; Fullerton, E. E.; Murray, C. B.; Terris, B. D. J. Am.
Chem. Soc. 2002, 124, 2884.
(59) Sun, S.; Anders, S.; Thomson, T.; Baglin, J. E. E.; Toney, M. F.;
Hamann, H. F.; Murray, C. B.; Terris, B. D. J. Phys. Chem. B 2003, 107,
5419.
(60) Teng, X.; Yang, H. J. Am. Chem. Soc. 2003, 125, 14559.
(61) De la Presa, P.; Rueda, T.; Morales, M. P.; Hernaldo, A. IEEE
Trans. Magn. 2008, 44, 2816.
(62) Chen, Y.; Luo, X.; Hue, G. H.; Luo, X.; Peng, D. L. Mater. Chem.
Phys. 2009, 113, 412.
(63) Li, Y.; Zhang, X. L.; Qiu, R.; Qiao, R.; Kang, Y. S. J. Phys. Chem.
C 2007, 111, 10747.
(64) Zhang, J.; Yang, H.; Fang, J.; Zou, S. Nano Lett. 2010, 10, 638.
(65) Guo, H.; Chen, Y.; Ping, H.; Wang, L.; Peng, D. L. J. Mater.
Chem. 2012, 22, 8336.
(66) Lee, D. C.; Ghezelbash, A.; Stowell, C. A.; Korgel, B. A. J. Phys.
Chem. B 2006, 110, 20906.
(67) Kang, Y.; Murray, C. B. J. Am. Chem. Soc. 2010, 132, 7568.
(68) Wang, C.; Daimon, H.; Onodera, T.; Koda, T.; Sun, S. Angew.
Chem., Int. Ed. 2008, 47, 3588.
(69) Wang, C.; Daimon, H.; Lee, Y.; Kim, J.; Sun, S. J. Am. Chem. Soc.
2007, 129, 6974.
(70) Wu, B.; Zheng, N.; Fu, G. Chem. Commun. 2011, 47, 1039.
(71) Kang, Y.; Ye, X.; Murray, C. B. Angew. Chem., Int. Ed. 2010, 49,
6156.
(72) Zhong, X.; Feng, Y.; Lieberwirth, I.; Knoll, W. Chem. Mater.
2006, 18, 2468.
(73) Mankin, M. N.; Mazumder, V.; Sun, S. Chem. Mater. 2011, 23,
132.
(74) Zhang, H. T.; Ding, J.; Chow, C. M. Langmuir 2008, 24, 375.
(75) Lacroix, L. M.; Gatel, C.; Arenal, R.; Garcia, C.; Lachaize, S.;
Blon, T.; Warot-Fonrose, B.; Snoeck, E.; Chaudret, B.; Viau, G. Angew.
Chem., Int. Ed. 2012, 51, 4690.
(76) Fang, Z.; Zhang, Y.; Du, F.; Zhong, X. Nano Res. 2008, 1, 249.
(77) Lim, S. I.; Jimenez, I. O.; Varon, M.; Casals, E.; Arbiol, J.;
Puntes, V. Nano Lett. 2010, 10, 964.
(78) Cheong, S.; Watt, J.; Ingham, B.; Toney, M. F.; Tilley, R. D. J.
Am. Chem. Soc. 2009, 131, 14590.
(79) Ren, J.; Tilley, R. D. J. Am. Chem. Soc. 2007, 129, 3287.
(80) Ren, J.; Tilley, R. D. Small 2007, 3, 1508.
(81) Yang, H.; Zhang, J.; Sun, K.; Zou, S.; Fang, J. Angew. Chem., Int.
Ed. 2010, 49, 6848.
(82) Lim, S. I.; Varon, M.; Ojea-Jimenez, I.; Arbiol, J.; Puntes, V.
Chem. Mater. 2010, 22, 4495.
(83) Peng, Z.; Yang, H. J. Am. Chem. Soc. 2009, 131, 7542.
(84) Wu, J.; Gross, A.; Yang, H. Nano Lett. 2011, 11, 798.
(85) Zhou, S.; Jackson, G. S.; Eichhorn, B. Adv. Funct. Mater. 2007,
17, 3099.

(86) Lu, Y. C.; Xu, Z.; Gasteiger, H. A.; Chen, S.; Hamad-Schifferli,
K.; Shao-Hom, Y. J. Am. Chem. Soc. 2010, 132, 12170.
(87) Zhou, S.; McIlwrath, K.; Jackson, G.; Eichhorn, B. J. Am. Chem.
Soc. 2006, 128, 1780.
(88) Wang, C.; Tian, W.; Ding, Y.; Ma, Y.; Wang, Z. L.; Markovic, N.
M.; Stamenkovic, V. R.; Daimon, H.; Sun, S. J. Am. Chem. Soc. 2010,
132, 6524.
(89) Peng, Z.; You, H.; Yang, H. ACS Nano 2010, 4, 1501.
(90) Liu, Q.; Yan, Z.; Henderson, N. L.; Bauer, J. C.; Goodman, D.
W.; Batteas, J. D.; Schaak, R. E. J. Am. Chem. Soc. 2009, 131, 5720.
(91) Xu, D.; Bliznakov, S.; Lu, Z.; Fang, J.; Dimitrov, N. Angew.
Chem., Int. Ed. 2010, 49, 1282.
(92) Mazumder, V.; Sun, S. J. Am. Chem. Soc. 2009, 131, 4588.
(93) Sato, R.; Kanehara, M.; Teranishi, T. Small 2011, 7, 469.
(94) Niu, Z.; Peng, Q.; Gong, M.; Rong, H.; Li, Y. Angew. Chem., Int.
Ed. 2011, 50, 6315.
(95) Ortiz, N.; Skrabalak, S. E. Angew. Chem., Int. Ed. 2012, 51,
11757.
(96) Hu, B.; Ding, K.; Wu, T.; Zhou, X.; Fan, H.; Jiang, T.; Wang, Q.;
Han, B. Chem. Commun. 2010, 46, 8552.
(97) Metin, O.; Duman, S.; Dinc, M.; Ozkar, S. J. Phys. Chem. C
2011, 115, 10736.
(98) Watt, J.; Young, N.; Haigh, S.; Kirkland, A.; Tilley, R. D. Adv.
Mater. 2009, 21, 2288.
(99) Watt, J.; Cheong, S.; Toney, M. F.; Ingham, B.; Cookson, J.;
Bishop, P. T.; Tilley, R. D. ACS Nano 2010, 4, 396.
(100) Kim, S. W.; Park, J.; Jang, Y.; Chung, Y.; Hwang, S.; Hyeon, T.
Nano Lett. 2003, 3, 1289.
(101) Yang, Z.; Klabunde, K. J. J. Organomet. Chem. 2009, 694, 1016.
(102) Park, K. H.; Jang, K.; Kim, H. J.; Son, S. U. Angew. Chem., Int.
Ed. 2007, 46, 1152.
(103) Stowell, C. A.; Korgel, B. A. Nano Lett. 2005, 5, 1203.
(104) Pazos-Perez, N.; Baranov, D.; Irsen, S.; Hilgendorff, M.; LizMarzan, L. M.; Giersig, M. Langmuir 2008, 24, 9855.
(105) Kura, H.; Ogawa, T. J. Appl. Phys. 2010, 107, 074310.
(106) Huo, Z.; Tsung, C.; Huang, W.; Zhang, X.; Yang, P. Nano Lett.
2008, 8, 2041.
(107) Lu, X.; Yavuz, M. S.; Tuan, H. Y.; Korgel, B. A.; Xia, Y. J. Am.
Chem. Soc. 2008, 130, 8900.
(108) Li, Z.; Tao, J.; Lu, X.; Zhu, Y.; Xia, Y. Nano Lett. 2008, 8, 3052.
(109) Li, Z.; Li, W.; Camargo, P. H. C.; Xia, Y. Angew. Chem., Int. Ed.
2008, 47, 9653.
(110) Feng, H.; Yang, Y.; You, Y.; Li, G.; Guo, J.; Yu, T.; Shen, X.;
Wu, T.; Xing, B. Chem. Commun. 2009, 1984.
(111) Halder, A.; Ravishankar, N. Adv. Mater. 2007, 19, 1854.
(112) Polavarapu, L.; Venkatram, N.; Ji, W.; Xu, Q. H. ACS Appl.
Mater. Interfaces 2009, 1, 2298.
(113) Wang, C.; Hu, Y.; Lieber, C. M.; Sun, S. J. Am. Chem. Soc.
2008, 130, 8902.
(114) Polavarapu, L.; Xu, Q. H. Nanotechnology 2009, 20, 185606.
(115) Peng, S.; Lee, Y.; Wang, C.; Yin, H.; Dai, S.; Sun, S. Nano Res.
2008, 1, 229.
(116) De la Presa, P.; Multigner, M.; De la Venta, J.; Garcia, M. A.;
Ruiz-Gonzalez, M. L. J. Appl. Phys. 2006, 100, 123915.
(117) Shen, C.; Hui, C.; Yang, T.; Xiao, C.; Tian, J.; Bao, L.; Chen,
S.; Ding, H.; Gao, H. Chem. Mater. 2008, 20, 6939.
(118) Chen, M.; Feng, Y. G.; Wang, X.; Li, T. C.; Zhang, J. Y.; Qian,
D. J. Langmuir 2007, 23, 5296.
(119) Mott, D.; Galkowski, J.; Wang, L.; Luo, J.; Zhong, C. J.
Langmuir 2007, 23, 5740.
(120) Son, S. U.; Park, I. K.; Park, J.; Hyeon, T. Chem. Commun.
2004, 778.
(121) Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Chem.
Commun. 2012, 48, 5626.
(122) Wang, C.; Yin, H.; Chan, R.; Peng, S.; Dai, S.; Sun, S. Chem.
Mater. 2009, 21, 433.
(123) Hou, X.; Zhang, X.; Yang, W.; Liu, Y.; Zhai, X. Mater. Res. Bull.
2012, 47, 2579.
1475

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

Chemistry of Materials

Review

(124) Li, N.; Zhang, X.; Chen, S.; Hou, X.; Liu, Y.; Zhai, X. Mater.
Sci. Eng., B. 2011, 176, 688.
(125) Yang, H.; Fan, N.; Luan, W.; Tu, S. Nanoscale Res. Lett. 2009,
4, 344.
(126) Zhang, H. T.; Wu, G.; Chen, X. H. Langmuir 2005, 21, 4281.
(127) Wang, Y.; Tang, A.; Li, K.; Yang, C.; Wang, M.; Ye, H.; Hou,
Y.; Teng, F. Langmuir 2012, 28, 16436.
(128) Cademartiri, L.; Bertolotti, J.; Sapienza, R.; Wiersma, D. S.;
Freymann, G.; Ozin, G. A. J. Phys. Chem. B 2006, 110, 671.
(129) Zhang, J.; Kumbhar, A.; He, J.; Das, N. C.; Yang, K.; Wang, J.
Q.; Wang, H.; Stokes, K. L.; Fang, J. J. Am. Chem. Soc. 2008, 130,
15203.
(130) Zhai, X.; Zhang, X.; Chen, S.; Yang, W.; Gong, Z. Colloids Surf.,
A 2012, 409, 126.
(131) Ghezelbash, A.; Korgel, B. A. Langmuir 2005, 21, 9451.
(132) Joo, J.; Na, H. B.; Yu, T.; Yu, J. H.; Kim, Y. W.; Wu, F.; Zhang,
J. Z.; Hyeon, T. J. Am. Chem. Soc. 2003, 125, 11100.
(133) Thomson, J. W.; Nagashima, K.; Macdonald, P. M.; Ozin, G. A.
J. Am. Chem. Soc. 2011, 133, 5036.
(134) Shavel, A.; Cadavid, D.; Ibanez, M.; Carrete, A.; Cabot, A. J.
Am. Chem. Soc. 2012, 134, 1438.
(135) Ou, K. L.; Fan, J. C.; Chen, J. K.; Huang, C. C.; Chen, L. Y.;
Ho, J. H.; Chang, J. Y. J. Mater. Chem. 2012, 22, 14667.
(136) Pein, A.; Baghbanzadeh, M.; Rath, T.; Haas, W.; Maier, E.;
Amenitsch, H.; Hofer, F.; Kappe, C. O.; Trimmel, G. Inorg. Chem.
2011, 50, 193.
(137) Stolle, C. J.; Panthani, M. G.; Harvey, T. B.; Akhavan, V. A.;
Korgel, B. A. ACS Appl. Mater. Interfaces 2012, 4, 2757.
(138) Liu, Y.; Yao, D.; Shen, L.; Zhang, H.; Zhang, X.; Yang, B. J. Am.
Chem. Sec. 2012, 134, 7207.
(139) Gerung, H.; Boyle, T. J.; Tribby, L. J.; Bunge, S. D.; Brinker, C.
J.; Han, S. M. J. Am. Chem. Soc. 2006, 128, 5244.
(140) Wu, H. P.; Ge, M. Y.; Wao, C. W.; Wang, Y. W.; Zeng, Y. W.;
Wang, L. N.; Zhang, G. Q.; Jiang, J. Z. Nanotechnology 2006, 17, 5339.
(141) Seo, W. S.; Jo, H. H.; Lee, K.; Park, J. T. Adv. Mater. 2003, 15,
795.
(142) Liu, Q.; Lu, W.; Ma, A.; Tang, J.; Lin, J.; Fang, J. J. Am. Chem.
Soc. 2005, 127, 5276.
(143) Selishcheva, E.; Parisi, J.; Kolny-Olesiak, J. J. Nanopart. Res.
2012, 14, 711.
(144) Zhang, Z.; Liu, S.; Chow, S.; Han, M. Y. Langmuir 2006, 22,
6335.
(145) Choi, S. H.; Na, H. B.; Park, Y. I.; An, K.; Kwon, S. G.; Jang, Y.;
Park, M.; Moon, J.; Son, J. S.; Song, I. C.; Moon, W. K.; Hyeon, T. J.
Am. Chem. Soc. 2008, 130, 15573.
(146) Peng, S.; Lei, C.; Ren, Y.; Cook, R. E.; Sun, Y. Angew. Chem.,
Int. Ed. 2011, 50, 3158.
(147) Xu, Z.; Hou, Y.; Sun, S. J. Am. Chem. Soc. 2007, 129, 8698.
(148) Yu, H.; Chen, M.; Rice, P. M.; Wang, S. X.; White, R. L.; Sun,
S. Nano Lett. 2005, 5, 379.
(149) Palchoudhury, S.; Xu, Y.; An, W.; Turner, C. H.; Bao, Y. J.
Appl. Phys. 2010, 107, 09B311.
(150) Schladt, T. D.; Graf, T.; Kohler, O.; Bauer, H.; Dietzsch, M.;
Mertins, J.; Branscheid, R.; Kolb, U.; Tremel, W. Chem. Mater. 2012,
24, 525.
(151) Si, R.; Zhang, Y. W.; You, L. P.; Yan, C. H. Angew. Chem., Int.
Ed. 2005, 44, 3256.
(152) Yu, T.; Joo, J.; Park, Y. I.; Hyeon, T. Angew. Chem., Int. Ed.
2005, 44, 7411.
(153) Cao, Y. C. J. Am. Chem. Soc. 2004, 126, 7456.
(154) Ostrowski, A. D.; Chan, E. M.; Gargas, D. J.; Katz, E. M.; Han,
G.; Schuck, P. J.; Milliron, D. J.; Cohen, B. E. ACS Nano 2012, 6, 2686.
(155) Bu, W.; Chen, Z.; Chen, F.; Shi, J. J. Phys. Chem. C 2009, 113,
12176.
(156) Wu, H.; Yang, Y.; Cao, Y. C. J. Am. Chem. Soc. 2006, 128,
16522.
(157) Matsushita, T.; Iwamoto, T.; Inokuchi, M.; Toshima, N.
Nanotechnology 2010, 21, 095603.

(158) Seo, J.; Jun, Y.; Ko, S. J.; Cheon, J. J. Phys. Chem. B 2005, 109,
5389.
(159) Du, Y.; Yin, Z.; Zhu, J.; Huang, X.; Wu, X. J.; Zeng, Z.; Yan, Q.;
Zhang, H. Nat. Commun. 2012, 3, 1177.
(160) Lu, X.; Zhuang, Z.; Peng, Q.; Li, Y. CrystEngComm 2011, 13,
4039.
(161) Lou, W.; Chen, M.; Wang, X.; Liu, W. Chem. Mater. 2007, 19,
872.
(162) Son, Y. S.; Choi, M. K.; Han, M. K.; Park, K.; Kim, J. Y.; Lim, S.
J.; Oh, M.; Kuk, Y.; Park, C.; Kim, S. J.; Hyeon, T. Nano Lett. 2012, 12,
640.
(163) Hiramatsu, H.; Osterloh, F. Chem. Mater. 2004, 16, 2509.
(164) Dinh, C. T.; Nguyen, T. D.; Kleitz, F.; Do, T. O. Chem.
Commun. 2011, 47, 7797.
(165) Yu, T.; Park, Y. I.; Kang, M. C.; Joo, J.; Park, J. K.; Won, H. Y.;
Kim, J. J.; Hyeon, T. Eur. J. Inorg. Chem. 2008, 855.
(166) Panda, A. B.; Glaspell, G.; El-Shall, M. S. J. Phys. Chem. C 2007,
111, 1861.
(167) He, J.; Kanjanaboos, P.; Frazer, N. L.; Weis, A.; Lin, X. M.;
Jaeger, H. M. Small 2010, 6, 1449.
(168) Shi, R.; Liu, X.; Gao, G.; Yi, R.; Qiu, G. J. Alloys Compd. 2009,
485, 548.
(169) Yamada, M.; Okumura, S.; Takajashi, K. J. Phys. Chem. Lett.
2010, 1, 2042.
(170) Chakravarthi, D. K.; Khabashesku, V. N.; Vaidyanathan, R.;
Blaine, J.; Yarlagadda, S.; Roseman, D.; Zeng, Q.; Barrera, E. V. Adv.
Funct. Mater. 2011, 21, 2527.

1476

dx.doi.org/10.1021/cm4000476 | Chem. Mater. 2013, 25, 14651476

You might also like