You are on page 1of 15

JOURNAL OF NANOSCIENCE LETTERS

72

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.



Atomically precise nanoparticles: A new frontier in nanoscience

Rongchao Jin
*
, Huifeng Qian, Yan Zhu, Anindita Das

Department of Chemistry, Carnegie Mellon University, Pittsburgh, Pennsylvania 15213, USA

*
Author for correspondence: Rongchao Jin, email: rongchao@andrew.cmu.edu
Received 26 Jan 2011; Accepted 5 Mar 2011; Available Online 20 Apr 2011


Abstract

Controlling nanoparticles with atomic precision has long been highly sought in nanoscience and is
crucially important in order to fully understand the physicochemical properties of nanoparticles, in particular those
surface-related phenomena, such as fluorescence blinking in quantum dots and catalysis of metal nanoparticles.
This short review article focuses on atomically precise metal nanoparticles that possess distinct quantum size
effects in the size range from about ten to a few hundred atoms in the particle (often called nanoclusters). We use
thiolate-protected gold nanoclusters (denoted as Au
n
(SR)
m
, where n=number of gold atoms) as an example in our
discussions. Significant advances have been achieved recently in synthesizing well defined Au
n
(SR)
m

nanoclusters. Interestingly, the prominent surface plasmon resonance of metallic Au nanoparticles disappears in
quantum-sized Au
n
(SR)
m
nanoclusters; the latter instead exhibit discrete electronic transitions. This new class of
nanomaterial holds great potential in a number of applications, particularly in catalysis, optics, and
nanoelectronics. A grand task yet to accomplish is to study the evolution from the molecule-like cluster state to the
metallic state. Understanding how the material properties of nanoclusters evolve from the quantum confinement
state to the plasmonic state is of paramount importance to both fundamental science and technological
applications, and will impact a number of fields, including chemistry, materials science, and condensed matter
physics.

Keywords: Gold; Nanoclusters; Quantum size effects; Atomic precision

1. Introduction

A variety of nanoparticles have been
extensively pursued in current nanoscience
research, including noble metal nanoparticles,
semiconductor quantum dots, metal and
semiconductor nanowires, carbon nanotubes,
fullerenes, graphene, polymer nanostructures,
and so on. All of these nanoparticle materials
constitute the central player in nanoscience and
nanotechnology. They hold great potential in a
wide range of technological applications, such as
catalysis, photovoltaics, electronics, biological
labelling, biomedicine, etc. This surge in new
applications of nanoparticle materials has led to
major efforts, driven by both experimental and
theoretical works, to understand the fundamental
properties of nanoparticles. To achieve this goal,
it is of paramount importance to obtain well-
defined nanoparticles for in-depth studies of
their new science.
Although significant progress has been
made in the synthesis of many kinds of
nanoparticles with control over size, shape and
composition, in particular metal nanoparticles
[1], such nanoparticles are still more or less
heterogeneous in terms of size and/or shape
uniformity. For example, the best quality
quantum dots (QDs) that researchers can
currently prepare still have a size distribution of
about 5% [2]. This heterogeneity is seemingly
small, but largely affects their material
properties; for example, QDs with a ~5% size
distribution show a distinctive optical absorption
spectrum composed of step-like multiple bands
(i.e., indicating a narrow size distribution),
however, careful studies reveal that the ensemble
spectrum is indeed contributed by numerous
individual sizes that are slightly different from
each other in optical properties. This is due to
the extreme sensitivity of quantum confinement
effects of QDs to particle size. Therefore,
atomically precise QDs are highly desirable,
though very difficult to attain. As for metal
nanoparticles, it is well known that they give rise
to surface plasmon resonance (SPR), which
accounts for the absorption, scattering, surface-
enhanced Raman scattering, and many other
important properties [3]. Since the SPR
properties of metal nanoparticles are not as
sensitive to particle size as are QDs, one would
argue that size control at the nanometer precision
is already sufficient and an atomic level of
control is not necessary. However, when metal
nanoparticles enter the quantum size regime
(typically <2 nm), these ultrasmall nanoparticles
(often called nanoclusters) indeed become
extremely sensitive to particle size [4-6], even
JOURNAL OF NANOSCIENCE LETTERS

73

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

more than QDs. Therefore, when semiconductor
and metal nanoparticles enter the quantum size
regime, size control with a nanometer precision
(e.g. a fraction of nm) is not sufficient; instead,
an atomic level of control should be sought.
Atomically precise nanoparticles are
critically important if one wants to understand
the surface-related properties; for examples, 1)
how do surface ligands protect and bind to
nanoparticles? 2) How does the surface structure
of nanoparticles affect the fluorescence blinking
in QDs? 3) How are molecules activated on
nanoparticle surfaces in a catalytic reaction? To
address all these fundamentally important
questions and many others, it is of paramount
importance to pursue atomically precise
nanoparticles for ultimate determination of their
total structure (i.e. not only the core structure,
but the surface structure as well).
This mini-review will focus on the
discussion of atomically precise metal
nanoparticles only. We choose gold as a model
system partly because it has fascinating optical
properties, which have drawn significant
attention since ancient times. Major advances
have been achieved in recent years in preparing
atomically precise gold nanoclusters. This new
class of nanomaterail is anticipated to find many
important applications in future studies.

2. Metal nanoparticles: From classic
confinement to quantum confinement

The optical properties (e.g. absorption,
scattering, fluorescence) of nanoparticles are
particularly interesting, as these properties
reflect the intrinsic electronic structure of
nanoparticles. Here, we provide a short
discussion on the evolution of optical properties
of metal nanoparticles from bulk to the
nanometer scale (typically < 100 nm or so). We
use gold as a pedagogical example, for that
tremendous research has been done on gold
nanoparticles. It is of particular interest to find
that metal nanoparticles experience two major
types of size effects with shrinking size, that is,
classical and quantum confinement effects.
Below we discuss the intriguing size evolution
behavior.
For bulk gold, it absorbs blue and part
of green light (< 540 nm), but strongly reflects
yellow and red light (Figure 1A), hence, bulk
gold appears yellowish red (a golden color) due
to optical reflection. Light can only penetrate a
limited depth of bulk gold (e.g. skin depth of ~30
nm for visible light). The short-wavelength
absorption (< 540 nm) is caused by interband
(dsp) electronic transitions, and reflection
spanning from green up to radio wavelengths is
due to free electron behavior of bulk gold.
When bulk gold is divided into small
particles, in particular down to the nanometer
scale (of the order of ~100 nm), significant
changes occur to the optical properties. As
shown in Figure 1B, colloids of ~520 nm
particles (e.g. spherical) display a strong
resonant absorption band in green (
peak
~520
nm), which is not seen in the optical spectrum of
bulk gold, hence, it is a new excitation mode
unique to nanoparticles (Figure 1C), often called
surface plasmon resonance (SPR). The nature of
SPR is due to a collective excitation of free
electrons in the particle upon absorbing a
photon. This collective excitation mode of
conduction electrons also accounts for the
intense optical scattering of metal nanoparticles
as well as the surface-enhanced Raman
scattering properties. For instance, the SPR
excitation in Au nanoparticles imparts the
colloids of diameter (D)~50 nm a beautiful
magenta color when viewed in transmission, but
a green color when viewed at a 90
o
geometry
(e.g., in dark field scattering experiments). Both
the absorption and scattering colors are size
tunable since the SPR band is size dependent;
the solution color of Au nanoparticles changes
from violet (D~100 nm,
peak
~570 nm) to
purplish red (D~50 nm,
peak
~540 nm) to wine
red (D~10 nm,
peak
~520 nm), whereas the
scattering color changes from orange (D~100
nm) to green (D~50 nm); smaller nanoparticles


Figure 1. (A) Reflectance spectrum of bulk gold, (B) Extinction spectra of gold colloids (3-100 nm
diameter), (C) Surface plasmon excitation (e.g. dipole mode) in metal nanoparticles.
JOURNAL OF NANOSCIENCE LETTERS

74

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

(< 30 nm) are weak in scattering. In the last
decade, there have been significant advances in
preparing uniform Au nanoparticles and
controlling particle shape. For non-spherical
metal nanoparticles, they may exhibit two or
multiple SPR bands depending on the particle
shape [3,7]. Beside gold, much effort has also
been extended into other precious metals, such
as Ag [8]. By tailoring the shape and size of such
metal nanoparticles, almost any color in the
visible spectrum and technical colors in the
near-infrared spectral region can be produced
through optical absorption or scattering [9]; note
that human eyes cannot sense near-infrared
wavelengths, hence, technical colors only.
For all of those plasmonic metal
nanoparticles, they are metallic and inherit the
electronic structure of bulk metal, that is, the free
electron picture is applicable to plasmonic
nanoparticles, even down to D~5 nm, although a
modification in free electron mean path may be
needed. Then, what causes the unique SPR mode
in nanoparticles given their electronic structure
is similar to that of the bulk metal? This is
indeed due to classical confinement of electrons
in the particle. Bulk Au shows very broad
reflection at wavelengths above the interband
threshold / bulk plasmon frequency (e.g. <2.5 eV
for gold), whereas Au nanoparticles (say,
spherical ones) exhibit an intense band at 520-
570 nm wavelength due to the fact that the
optical excitation mode in small particles is
mainly dipolar in nature. The dipole mode takes
the majority of weight of plasmon excitation,
hence, a strong resonance band is seen. With
increasing particle size, higher-order plasmon
modes such as quadruple, octupole, and even
higher ones will come into play and partition the
oscillator strength, resulting in significant
broadening of the overall plasmon band, and
eventually the optical properties of very large
particles (e.g. > micron) will resemble bulk
metal. Thus, the distinct SPR (e.g. dipole mode)
is a size effect of conduction electron confined in
the nanoparticle. The new SPR properties can be
well modeled by classical electrodynamics, even
by electrostatics in simplified scenarios. In other
words, the free electron picture of bulk metals is
nicely retained in not-too-small nanoparticles.
Therefore, we may regard the surface plasmon
mode as a classical size effect, for that the
underlying physics is essentially handled by
classical physics and no quantum physics is
needed for particles in this size regime.
As the size of gold nanoparticles
shrinks below about 2 nm (diameter), significant
and intriguing changes occur to such ultrasmall
nanoparticles; for instances, the SPR band
disappears and step-like multibands are emerged
[10-12]. The discrete absorption bands [12], are
due to quantum size effects. If one estimates the
intrinsic characteristic length scales of
conduction electrons, one of such length scales is
the de Broglie wavelength; for free electrons in
metal, this length scale is about 1 nm. When
particle size approaches this length scale,
valence electrons in the particle will feel strong
confinement. Such effects bear quantum physics
character, hence, only at the bottom of the
nanoscale metal nanoparticles start to exhibit
quantum confinement effects. For comparison,
semiconductor nanoparticles start to show the
quantum size effects at a much larger size
threshold (e.g. ~12 nm exciton diameter for
CdSe nanoparticles). For quantum sized metal
nanoparticles, the disappearance of the distinct
SPR modes is in striking contrast with their
larger counterparts metallic nanoparticles that
exhibit intense SPR bands. Quantum
confinement effects make metal nanoclusters no
longer metallic (i.e. meaning quasi-continuous
electron energy levels with spacing <<k
B
T).
Indeed, nanoclusters manifest themselves a
significantly quantized, discrete electronic
structure, which is in close resemblance of atoms
or molecules. Overall, ultrasmall metal
nanoparticles are in the quantum size regime,
and quantum mechanical treatment is essential to
understand the fundamental properties of
nanoclusters.
Due to the extreme sensitivity of
quantum confinement effects to particle size,
control over metal nanocluster size with a
nanometer precision (e.g. a fraction of nm) is
not sufficient, rather, an atomic level of control
is of necessity in order to observe distinctive
quantum size effects. Therefore, to precisely
control the number of metal atoms in the particle
is essential for metal nanoclusters. Such
atomically precise metal nanoclusters are
accordingly represented with a formula (M
n
L
m
,
where n is the number of metal atoms, and L
represents ligands), much like the case of
molecules.
Overall, metal nanoclusters lie in
between two types of chemistry (Scheme 1), [4]
i.e. molecular chemistry which handles small
molecules (e.g. organometallic compounds of a
few angstroms in size) and nanocrystal
chemistry which focuses on crystalline
nanoparticles (typically > 2 nm). There is
tremendous new science to dig out in the narrow
size range from subnanometer to ~2 nm [13-20].
Many transitions in terms of material properties
of metal nanoparticles occur over this size
regime. To understand the new science of metal
nanoclusters, it is of paramount importance to
first prepare atomically precise nanoclusters.
Below we discuss the chemical synthesis of gold
nanoclusters, especially thiolate-protected ones,
JOURNAL OF NANOSCIENCE LETTERS

75

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

as such nanoclusters are particularly robust and
are highly attractive for practical applications.

3. Controlling gold nanoclusters with atomic
precision

The atomic level control in nanocluster
synthesis apparently brings major difficulties to
materials chemists. Nevertheless, impressive
progress has been attained in recent years.
Solution phase synthesis of metal nanoclusters
typically involves reduction of a metal salt in
solution by a reducing agent in the presence of
appropriate protecting ligands. Upon addition of
the reducing agent, metal salt is reduced to metal
atoms (e.g. Au(0)), and coalescence of them
leads to dimers, trimers, tetramers, and so on; in
the meanwhile, ligands concurrently attach to the
growing particles. Thus, a solution process
involves two simultaneous competing processes,
i.e. particle growth and ligand passivation. Once
the particles are sufficiently passivated by
ligands, further growth is terminated or at least
retarded, and nanoclusters may be separated. In
regard to the synthesis, an important aspect is to
find appropriate protecting ligands that can
terminate the growth of particles at an early
stage before the initial nuclei (e.g. few-atom
clusters) grow into metallic nanoparticles. Such
ligands should be capable of rapidly adsorbing
onto the growing nanoclusters, terminating their
size evolution and protecting them from
aggregation or Ostwald ripening over time. To
obtain ultrasmall nanoclusters (instead of larger
metallic particles), strong passivating ligands are
often chosen, such as thiols.
The gold-thiol chemistry was initiated
in the 1980s due to the discovery of self-
assembled monolayer (SAM) of thiols on bulk
gold surfaces. Inspired by the SAM work,
researchers started to exploit thiols for the
synthesis and functionalization of gold
nanoparticles in the 1990s [21-23]. These early
seminal works have stimulated tremendous
research in gold thiolate nanoparticles. With
respect to ultrasmall gold thiolate nanoparticles,
early works by Whetten, Murray and others
could produce quite narrow size distributed Au
nanoparticles [5, 23-26], but an atomic level of
control was not achieved until recently. Below
we summarize the research progress with a focus
on the recent progress in atomically precise
Au
n
(SR)
m
nanoclusters, in particular a size-
focusing methodology recently developed by Jin
and coworkers.
To make ultrasmall Au nanoparticles
that exhibit distinct quantum confinement effect,
Whetten and others made important
modifications to the original Brust protocol,
including i) using a large excess of thiol (e.g.
thiol-to-gold ratios of 3:1 or higher), ii) using a
greater excess of the reducing agent (typically 10
equivalents of NaBH
4
per mole of gold) [23].
Under these conditions, small Au nanoparticles
(down to ~1.5 nm) could be produced, but the
as-prepared nanoparticles were quite
polydisperse. Alvarez et al. performed extensive
work on fractionation of polydisperse
Au:thiolate nanoparticles and obtained several
fractions in the 1.5-3.5 nm range (containing
some 100 to 1300 Au atoms), including species
of 93-92 k, ~57 k, 46-45 k, and 29-27 k (k =
1000 amu units, measured by laser desorption
ionization mass spectrometry, LDI-MS); note
that these core masses refer to the form of Au
n
S
m

since the S-C bonds are typically broken in LDI
analysis) [24, 27]. Although the early works did
not give rise to atomically precise nanoclusters,
the surface plasmon resonance was interestingly
found to fade away with decreasing size, which
is a manifestation of quantum size effects. At
around ~2 nm size (diameter), the SPR
completely disappeared, implying a transition
from metallic to molecular state. Subsequent
work by Whetten and coworkers produced even
smaller Au nanoclusters using improved
chemistry involving longchain thiolates (n-
C
n
H
2n+1
SH, where, n = 4, 6, 12, 18), albeit still
not atomically monodisperse, such as species
with core masses of 29-28 k (1.7 nm, ~146
atoms), 23-22 k (1.5 nm, ~101 atoms), and 15-14
k (1.3 nm, ~75 atoms), and the ever smallest
cluster 8 k (1.1 nm, ~38 atoms) [5]. Step-like


Scheme 1. Metal nanoclusters bridge organometallic complexes and nanocrystals.
JOURNAL OF NANOSCIENCE LETTERS

76

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

optical absorption was for the first time observed
in the 8 k species [5].

3.1. Synthesis of atomically precise gold
nanoclusters
In early days, the gold thiolate
nanoclusters made were hydrophobic (capped by
alkanethiolates or arylthiolates) and were only
amenable to LDI or matrix-assisted laser
desorption ionization (MALDI) MS
characterization at that time. Unfortunately, LDI
or MALDI are not good tools as they often break
up the cluster (e.g. loss of thiolate ligands,
breaking of Au-S or S-C bonds, or even loss of
gold atoms), hence, it was not possible to
determine the intact cluster mass for deducing
the cluster formula. Thus, it is highly desirable to
prepare nanoclusters amenable to electrospray
ionization (ESI) mass spectrometry
characterization. In early days, ESI-MS was
usually done for aqueous samples, although
nowadays organic soluble clusters can be
handled in ESI analysis. Compared to LDI and
MALDI, ESI is a much softer ionization
technique and allows for the determination of the
intact cluster mass, based upon which one can
deduce the exact cluster composition in
conjunction with other characterization such as
elemental analysis. To utilize ESI-MS, Schaaff
et al. explored a water-soluble tripeptide thiol,
glutathione (GSH=-Glu-Cys-Gly) for the
synthesis of water soluble gold clusters and
successfully obtained a new 5k (core mass)
species. Following polyacrylamide gel
electrophoresis (PAGE) separation of the crude
product of Au
n
(SG)
m
(a mixture), ESI-MS
successfully determined that the mass of intact
clusters to be 10.4 kDa [12]; unfortunately, the
formula was erroneously assigned to Au
28
(SG)
16

(later corrected as Au
25
(SG)
18
by Tsukuda et al
[6], vide infra). The new cluster exhibits strong
quantum confinement effects, manifested by
multiple absorption bands in the optical
spectrum. Murray and coworkers reported the
synthesis and isolation of a molecule-like cluster
initially erroneously formulated as Au
38
(SR)
24

but later corrected as Au
25
(SR)
18
[28, 29].
Tsukuda and coworkers performed a
high resolution PAGE separation of the
Au
n
(SG)
m
clusters and conducted an elegant ESI-
MS analysis of all the isolated cluster species
[6]. Very clean ESI-MS spectra were for the first
time obtained under largely improved ESI-MS
conditions (i.e. suppressing fragmentation of
clusters in ESI), indicating high purity of the as-
separated Au
n
(SG)
m
nanoclusters. Negishi et al.
identified and assigned nine distinct
nanoclusters, including Au
10
(SG)
10
, Au
15
(SG)
13
,
Au
18
(SG)
14
, Au
22
(SG)
16
, Au
22
(SG)
17
, Au
25
(SG)
18
,
Au
29
(SG)
20
, Au
33
(SG)
22
, and Au
39
(SG)
24
, Figure
2. Of note, the Au
25
(SG)
18
cluster was for the
first time unequivocally identified, and the
previous incorrect formula of Au
28
(SG)
16
was
corrected [6]. Murray et al also corrected their
earlier incorrect Au
38
(SR)
24
assignment to
Au
25
(SR)
18
[29-31]. The three nanocluster
species were finally unified as one species.
In terms of the synthesis, almost all the
early works produced only size mixed
nanoclusters, albeit pure Au
25
(SR)
18
clusters
(where, R=-C
2
H
4
Ph, -G, -C
6
H
13
, etc) were
obtained following solvent extraction or PAGE
separation. In later work, Zhu et al significantly
improved the two-phase synthetic method by
recognizing the importance of kinetic control in
the cluster synthesis, especially the Au(I)
intermediates, and for the first time obtained
highly pure Au
25
(SC
2
H
4
Ph)
18
clusters in the
crude product [32]. The high purity of the
Au
25
(SR)
18
nanoclusters was evidenced by the
distinct absorption spectrum of the crude product
without any purification or separation [32]. This
major progress permitted straightforward growth
of single crystals of the Au
25
(SC
2
H
4
Ph)
18

nanocluster for structural analysis [15, 33]. The
finding of the kinetic control conditions for high
yielding synthesis of Au
25
(SR)
18
in high purity is
interesting; this approach has been extended to
prepare various Au
25
nanoclusters with
functionalized thiols [34-36]. Moreover, new
clusters such as Au
20
(SCH
2
CH
2
Ph)
16
and
Au
24
(SCH
2
CH
2
Ph)
20
have also been made by this
method [37, 38].
Another quite useful method for
making monodisperse Au
n
(SR)
m
nanoclusters is
the thiol etching method. In early work, Whetten
et al conducted thiol etching on a mixture of 14
k and 8 k species, and found that the 8 k species
became enriched after thiol etching of the
mixture [39]. In some other cases, phosphine-
protected gold clusters were exploited to prepare
gold thiolate nanoclusters. Hutchison and
coworkers reported an improved synthesis of
phosphine-capped Au
11
clusters and ligand
exchange with thiols to make thiolated
nanoclusters [40-42]. In another report,
phosphine-capped Au
~55
clusters were converted
to thiolated Au
~75
following thiol etching with
C
6
-SH [43]. Shichibu et al. reported the
conversion of Au
11
phosphine clusters to
Au
25
(SG)
18
in the presence of excess glutathione
(GSH) [44]. They also reported another 25-atom
gold nanocluster capped by mixed phosphine
and thiolate ligands with the structure
determined to be a biicosahedral one [45]. Price
and Whetten prepared benzenethiolate-capped
8.7 kDa gold clusters following an oxidative
etching of the 22 k clusters [46]. Qian et al.
found that phosphine-capped, polydisperse Au
nanoparticles (predominantly 13.5 nm
JOURNAL OF NANOSCIENCE LETTERS

77

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

diameter) could be utilized as a common starting
material for conversion into monodisperse
[Au
25
(PPh
3
)
10
(SR)
5
Cl
2
]
2+
nanorods and
Au
25
(SR)
18
nanospheres via respective one-phase
and two-phase thiol etching [47]. This
conversion process is striking in two features:
size focusing and shape control. From these
works, one sees that the phosphine to thiol
ligand exchange often results in a core size
change. In another system, Tsunoyama et al.
used octadecanethiol (C
18
-SH) to etch pre-
formed gold clusters stabilized by poly(N-vinyl-
2-pyrrolidone) (PVP) and obtained four fractions
after recycling size exclusion chromatography
separation [48, 49]. The as-obtained Au:SC
18

clusters were determined to be 8, 11, 21, and 26
kDa by LDI mass spectrometry. This sequence
of core sizes is different from the series of 8, 14,
22, and 29 kDa obtained by direct reduction of
Au(I)SC
n
polymers by NaBH
4
[49]. Recently
they determined the number of gold atoms in a
series of PVP-stabilized gold nanoclusters by
MALDI-MS and found out discrete sizes of 35
1, 43 1, 58 1, 70 3, 107 4, 130 1, and
150 2 atoms [50].
Chaki et al. reported the synthesis of 8
k and 29 k species following a thiol etching
method, but the yield was quite low. These
clusters are assigned to be Au
38
(SC
n
)
24
and
Au
144
(SC
n
)
59
on the basis of detailed ESI-MS
analysis [51]. Quinn and coworkers also reported
the synthesis of Au
38
(SC
6
)
22
, but atomically
monodisperse nanoclusters were not obtained
[52,

53]. The assignment of the formula was
based upon LDI-MS characterization and ligand
loss was perhaps involved. Qian et al. developed
a two-step method for preparing atomically
monodisperse Au
38
(SR)
24
clusters [54]. This
method includes two main steps: i) the synthesis
of a Au
n
(SG)
m
mixture with a controlled size
distribution, ii) such Au
n
(SG)
m
clusters are then
used as the starting material for subsequent
thermal thiol etching. This method allows, for
the first time, a large-scale synthesis of
Au
38
(SC
12
)
24
nanoclusters. Following some
modifications of this method [55, 56], Qian et al
further attained the synthesis of phenylethiolate-
capped Au
38
(SC
2
H
4
Ph)
24
nanoclusters and
successfully crystallized this cluster [16].
Interestingly, the crystal structure of


Figure 2. ESI mass spectra of the fractionated Au
n
(SG)
m
clusters. Left panel: wide range ESI spectra of
species 1-9 (each showing several multiply charged peaks); Right panel: zoom-in spectra. Reproduced from
ref. 6 by permission. Copyright 2005 American Chemical Society.
JOURNAL OF NANOSCIENCE LETTERS

78

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

Au
38
(SC
2
H
4
Ph)
24
is chiral and two isomers (left-
and right-handed ones) were found. With respect
to the 29 k species, early work by Whetten and
Murray et al. determined its composition to be
Au
~140
[5, 23, 57, 58]. The precise Au
144
(SR)
59

formula reported by Chaki et al. is interesting
[51]. Qian et al. developed a two-step method
for the synthesis of atomically monodisperse
Au
144
(SR)
60
nanoclusters (where, R=-C
12
H
25
or -
CH
2
CH
2
Ph) [59, 60], note that this formula is
one-ligand different from the one reported by
Chaki et al. Murray et al obtained the same
Au
144
(SR)
60
formula in their work [61]. The one-
ligand difference between Au
144
(SR)
59
and
Au
144
(SR)
60
is not clear yet.

It is worth noting that, in the
determination of the cluster formula, ESI-MS is
an indispensable tool. Gold thiolate nanoclusters
are typically charge neutral (except the case of
Au
25
(SR)
18

). To impart charges to the clusters,


Tsukuda et al used an oxidation method [51].
Tracy et al found that adding alkali acetate can
promote cluster ionization [31]. For MALDI
analysis, Dass et al. found a good matrix
compound, trans-2-[3-(4-tert-butylphenyl)-2-
methyl-2-propenyldidene] malononitrile
(abbreviated as DCTB), and reported the first
MALDI-MS analysis of intact Au
25
(SR)
18

clusters, albeit more or less fragmentation may
still be resulted [62-64]. Using MALDI-MS,
Dass et al also identified an interesting cluster
species formulated as Au
68
(SR)
34
[65]. The
advances in mass spectrometry analysis of
clusters have significantly pushed forward the
research on gold thiolate clusters. However, the
mass determination of large Au
n
(SR)
m

nanoclusters still remains a challenge. So far,
Au
144
(SR)
60
is the largest nanocluster that has
been successful determined by ESI-MS.

3.2. Size focusing methodology for the synthesis
of Au
n
(SR)
m
nanoclusters
With respect to the synthesis of
Au
n
(SR)
m
nanoclusters, a major method is the
size focusing methodology recently established
by us on the basis of the kinetic control and thiol
etching [66]. This method has been
demonstrated to be universal in the synthesis of a
number of atomically precise Au
n
(SR)
m

nanoclusters, such as Au
25
(SR)
18
[32, 34],
Au
38
(SR)
24
[54, 55], and Au
144
(SR)
60
[59],
(where, R=CH
2
CH
2
Ph, C
12
H
25
, etc). Below, we
briefly discuss the size-focusing features in
preparing well-defined Au
n
(SR)
m
nanoclusters.
In the synthesis of the Au
25
(SR)
18

cluster, we discovered that the kinetics for the
formation of Au(I):SR intermediates is critical
for high yield synthesis of monodisperse
Au
25
(SR)
18
clusters; specifically, it was found
that control over the reaction temperature and
aggregation conditions of [Au(I):SR]
x
polymeric
intermediates can generate particular aggregates
that can lead to the exclusive formation of
Au
25
(SR)
18
clusters in high yield and high purity.
An interesting size-focusing process was
observed during the course of prolonged aging
of the crude product, evidenced by the spectral
evolution (Figure 3) [34]. Monodisperse
Au
25
(SR)
18
clusters (Figure 4A) were formed in
high yield (~50%, Au atom basis). The
Au
25
(SR)
18
clusters show three distinct
absorption bands at 670, 450, and 400 nm
(Figure 4B). Of note, ultrasmall nanoclusters are
barely observable under bright field TEM [35],
and sometimes electron beam bombardment
results in severe aggregation of clusters. This
size focusing growth process was found to be
quite common to various types of thiols (e.g.,
HS-C
2
H
4
Ph, HS-C
6
H
13
, HS-C
12
H
25
, HS-G
(glutathione), HS-C
10
H
22
COOH, etc) [34, 64, 67,
68].
Guided by the size-focusing method,
Qian et al. have also succeeded in preparing
atomically monodisperse Au
38
(SR)
24

nanoclusters and successfully determined the
structure by X-ray crystallography. A two-step
procedure, rather than a one-pot process as the
Au
25
(SR)
18
synthesis, was developed for


Figure 3. Evolution of the UV-vis spectra of the crude product with aging time. Adapted from ref. 66.
JOURNAL OF NANOSCIENCE LETTERS

79

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

synthesizing Au
38
(SR)
24
[54, 55]. First, a crude
mixture of glutathionate (-SG) capped Au
n
(SG)
m

nanoclusters was made. Then, the mixture was
subject to a thermal thiol etching process in a
two-phase (water/organic) system (Figure 5). A
prolonged reaction process in the presence of
excess thiol (e.g. neat dodecanethiol or diluted
phenylethylthiol) caused gold core etching
(perhaps involving air as well), and eventually
the starting polydisperse particles were
converted to monodisperse Au
38
(SR)
24
with high
purity (Figure 4C). Detailed investigation on the
conversion process clearly shows a size focusing
process, evidenced both by optical spectra and
MALDI mass spectra. The initial optical
spectrum of the starting Au
n
(SC
2
H
4
Ph)
m
shows a
decaying curve, indicating a mixture; with
reaction going on, several distinct peaks start to
emerge in the spectrum and the final product
shows a distinctive optical spectrum exhibiting
multiple bands (Figure 4D). MALDI-MS
analysis reveals that, with the increase of
reaction time, those relatively large Au
nanoclusters (n>38) seem to convert to
Au
38
(SC
2
H
4
Ph)
24
. After ~40 hrs, very clean
Au
38
(SC
2
H
4
Ph)
24
nanoclusters were obtained;
note that a fragment (9342 Da) is seen in the
MALDI spectrum due to minor fragmentation
(Figure 5, right panel), but this fragment is not
seen in ESI-MS (Figure 4C) [55].
A key condition to obtain single-size
nanoclusters is to control the initial size
distribution of the Au
n
(SG)
m
mixture prior to the
size focusing step. In the synthesis of
Au
38
(SC
2
H
4
Ph)
24
, Qian et al. realized that the
solvent plays an important role in controlling the
size range of the Au
n
(SG)
m
starting mixture.
Acetone was found to be a good solvent for high
`
Figure 4. ESI-MS (left panel) and UV-vis absorption (right panel) spectra of Au
25
(SR)
18
, Au
38
(SR)
24
and
Au
144
(SR)
60
clusters, respectively. Note: R=C
2
H
4
Ph. Adapted from refs. 60, 55 and 59.


Figure 5. Size focusing synthesis of monodisperse Au
38
(SR)
24
nanoclusters. Adapted from ref. 55.
JOURNAL OF NANOSCIENCE LETTERS

80

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

yielding synthesis of Au
38
(SR)
24
nanoclusters.
Mass spectrometry analysis shows that the
acetone mediated synthesis produces a dominant
size range from 8 to 18 kDa for the starting
mixture, size focusing of this mixture leads to
Au
38
(SC
2
H
4
Ph)
24
in ~25% yield (Au atom basis).
Note that the remaining ~75% gold is lost to
Au(I)-SR [55].
The size-focusing methodology has also
been employed to synthesize atomically
monodisperse Au
144
(SR)
60
nanoparticles. In a
two-step approach, polydisperse Au nanoclusters
capped by SC
2
H
4
Ph thiolate were first made by
NaBH
4
reduction of polymeric Au(I):SR
intermediates [59].

Again, it is important to
control the size range of the Au
n
(SC
2
H
4
Ph)
m

starting mixture. One should control the size
range to be >Au
38
, otherwise both Au
144
and
Au
38
would be resulted after size focusing and
separation of them is difficult. By adjusting the
experimental conditions, Qian et al. obtained
appropriately size-distributed Au clusters (< 2
nm) as the starting material for subsequent size-
focusing to Au
144
(SR)
60
. MALDIMS analysis
showed a dominant size range from ~24 to 36
kDa [59]. These polydisperse nanoclusters were
used as the precursor for size focusing. In the
presence of excess PhC
2
H
4
SH and at 80
o
C, the
polydisperse Au nanoclusters were gradually
converted to monodisperse Au
144
within ~24 hrs.
ESI-MS confirms that the as-prepared
nanoclusters were monodispersed (Figure 4E); a
set of peaks (z= 2+) corresponding to Cs-adducts
of [Au
144
(SC
2
H
4
Ph)
60
Cs
x
]
2+
(x=0 to 4) was
observed in ESI-MS. The UV-vis absorption
spectrum of the initial Au nanoclusters shows a
featureless curve, implying polydisperse clusters,
but after size focusing, the final
Au
144
(SC
2
H
4
Ph)
60
nanoclusters show prominent
multiple absorption bands at ~510 nm (2.44 eV)
and ~700 nm (1.78 eV), Figure 4F [59].
Taken together, the size focusing
methodology is remarkable and represents a
major step toward the rational synthesis of
atomically monodisperse Au
n
nanoclusters [66].
This method should be extendable to the
synthesis of other stable Au
n
(SR)
m
nanoclusters.
As discussed above, a general scheme of the
synthesis include two steps: i) to prepare an
appropriately size distributed Au
n
(SR)
m
mixture,
and ii) to effect size focusing, e.g. room
temperature spontaneous focusing in the
synthesis of Au
25
(SR)
18
or thermally induced
size focusing with excess thiol in the cases of
Au
38
(SR)
24
and Au
144
(SR)
60
. A key condition is
to control the size distribution of the starting
nanoclusters prior to size focusing: a too broad
size distribution would result in several robust
nanoclusters after size-focusing and isolation of
them would be very difficult. Apparently, the
fundamental basis of this size focusing
methodology lies in the stability property of
Au
n
(SR)
m
nanoclusters of certain sizes. This
leads to several important questions: what sizes
are particularly robust? And, what factors
determine their stability? These are still open
questions.

4. Optical properties of Au
n
(SR)
m

nanoclusters

To date, a number of atomically precise
nanoclusters have been made. They show many
interesting physical and chemical properties [28,
69-71]. Herein, we only illustrate the optical
absorption properties of gold nanoclusters, as
these are among the most interesting properties
of metal nanoparticles.
Unlike metallic Au nanoparticles (> 2
nm), which exhibit a collective electronic
response that gives rise to a plasmon band in
optical absorption/scattering, Au
n
(SR)
m

nanoclusters undergo discrete electronic
transitions due to excitation of a single electron
between molecular-like orbitals (e.g. HOMO and
LUMO). Using Au
25
(SR)
18
as an example,
density functional theory (DFT) calculations on
the basis of the crystal structure provide a
precise correlation of the Au
25
structure with its
optical absorption property [15, 72-74].
The electronic structure of Au
25
(SR)
18

nanoclusters clearly shows quantum size effects:
a discrete electronic structure was observed
(Figure 6A) [15], which is composed of a
significantly quantized sp band and dense d
bands. This is in striking contrast with the
continuous band structure of plasmonic Au
nanoparticles or bulk gold. The simulated optical
absorption spectrum agrees well with the
experimental one. In experiment, three
absorption bands are observed in Au
25
(SR)
18

nanoclusters (Figure 6B) [32]. The origins of
these optical absorption peaks are identified. The
lowest energy peak at 670 nm/1.85 eV (peak a in
Figure 6) corresponds to a LUMO

HOMO
transition, which is essentially an intraband
(sp

sp) transition (note that we borrowed the


term sp band from solid state theory). Because
only the three orbitals in the HOMO (triply
degenerate) have more s character than d
character, transitions arising out of the other
occupied HOMO-n orbitals tend to be interband
(sp

d) transitions, such as peak b at 450 nm/


2.63 eV, which arises from mixed intraband
(sp

sp) and interband (sp

d) transitions (i.e.
three quasi-degenerate transitions, Figure 6A).
The peak c at 400 nm/ 2.91 eV arises principally
from an interband transition (sp

d). Taken
together, the optical absorption properties of
JOURNAL OF NANOSCIENCE LETTERS

81

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

Au
25
clusters apparently root in strong quantum
confinement of valence electrons in the cluster.
Interestingly, the HOMO and LUMO orbitals are
comprised almost exclusively of atomic orbital
contributions from the 13 Au atoms in the
icosahedral core rather than the 12 exterior Au
atoms. Thus, the first 670 nm (1.85 eV) peak in
the absorption spectrum can be viewed as a
transition that is due entirely to the electronic
and geometric structure of the Au
13
core [15].
Surprisingly, the anionic [Au
25
(SR)
18
]


can be spontaneously oxidized to form charge
neutral [Au
25
(SR)
18
]
0
[33].

Zhu et al grew single
crystals of the product; the structure was found
to be almost the same as the anion but with some
small differences. The charge state shows some
noticeable effects in the optical spectra of the
clusters, albeit the overall spectral profiles are
still similar. Specifically, the 800-nm band
observed in the anion disappears in the neutral
[Au
25
(SR)
18
]
0
cluster, and the 400 nm band
less pronounced in the anion become very
prominent in the neutral cluster [33]. These fine
features can serve as spectroscopic fingerprints
for convenient identification of the cluster
charge state in practical work. Interestingly, the
charge-neutral cluster was found to be
paramagnetic, while the anion is diamagnetic
[75].
The Au
25
cluster serves as a good
example to illustrate the quantum size effect on
the optical properties of gold nanoclusters.
Unlike Au nanocrystals in which surface
plamson (collective excitation of conduction
electrons) dominates the optical properties, one-
electron transitions are manifested in quantum
sized Au
n
(SR)
m
nanoclusters. An interesting
question is how many electrons are needed in
order to support surface plasmon excitation.
And, at what size does the molecule-like state
transit to the metallic state? These fundamental
questions remain to be addressed in future work.
Overall, the structure and
physicochemical properties of ultrasmall
Au
n
(SR)
m
nanoclusters are very sensitive to the
number of atoms in the particle; for example,
Qian et al observed drastic differences in the
optical absorption spectra of Au
38
(SR)
24
and
Au
40
(SR)
24
[56]. Thus, precise control over the
number of atoms in the cluster is very critical. In
future work, it is essential to prepare a series of
size discrete, atomically precise Au
n
(SR)
m

clusters in order to reveal the evolution of the
optical properties of Au
n
(SR)
m
clusters with size.

5. Applications of Au
n
(SR)
m
nanoclusters

Much work remains to be carried out to
explore the practical utility of nanocluster
materials. Herein, we illustrate the catalytic
opportunities of Au
n
(SR)
m
nanoclusters. Such
well defined nanoclusters (with ligands on or


Figure 6. (A) Kohn-Sham orbital level diagram for Au
25
(SR)
18
. (B) Peak assignment of
the absorption spectrum of Au
25
(SR)
18
clusters. Adapted from ref. 15.
JOURNAL OF NANOSCIENCE LETTERS

82

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

removed) will provide important insight into the
size dependent catalytic activity and selectivity
and the structure-property correlation. The
Au
n
(SR)
m
nanoclusters are indeed found to be
highly active in catalyzing some chemical
processes, such as selective oxidation of alcohol
to aldehyde [76-78], selective oxidation of
alkene to epoxide or aldehyde [79-82],
chemoselective hydrogenation of o, |-
unsaturated ketones to unsaturated alcohols [83,
84].

5.1. Selective oxidation using Au
n
L
m

nanocluster catalysts
Tsukuda and coworker found PVP-
protected Au
n
nanoclusters smaller than 1.5 nm
showed higher activity for aerobic oxidation of
alcohol than those of larger size or stabilized by
poly(allylamine) (PAA) [77]. Extensive
characterization revealed that the catalytically
active Au nanoclusters are indeed negatively
charged by electron donation from PVP ligands.
The catalytic activity was found to enhance with
increasing electron density on the Au core. They
proposed that electron transfer from the anionic
Au
n
core to the LUMO (*) of O
2
generates
superoxo- or peroxo-like species, which in turn
oxidize alcohol to aldehyde (or ketone), Scheme
2 [77]. Our recent work has demonstrated that
thiolate-protected Au
n
(SR)
m
nanoclusters are
indeed excellent catalysts for the selective
oxidation reaction of styrene [81]. The reaction
was carried out in solution phase by adding free
(unsupported) Au
25
(SC
2
H
4
Ph)
18
catalysts to a
toluene solution of styrene, followed by O
2

introduction. At 80
o
C (oil bath) and 1 atm
pressure for ~18 hrs, conversion of styrene up to
~36% was observed, and the reaction product
consists of benzadehyde (~65% selectivity),
styrene epoxide (~30%), and minor
acetophenone (<5%) [81]. Using organic
peroxide (e.g. TBHP) as oxidant, 100%
conversion of styrene and high selectivity for
benzaldehyde or epoxide can be obtained [79,
82].
Using Au
25
, Au
38
and Au
144
nanocluster
catalysts, some interesting size dependent results
were observed in the selective oxidation of
styrene. Small size nanoclusters (e.g. Au
25
) were
found to be critical in activating O
2
, while TBHP
activation does not show a strong dependence on
Au
n
size [82]. Liu et al. investigated Au
n
(n = 10,
18, 25, 39) supported on hydroxyapatite (HAP)
for aerobic oxidation of cyclohexane. These
Au
n
/HAP catalysts were found to efficiently
oxidize cyclohexane to cyclohexanol and
cyclohexanone. A monotonic increase in the
turnover frequency was found with n up to 39,
then followed by an activity decrease with a
further increase in n up to 85 [85].

5.2. Selective hydrogenation of , -
unsaturated ketones and aldehydes
Zhu et al. have demonstrated that the


Scheme 2. PVP capped Au nanoclusters as catalysts for selective oxidation of alcohols. Adapted from ref. 77.

Scheme 3. Chemoselective hydrogenation of o, |-unsaturated ketone to unsaturated
alcohol catalyzed by Au
25
(SR)
18
nanoclusters. Adapted from ref. 83.
JOURNAL OF NANOSCIENCE LETTERS

83

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

Au
25
(SR)
18
nanoclusters are also excellent
catalysts for selective hydrogenation of o, |-
unsaturated ketones to unsaturated alcohols at
room temperature [83]. A complete selectivity
(100%) to o, |-unsaturated alcohols was
obtained in such reactions. A mechanism has
also been proposed (Scheme 3) [83]. For future
work, further investigation on the
stereoselectivity and enantioselectivity of
nanocluster catalysts would be interesting. The
two chiral systemsAu
25
with chiral thiolate
and intrinsically chiral Au
38
with achiral or chiral
ligands serve as good models. The doped
nanoclusters, such as Pd
1
Au
24
(SR)
18
[86-88], are
also interesting for selective hydrogenation
reactions.
The results discussed above clearly
demonstrate the great promise of nanocluster
catalysts for catalysis as well as their unique
potential in revealing some fundamental aspects
of nanogold catalysis. In future work, with a
more complete size series that spans the
quantum confinement to the plasmonic regimes,
one will be able to elucidate some interesting
questions: 1) Will the size dependent catalytic
activity/selectivity show a linear or irregular
relationship? 2) Will the presence of thiolate
ligands inhibit the catalytic activity? 3) How
does the plasmonic excitation affect the catalytic
properties? The available crystal structures of
Au
n
(SR)
m
nanoclusters will permit identification
of catalytic active sites and reaction pathways. In
addition, these well defined catalysts provide a
unique opportunity to investigate the intriguing
interactions between nanoclusters and supports.
Charge transfer may occur between transition
metal oxide and supported nanoclusters; for
instance, in the literature Landman and Heiz et
al discussed charge transfer from oxide defect
sites to Au
n
and its major effect on CO oxidation
[89]. By correlating the catalytic performance
with structure and electronic properties
Au
n
(SR)
m
nanoclusters, one will be able to map
out how these effects influence catalytic
performance (e.g. activity and selectivity).

6. Outlook

The atomically monodisperse
nanoclusters discussed above represent a very
important intermediate size regime between
localized atomic states and delocalized band
structure. A lot of transitions in terms of material
properties occur in the size range of
nanoclusters, i.e. from a dozen to a few hundred
atoms (equivalent diameter from subnanometer
to ~2 nm).
With respect to the synthesis, the size-
focusing methodology has opened up the
possibility of preparing robust nanoclusters with
excellent control over size. We anticipate that
this methodology will be quite useful in
fulfilling the synthesis of a series of size
discrete, atomically precise nanoclusters ranging
from the quantum confinement regime to the
plasmonic regime. In future work, extension to
the synthesis of well defined silver nanoclusters
[90-95], as well as other metals [96, 97], is also
highly desirable.
For optical properties, metal
nanoparticles (> 2 nm in diameter) are
dominated by the surface plasmon resonance,
and its presence is taken to be indicative that a
particle is metallic since only free electrons can
be collectively excited and create a surface
plasmon. For Au
n
(SR)
m
nanoclusters, discrete
energy states and accordingly one-electron
excitations appear. The general trend of the
HOMO-LUMO gap evolution is that the
absorption band will red-shift and its intensity
will decrease with increasing cluster size, and
eventually the plasmon excitation mode will
emerge. However, it is not clear whether the
transition is smooth over a size range or abrupt
at a critical size. In addition, it is of particular
interest to find out whether the structural
evolution to face-centered cubic (fcc) occurs
concurrently with the optical property transition.
One of the major applications of metal
nanoclusters is in catalysis. Compared to fcc-
crystalline Au nanoparticles (i.e. metallic),
Au
n
L
m
nanoclusters not only open up the
possibility of tuning a catalytic process by
changing cluster size, but also the possibility of
catalyzing chemical reactions at low
temperatures (e.g. <100
o
C), which is a quite
attractive feature of nanocluster catalysts. The
size series of Au
n
(SR)
m
spanning the non-
plasmonic and plasmonic size regimes will
provides a new model system for investigating
the quantum confinement and plasmonic effects
on the catalytic properties of gold nanoparticles.
Well defined nanocluster catalysts will lead to
fundamental understanding of the origin of
nanogold catalysis, especially how quantum
confinement and surface plasmons influence the
catalytic properties. By correlating with the
crystal structures of Au
n
(SR)
m
nanoclusters, the
catalytic active sites and reaction pathways may
be identified ultimately, which will provide
important feedback to the design of highly active
and selective nanocatalysts for certain reaction
processes.
Overall, a central task of nanocluster
research is to study the evolution from the
molecular-like cluster state to the metallic state.
Future studies will significantly expand
understanding of the evolution of the material
properties from the atomic state to nanoclusters
to fcc-structured nanocrystals. The evolution
JOURNAL OF NANOSCIENCE LETTERS

84

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

behavior across length scales will lead to
fundamental insights into structure-property
correlation and new principles for design of
novel functional metal nanomaterials for
catalysis, energy conversion, sensing, and many
others.

References

1. C. Burda, X. Chen, R. Narayanan, M. A. El-
Sayed, Chem. Rev. 105 (2005) 1025.
2. X. G. Peng, M. C. Schlamp, A. V.
Kadavanich, A. P. Alivisatos, J. Am. Chem.
Soc. 119 (1997) 7019.
3. U. Kreibig, M. Vollmer, Optical Properties
of Metal Clusters, Springer-Verlag, New
York (1995).
4. R. Jin, Nanoscale 2 (2010) 343.
5. T. G. Schaaff, M. N. Shafigullin, J. T.
Khoury, I. Vezmar, R. L. Whetten, W. G.
Cullen, P. N. First, C. Gutierrez-Wing, J.
Ascensio, M. J. Jose-Yacaman, J. Phys.
Chem. B 101 (1997) 7885.
6. Y. Negishi, K. Nobusada, T. Tsukuda, J.
Am. Chem. Soc. 127 (2005) 5261.
7. C. J. Murphy, A. M. Gole, S. E. Hunyadi, C.
J. Orendorff, Inorg. Chem. 45 (2006) 7544.
8. R. Jin, Y. Cao, C. A. Mirkin, K. L. Kelly, G.
C. Schatz, J. G. Zheng, Science 294 (2001)
1901.
9. R. Jin, Y. C. Cao, E. Hao, G. S. Metraux, G.
C. Schatz, C. A. Mirkin, Nature 425 (2003)
487.
10. R. L. Whetten, M. N. Shafigullin, J. T.
Khoury, T. G. Schaaff, I. Vezmar, M. M.
Alvarez, A. Wilkinson, Acc. Chem. Res. 32
(1999) 397.
11. O. Varnavski, G. Ramakrishna, J. Kim, D.
Lee, T. Goodson, J. Am. Chem. Soc. 132
(2010) 16.
12. T. G. Schaaff, G. Knight, M. N. Shafigullin,
R. F. Borkman, R. L. Whetten, J. Phys.
Chem. B 102 (1998) 10643.
13. P. D. Jadzinsky, G. Calero, C. J. Ackerson,
D. A. Bushnell, R. D. Kornberg, Science
318 (2007) 430.
14. M. W. Heaven, A. Dass, P. S. White, K. M.
Holt, R. W. Murray, J. Am. Chem. Soc. 130
(2008) 3754.
15. M. Zhu, C. M. Aikens, F. J. Hollander, G.
C. Schatz, R. Jin, J. Am. Chem. Soc. 130
(2008) 5883.
16. H. Qian, W. T. Eckenhoff, Y. Zhu, T.
Pintauer, R. Jin, J. Am. Chem. Soc. 132
(2010) 8280.
17. M. Walter, J. Akola, O. Lopez-Acevedo, P.
D. Jadzinsky, G. Calero, C. J. Ackerson, R.
L. Whetten, H. Gronbeck, H. Hakkinen,
Proc. Natl. Acad. Sci. U.S.A. 105 (2008)
9157.
18. C. Gautier, T. Burgi, ChemPhysChem 10
(2009) 483.
19. D. Jiang, M. L. Tiago, W. Luo, S. Dai, J.
Am. Chem. Soc. 130 (2008) 2777.
20. D.-e. Jiang, M. Walter, S. Dai, Chem. Eur.
J. 16 (2010) 4999.
21. M. Brust, M. Walker, D. Bethell, D. J.
Schiffrin, R. Whyman, J. Chem. Soc.,
Chem. Commun. (1994) 801.
22. C. A. Mirkin, R. L. Letsinger, R. C. Mucic,
J. J. Storhoff, Nature 382 (1996) 607.
23. R. L. Whetten, J. T. Khoury, M. M. Alvarez,
S. Murthy, I. Vezmar, Z. L. Wang, P. W.
Stephens, C. L. Cleveland, W. D. Luedtke,
U. Landman, Adv. Mater. 8 (1996) 428.
24. M. M. Alvarez, J. T. Khoury, T. G. Schaaff,
M. Shafigullin, I. Vezmar, R. L. Whetten,
Chem. Phys. Lett. 266 (1997) 91.
25. M. J. Hostetler, J. E. Wingate, C.-J. Zhong,
J. E. Harris, R. W. Vachet, M. R. Clark, J.
D. Londono, S. J. Green, J. J. Stokes, G. D.
Wignall, G. L. Glish, M. D. Porter, N. D.
Evans, R. W. Murray, Langmuir 14 (1998)
17.
26. M. J. Hostetler, S. J. Green, J. J. Stokes, R.
W. Murray, J. Am. Chem. Soc. 118 (1996)
4212.
27. M. M. Alvarez, J. T. Khoury, T. G. Schaaff,
M. N. Shafigullin, I. Vezmar, R. L.
Whetten, J. Phys. Chem. B 101 (1997)
3706.
28. R. L. Donkers, D. Lee, R. W. Murray,
Langmuir 20 (2004) 1945.
29. R. L. Donkers, D. Lee, R. W. Murray,
Langmuir 24 (2008) 5976.
30. J. B. Tracy, G. Kalyuzhny, M. C. Crowe, R.
Balasubramanian, J.-P. Choi, R. W. Murray,
J. Am. Chem. Soc. 129 (2007) 6706.
31. J. B. Tracy, M. C. Crowe, J. F. Parker, O.
Hampe, C. A. Fields-Zinna, A. Dass, R. W.
Murray, J. Am. Chem. Soc. 129 (2007)
16209.
32. M. Zhu, E. Lanni, N. Garg, M. E. Bier, R.
Jin, J. Am. Chem. Soc. 130 (2008) 1138.
33. M. Zhu, W. T. Eckenhoff, T. Pintauer, R.
Jin, J. Phys. Chem. C 112 (2008) 14221.
34. Z. Wu, J. Suhan, R. Jin, J. Mater. Chem. 19
(2009) 622.
35. Z. Wu, C. Gayathri, R. R. Gil, R. Jin, J. Am.
Chem. Soc. 131 (2009) 6535.
36. Z. Wu, R. Jin, ACS Nano 3 (2009) 2036.
37. M. Zhu, H. Qian, R. Jin, J. Am. Chem. Soc.
131 (2009) 7220.
38. M. Zhu, H. Qian, R. Jin, J. Phys. Chem.
Lett. 1 (2010) 1003.
39. T. G. Schaaff, R. L. Whetten, J. Phys.
Chem. B 103 (1999) 9394.
40. G. H. Woehrle, M. G. Warner, J. E.
Hutchison, J. Phys. Chem. B 106 (2002)
9979.
41. G. H. Woehrle, J. E. Hutchison, Inorg.
Chem. 44 (2005) 6149.
42. G. H. Woehrle, L. O. Brown, J. E.
Hutchison, J. Am. Chem. Soc. 127 (2005)
2172.
JOURNAL OF NANOSCIENCE LETTERS

85

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

43. R. Balasubramanian, R. Guo, A. J. Mills, R.
W. Murray, J. Am. Chem. Soc. 127 (2005)
8126.
44. Y. Shichibu, Y. Negishi, T. Tsukuda, T.
Teranishi, J. Am. Chem. Soc. 127 (2005)
13464.
45. Y. Shichibu, Y. Negishi, T. Watanabe, N. K.
Chaki, H. Kawaguchi, T. Tsukuda, J. Phys.
Chem. C 111 (2007) 7845.
46. R. C. Price, R. L. Whetten, J. Am. Chem.
Soc. 127 (2005) 13750.
47. H. Qian, M. Zhu, E. Lanni, Y. Zhu, M. E.
Bier, R. Jin, J. Phys. Chem. C 113 (2009)
17599.
48. H. Tsunoyama, Y. Negishi, T. Tsukuda, J.
Am. Chem. Soc. 128 (2006) 6036.
49. H. Tsunoyama, P. Nickut, Y. Negishi, K.
Al-Shamery, Y. Matsumoto, T. Tsukuda, J.
Phys. Chem. C 111 (2007) 4153.
50. H. Tsunoyama, T. Tsukuda, J. Am. Chem.
Soc. 131 (2009) 7086.
51. N. K. Chaki, Y. Negishi, H. Tsunoyama, Y.
Shichibu, T. Tsukuda, J. Am. Chem. Soc.
130 (2008) 8608.
52. O. Toikkanen, V. Ruiz, G. Ronholm, N.
Kalkkinen, P. Liljeroth, B. M. Quinn, J. Am.
Chem. Soc. 130 (2008) 11049.
53. O. Toikkanen, S. Carlsson, A. Dass, G.
Ronnholm, N. Kalkkinen, B. M. Quinn, J.
Phys. Chem. Lett. (2009) 32.
54. H. Qian, M. Zhu, U. N. Andersen, R. Jin, J.
Phys. Chem. A 113 (2009) 4281.
55. H. Qian, Y. Zhu, R. Jin, ACS Nano 3 (2009)
3795.
56. H. Qian, Y. Zhu, R. Jin, J. Am. Chem. Soc.
132 (2010) 4583.
57. T. G. Schaaff, M. N. Shafigullin, J. T.
Khoury, I. Vezmar, R. L. Whetten, J. Phys.
Chem. B 105 (2001) 8785.
58. J. F. Hicks, D. T. Miles, R. W. Murray, J.
Am. Chem. Soc. 124 (2002) 13322.
59. H. Qian, R. Jin, Nano Lett. 9 (2009) 4083.
60. M. A. MacDonald, P. Zhang, H. Qian, R.
Jin, J. Phys. Chem. Lett. 1 (2010) 1821.
61. C. A. Fields-Zinna, R. Sardar, C. A.
Beasley, R. W. Murray, J. Am. Chem. Soc.
131 (2009) 16266.
62. A. Dass, A. Stevenson, G. R. Dubay, J. B.
Tracy, R. W. Murray, J. Am. Chem. Soc.
130 (2008) 5940.
63. A. Dass, K. Holt, J. F. Parker, S. W.
Feldberg, R. W. Murray, J. Phys. Chem. C
112 (2008) 20276.
64. A. C. Dharmaratne, T. Krick, A. Dass, J.
Am. Chem. Soc. 131 (2009) 13604.
65. A. Dass, J. Am. Chem. Soc. 131 (2009)
11666.
66. R. Jin, H. Qian, Z. Wu, Y. Zhu, M. Zhu, A.
Mohanty, N. Garg, J. Phys. Chem. Lett. 1
(2010) 2903.
67. J. F. Parker, J. E. F. Weaver, F. McCallum,
C. A. Fields-Zinna, R. W. Murray,
Langmuir 26 (2010) 13650.
68. M. A. H. Muhammed, A. K. Shaw, S. K.
Pal, T. Pradeep, J. Phys. Chem. C 112
(2008) 14324.
69. D. Garcia-Raya, R. Madueno, M. Blazquez,
T. Pineda, J. Phys. Chem. C 113 (2009)
8756.
70. R. Hamouda, B. Bellina, F. Bertorelle, I.
Compagnon, R. Antoine, M. Broyer, D.
Rayane, P. Dugourd, J. Phys. Chem. Lett. 1
(2010) 3189.
71. N. Sakai, T. Tatsuma, Adv. Mater. 22
(2010) 3185.
72. J. Akola, M. Walter, R. L. Whetten, H.
Hakkinen, H. Gronbeck, J. Am. Chem. Soc.
130 (2008) 3756.
73. C. M. Aikens, J. Phys. Chem. Lett. 1 (2010)
2594.
74. C. M. Aikens, J. Phys. Chem. Lett. 2 (2011)
99.
75. M. Zhu, C. M. Aikens, M. P. Hendrich, R.
Gupta, H. Qian, G. C. Schatz, R. Jin, J. Am.
Chem. Soc. 131 (2009) 2490.
76. H. Tsunoyama, H. Sakurai, Y. Negishi, T.
Tsukuda, J. Am. Chem. Soc. 127 (2005)
9374.
77. H. Tsunoyama, N. Ichikuni, H. Sakurai, T.
Tsukuda, J. Am. Chem. Soc. 131 (2009)
7086.
78. H. Tsunoyama, H. Sakurai, T. Tsukuda,
Chem. Phys. Lett. 429 (2006) 528.
79. Y. Liu, H. Tsunoyama, T. Akita, T.
Tsukuda, Chem. Commun. 46 (2010) 550.
80. Y. Liu, H. Tsunoyama, T. Akita, S. Xie, T.
Tsukuda, ACS Catalysis 1 (2011) 2.
81. Y. Zhu, H. Qian, M. Zhu, R. Jin, Adv.
Mater. 22 (2010) 1915.
82. Y. Zhu, H. Qian, R. Jin, Chem. Eur. J. 16
(2010) 11455.
83. Y. Zhu, H. Qian, B. A. Drake, R. Jin,
Angew. Chem. Int. Ed. 49 (2010) 1295.
84. Y. Zhu, Z. Wu, C. Gayathri, H. Qian, R. R.
Gil, R. Jin, J. Catal. 271 (2010) 155.
85. Y. Liu, H. Tsunoyama, T. Akita, S. Xie, T.
Tsukuda, ACS Catalysis (2011).
86. C. A. Fields-Zinna, M. C. Crowe, A. Dass,
J. E. F. Weaver, R. W. Murray, Langmuir
25 (2009) 7704.
87. Y. Negishi, W. Kurashige, Y. Niihori, T.
Iwasa, K. Nobusada, Phys. Chem. Chem.
Phys. 12 (2010) 6219.
88. H. Qian, E. Barry, Y. Zhu, R. Jin, Acta
Phys. Chim. Sin. 27 (2011) 513.
89. U. Landman, B. Yoon, C. Zhang, U. Heiz,
M. Arenz, Top. Catal. 44 (2007) 145.
90. Z. Wu, E. Lanni, W. Chen, M. E. Bier, D.
Ly, R. Jin, J. Am. Chem. Soc. 131 (2009)
16672.
91. S. Kumar, M. D. Bolan, T. P. Bigioni, J.
Am. Chem. Soc. 132 (2010) 13141.
92. N. Cathcart, P. Mistry, C. Makra, B.
Pietrobon, N. Coombs, M. Jelokhani-
Niaraki, V. Kitaev, Langmuir 25 (2009)
5840.
JOURNAL OF NANOSCIENCE LETTERS

86

Journal of Nanoscience Letters | Volume 1 | Issue 2 | 2011
www.simplex-academic-publishers.com
2011 Simplex Academic Publishers. All rights reserved.

93. N. Cathcart, V. Kitaev, J. Phys. Chem. C
114 (2010) 16010.
94. O. M. Bakr, V. Amendola, C. M. Aikens,
W. Wenseleers, R. Li, L. D. Negro, G. C.
Schatz, F. Stellacci, Angew. Chem. Int. Ed.
48 (2009) 5921.
95. T. U. B. Rao, B. Nataraju, T. Pradeep, J.
Am. Chem. Soc. 132 (2010) 16304.
96. C. Femoni, M. C. Iapalucci, G. Longoni, S.
Zacchini, S. Zarra, J. Am. Chem. Soc. 133
(2011) 2406.
97. E. G. Mednikov, L. F. Dahl, J. Am. Chem.
Soc. 130 (2008) 14813.

You might also like