You are on page 1of 17

G Model

NANTOD-639; No. of Pages 17 ARTICLE IN PRESS


Nano Today xxx (2018) xxx–xxx

Contents lists available at ScienceDirect

Nano Today
journal homepage: www.elsevier.com/locate/nanotoday

Review

Heterogeneous catalysis by gold and gold-based bimetal nanoclusters


Jianbo Zhao a,b,∗ , Rongchao Jin b,∗
a
Henan Provincial Key Laboratory of Surface and Interface Science, Zhengzhou University of Light Industry, No.136 of Science Road, Zhengzhou 450001,
China
b
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213, United States

a r t i c l e i n f o a b s t r a c t

Article history: Gold and gold-based bimetallic catalysts have shown great potential for many important chemical trans-
Received 8 September 2017 formation reactions owing to their good activity and high selectivity under relatively mild conditions.
Received in revised form 8 November 2017 However, elucidating their structure-property relationships remains a major challenge because of size
Accepted 23 December 2017
distributions and ill-defined structures (i.e. structural heterogeneity) of conventional nanoparticle cata-
Available online xxx
lysts. Recently, controlled synthesis of homogold and gold-based bimetal nanoclusters with precise atom
numbers have been accomplished, and more importantly their atomic structures have been elucidated,
Keywords:
which provide new model systems for understanding the catalytic behavior at the atomic level. Due
Catalysis
Homogold nanoclusters
to their ultra-small size, gold and bimetal nanoclusters exhibit unique properties, which contribute to
Gold-based bimetal nanoclusters novel catalysts or precatalysts for many chemical reactions. Herein, we review the recent progress in
Ligand the catalytic research of atomically precise gold and gold-based bimetal nanoclusters. Specifically, this
Structure-property relationship review covers the application of nanoclusters with or without ligands as catalysts for reactions such as
selective oxidation, selective hydrogenation, C–C coupling and photocatalysis. Finally, we provide some
future perspectives about metal nanocluster catalysis.
© 2017 Elsevier Ltd. All rights reserved.

Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Synthetic breakthroughs of gold and bimetallic nanoclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Homogold nanoclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Size-focusing methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
LEIST methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Gold-based bimetal nanoclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Catalytic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Homogold nanoclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Hydrogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Carbon–Carbon coupling reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Photocatalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Gold-based bimetal nanoclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Selective hydrogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Carbon–Carbon coupling reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Future perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00

∗ Corresponding authors at: Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213, United States.
E-mail addresses: zhaojianbo@zzuli.edu.cn (J. Zhao), rongchao@andrew.cmu.edu (R. Jin).

https://doi.org/10.1016/j.nantod.2017.12.009
1748-0132/© 2017 Elsevier Ltd. All rights reserved.

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
2 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Introduction

Heterogeneous catalysis occupies a pivotal position in the man-


ufacture of essential products in the fields of chemical industry,
agriculture, pharmaceuticals and numerous materials [1]. One of
the core missions of heterogeneous catalysis is to design and
develop high-performance catalysts. As an important type of het-
erogeneous catalysts, metal catalysts have been extensively applied
in many important reactions such as partial oxidation and selec-
tive hydrogenation [2,3]. Particularly, with the rapid development
of nanoscience and nanotechnology, both academia and industry
communities have shown increasing interest in controlling metal
nanoparticles for enhancing the catalytic activity, tuning the selec-
tivity and even obtaining unique chemo- and stereo-selectivity for
important catalytic processes [4–7].
Gold was once considered to be catalytically inactive for a
long time. However, since the 1980s there has been increasing
interest in gold catalysis [8,9]. It was the first time for gold to
become a high-performance catalyst, in sharp contrast to the ear-
lier reports about no or poor activity of gold. Since then, much
work has further explored the highly dispersed Au catalysts for
a range of chemical transformations, such as selective oxidations,
selective hydrogenations, C–C coupling reactions, and the addi- Scheme 1. Schematic Diagram of the Size-Focusing Methodology. Reproduced with
tion of multiple C C bonds [10–14]. The Au nanocatalysts are permission from [18]. Copyright 2016 American Chemical Society.

usually obtained by conventional preparation methods, including


the deposition-precipitation, co-precipitation and sol precipitation
[21]. Subsequently, thiol was employed in the preparation and
methods [8,10]. Such nanocatalysts are apparently more or less
functionalization of gold nanoparticles [22,23]. However, it was not
polydispersed in size and have no uniform composition in the case
until 2005 that ultra-small gold clusters (∼1 nm) of a series of sizes
of bimetallic catalysts, not to mention the lack of well-defined
protected by glutathione were for the first time isolated from a
structures, which result in major difficulties in identifying the cat-
mixture by the Tsukuda group [24]. The gel electrophoresis method,
alytic sites and determining the structure-property relationships of
however, did not offer large quantities of nanoclusters and the yield
Au catalysts. Thus, designing novel Au nanocatalysts with precise
of each size was low, thus bulk solution synthesis with size con-
particle size and well-defined structure is highly desirable in order
trol remained to be developed. The Jin group made major progress
to solve the above issues.
in 2007 [25] and established two systematic methodologies in the
Controlled synthesis of nanoparticles with precise numbers of
following years, including the size-focusing methodology [26] and
atoms has long been a major dream of nanochemists. In recent
the ligand exchange-induced size/structure transformation (LEIST)
years major breakthroughs have been made in preparing atomically
methodology [27]. The rational synthesis of atomically precise Au
precise nanoparticles with diameters from subnanometer to a few
nanoclusters with molecular purity and high yield can be accom-
nanometers, which are often called nanoclusters [15–19]. Among
plished by these methodologies, which provides opportunities for
them, controllable synthesis of atomically precise Au nanoclusters
investigating the structure, properties and applications of Au nan-
was first attained [15–18]. These Au nanoclusters with excellent
oclusters.
size control provide new opportunities for investigating the precise
size-dependent catalytic properties and also obtaining structural
insights into catalytic mechanisms. Besides, due to the ultra-small Size-focusing methodology
size (less than 2 nm), these Au nanoclusters exhibit unique and An ideal synthesis is to achieve one-sized Au nanoclusters
even unexpected properties that are not observed in bulk gold and in high yield. In 2007, Zhu et al. achieved for the first time
conventional gold nanoparticles. Finally, the precise atomic struc- Au25 (SCH2 CH2 Ph)18 nanoclusters in molecular purity and high
tures of more nanoclusters of gold and bimetals have also been yield through kinetic control [25]. Particularly, it was discovered
determined by X-ray crystallography [17,18,20], which allows for that the kinetic formation for the Au(I)-SR intermediate was crit-
precise correlation of the structure with the catalytic property of ical for the high yield of Au25 (SCH2 CH2 Ph)18 . Subsequently, the
nanoclusters. Herein, we review the recent progress in the cataly- size-focusing method was extended to successful syntheses of
sis by homogold and gold-based bimetal nanoclusters in the past Au38 (SR)24 [28,29], Au64 (SR)32 [30], Au99 (SR)42 [31], Au144 (SR)60
five years. We first introduce the synthetic strategies briefly, and [32], Au246 (SR)80 [17] and Au333 (SR)79 [33]. It has since become
then focus on the catalytic properties of homogold and gold-based a universal method for rational synthesis of atomically precise
bimetal nanoclusters reported in the last few years. Finally, we pro- Aun (SR)m nanoclusters in molecular purity, which are also effective
vide some perspectives of metal nanoclusters for future catalysis for preparing both the organic-soluble nanoclusters and aque-
research. ous nanoclusters [34,35]. In general, this methodology involves
two steps: the kinetic control and the thermodynamic selection
(Scheme 1) [18]. The key point lies in the effective control of the
Synthetic breakthroughs of gold and bimetallic size distribution of the initial Aux (SR)y product prior to the size-
nanoclusters focusing step. A too-broad size range would lead to more than one
stable size in the final product after the size-focusing and their iso-
Homogold nanoclusters lation would be difficult. Controlling the Aux (SR)y mixture largely
depends on the kinetic control. Some important factors are such as
Gold-thiol chemistry was traced back to the 1980s due to the the temperature, solvent and reactant ratio. For more details of the
research on thiol self-assembled monolayers on bulk gold surfaces methodology, the reader is referred to ref [26].

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 3

Scheme 2. Reaction pathway for conversion of Au38 (PET)24 to Au36 (TBBT)24 . Stage I, ligand exchange; II, structure distortion; III, disproportionation; IV, size focusing.
Reproduced with permission from [37]. Copyright 2013 American Chemical Society.

Fig. 1. X-ray structure of Au25 (SCH2 CH2 Ph)18 (magenta, gold atoms of the core; cyan, gold atoms of the shell (or surface)). (A) ball stick view; (B) space-filling view. Adapted
with permission from ref [41]. Copyright 2008 American Chemical Society.

LEIST methodology discussed herein [41]. The Au25 (SR)18 (where R = CH2 CH2 Ph) con-
This methodology, which is based on the size-focusing one, tained a 13-atom Au13 icosahedral core (Fig. 1A, magenta portion)
employed the Aun (SR)m nanoclusters as the starting material for and a Au12 (SR)18 shell (Fig. 1A, cyan and yellow) [41,42]. Due to
reaction with sufficiently different ligands than the surface lig- the unique structure, the shell Au atoms and even some of the
ands on Aun (SR)m to create new sizes of Au nanoclusters [27]. icosahedral core Au atoms are accessible to reactant molecules,
A number of new Au nanoclusters have been prepared by this thereby providing potential catalytic sites (Fig. 1B). The electronic
methodology, which are otherwise difficult to synthesize through structure of Au25 (SR)18 showed discrete energy levels caused by
the size-focusing methodology. Taking Au36 (TBBT)24 (where TBBT the quantum-size effect [41]. The molecular orbitals can be divided
represents 4-tert-butylbenzenethiolate) as an example (Scheme 2) into the core orbitals (primarily contributed by Au13 ) and the shell
[36,37]. Starting with molecularly pure Au38 (SCH2 CH2 Ph)24 , ligand orbitals s (primarily contributed by Au12 (SR)18 ), thus exhibiting the
exchange reaction with a large excess of TBBT under thermal condi- geometric and electronic core-shell structures [41,43]. It was also
tions (∼80 ◦ C) resulted in the formation of Au36 (TBBT)24 in greater proven that the ligand type has no effect on the geometric struc-
than 90% yield and molecular purity. Two key points affecting ture [44], evidenced by the same optical absorption spectrum of
the LEIST process were observed. First, the high ratio of incom- Au25 (SR)18 regardless the −R groups of the thiolates [45,46]. The
ing TBBT to the original PET (where PET represents SCH2 CH2 Ph) crystal structures of other Au nanoclusters have been reviewed
on the nanoclusters is required. Second, the thermal process is elsewhere [18].
required to overcome the energy barrier between different stable
sizes (i.e. Au38 vs Au36 ). On the basis of the method, the conversions Gold-based bimetal nanoclusters
of Au25 (SR)18 to Au28 (SR )20 [38], Au25 (SR)18 to Au20 (SR )16 [39],
Au38 (SR)24 to Au36 (SR )24 [36,37], and Au144 (SR)60 to Au133 (SR )52 A common synthetic strategy for gold-based bimetal nan-
[40] were also accomplished. oclusters is similar to that of gold nanoclusters except the
Further, the aforementioned methodologies have led to suc- use of combined gold and heterometal salts as the precursor
cessful crystallization and structure determination of many Au [18,20,47–53]. Generally, two kinds of heteroatom substitution
nanoclusters [18]. Due to the extensive application of Au25 (SR)18 behavior have been found: (1) a specific number of foreign atoms
in many fields, its structure and electronic properties are briefly (e.g. Pd or Pt) can be incorporated into gold nanocluster irrespective

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
4 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Fig. 2. X-ray crystal structure of the Au25-x Agx (SR)18 . The X-ray crystallography averaged composition is Au5.6 Ag19.4 (SC6 H11 )18 . (a) The 13-atom icosahedral alloy core. (b)
Six [–SR–Au(Ag)–SR–Ag–SR–] dimeric staples with different occupancy of Au/Ag atoms. (c) Total structure of Au25-x Agx (SR)18 . Magenta: Au, gray: Ag/Au, yellow: S, grey: C.
Reproduced with permission from ref [54]. Copyright 2016 Royal Society of Chemistry.

Fig. 3. (A) Reaction temperature dependence of CO conversion over Au25 (SR)18 /MOx catalysts. Pretreatment condition: N2 at room temperature (rt) for 0.5 h. Reaction
conditions: GHSV = 7500 mL g−1 h−1 , catalyst = 0.1 g. (B) Reaction temperature dependence of CO conversion over Au25 (SR)18 /CeO2 catalyst after different pretreatments.
GHSV = 7500 mL g−1 h−1 . Adapted with permission from ref [65]. Copyright 2012 American Chemical Society.

of the salt precursor ratio; (2) a range of heteroatom substitution is the reaction that the superior activity of gold was discovered for
(e.g. Ag or Cu) depending on the initial precursor ratio but the total the first time [8], so it has been employed extensively as a proto-
number of metal atoms in the parent cluster is preserved [20]. For typical reaction for understanding gold catalysis [60]. Besides, CO
Pd or Pt doping, experimental and theoretical works show that oxidation can also readily reveal the ligand effect on the property of
Pd or Pt is located in the center of Au25 (SR)18 clusters [20]. For Au nanoclusters. Therefore we first focus on CO oxidation catalyzed
Ag or Cu substitution, here we illustrate the case of heteroatom by gold nanoclusters.
substitution in Au25 (SR)18 since such bimetallic clusters have been It is worth noting some results of bare gold cluster catalysts.
investigated for catalysis recently. It is noted that Ag or Cu substi- Valden et al. [61] found that layered Au nanoclusters with thickness
tution in Au25 (SR)18 clusters with a specific number has not been smaller than 3 atoms were very active for low temperature CO oxi-
obtained yet [20]. The similarities between Ag and Au readily lead dation. Herzing et al. [62] investigated several iron oxide-supported
to alloy clusters with a distribution of Ag atoms rather than a spe- gold catalysts by aberration-corrected scanning transmission elec-
cific number. As an example, the total structure of Agx Au25-x (SR)18 tron microscopy and observed 0.5 nm bilayer Au clusters containing
(where R = C6 H11 ) is shown in Fig. 2 [54], in which the central atom about 10 gold atoms were responsible for the high activity for CO
remains to be gold, whereas the icosahedral shell is composed of Au oxidation. Besides, experimental and theoretical results showed
and Ag (Fig. 2a), and heavy doping can also lead to Ag distribution that partial electron transfer from the support surface to the gold
into the external staple motifs (Fig. 2b). cluster and oxygen-vacancy/F-center defects of MgO accounted for
Besides the common method of mixed metal salt precursors for the unusual catalytic activity of Au8 /MgO in the low-temperature
bimetallic nanocluster synthesis, a novel metal-exchange method oxidation of CO [63,64]. The aforementioned tailoring of the prop-
using gold nanoclusters as the template for bimetallic nanoclus- erties of small metal clusters by altering the cluster size and its
ters with atomic precision have also been reported [54–58], in support has proven to be quite useful in design of nanostructured
which heterometal-thiolate complexes react with the presynthe- materials for catalytic applications.
sized homogold nanoclusters. In addition, Pradeep and coworkers The advent of Au nanoclusters with uniform size and well-
have also reported an inter-cluster reaction method for obtaining defined structure has provided fundamental insights into Au
bimetallic nanoclusters [59]. catalysis. For instance, the support-dependent activity of
thiolate-ligated Au25 (SR)18 (where R = CH2 CH2 Ph) nanoclus-
ters for CO oxidation reaction was observed in the order:
Catalytic properties
Au25 (SR)18 /TiO2 < Au25 (SR)18 /Fe3 O4 < Au25 (SR)18 /CeO2 (Fig. 3A)
[65]. Surprisingly, Au25 (SR)18 /TiO2 had no catalytic activity even
Homogold nanoclusters
if the temperature was up to 200 ◦ C. This is significantly different
from conventional Au/TiO2 , which was considered to be one of the
Oxidation
most active catalysts [8]. The catalytic activity of Au25 (SR)18 /CeO2
CO oxidation. CO oxidation is the model reaction most widely stud-
was found to be related to the pretreatment atmosphere and
ied due to its simplicity and significance [60]. More importantly, it

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 5

Fig. 4. (A) Amount of CO2 production during the copulse (black squares) and sequential pulses (red squares) of CO and O2 at 80 ◦ C on the Au25 (SR)18 /CeO2 rod sample
pretreated at 150 ◦ C in O2 along with their ratios (blue squares). (B) Percentage of MvK and L−H reaction pathways on Au25 (SR)18 /CeO2 rod catalysts pretreated at different
temperatures with O2 . Reproduced with permission from ref [70]. Copyright 2014 American Chemical Society.

temperature. The O2 pretreatment at 150 ◦ C markedly increased vinyl-2-pyrrolidone) (PVP) protected Au nanoclusters of less than
the activity, but higher temperatures (above the thiolate desorp- 1.5 nm were found to exhibit superior activity for the aerobic oxida-
tion temperature) led to no further increase in activity (Fig. 3B) tion of alcohols compared with larger size Au nanoclusters [71]. The
[65]. In addition, the thermal pretreatment in N2 had no signif- catalytically active Au nanoclusters were revealed to be negatively
icant effect on its activity. Characterizations of Au25 (SR)18 /CeO2 charged via electron donation from PVP. The increasing electronic
indicated that thiolate ligands remained on Au nanoclusters after density of Au nanocluster with decreasing size led to the enhanced
the pretreatment at 150 ◦ C. In others words, CO oxidation can catalytic activity. The electron was transferred from the anionic Au
be catalyzed by intact Au25 (SR)18 supported on CeO2 , which is nanoclusters to LUMO of O2 to generate superoxo or peroxo-like
also reported by the latest work [66]. These showed that O2 species and subsequently facilitated the oxidation of alcohols. The
pretreatment was closely relevant to the product of active oxygen plausible mechanism was illustrated in Scheme 4 [71]. It is worth
species on CeO2 instead of ligand removal of Au nanoclusters. The nothing that PVP not only acts as a stabilizer but also modifies the
catalytic properties of Au38 (SR)24 [67,68] and Au144 (SR)60 [69] for electronic structure of Au clusters, which are different from the role
CO oxidation have also been evaluated. of the thiolate ligand in Au25 (SR)18 /CeO2 [70].
It is noteworthy that thiolate ligands have a substantial influ- Yoskamtorn et al. [72] synthesized Au25 (SC12 H25 )18 on hierar-
ence on catalytic activity of gold nanoclusters for CO oxidation chically porous carbon sheets by the removal of ligands through
[65–69]. For instance, the intact Au38 (SCH2 CH2 Ph)18 /CeO2 exhib- calcination under vacuum. The as-prepared catalysts were eval-
ited low catalytic activity, while the Au catalysts after partial uated for the selective oxidation of benzyl alcohol. Interestingly,
removal of thiolates showed better activity. Surprisingly, complete the selectivity for benzaldehyde was significantly improved when
removal of ligands led to a decrease in activity [64]. To disclose the the amount of residual thiolates on Au nanoclusters increased,
role of thiolate ligands, Wu et al. [70] meticulously investigated the although the activity was reduced, as shown in Fig. 5. TEM dis-
catalytic property of Au25 (SR)18 /CeO2 -rod (where R = SCH2 CH2 Ph) closed that the size of Au nanoclusters remained almost unchanged
for CO oxidation reaction by a series of experiments and DFT cal- after thermal pretreatment and EXAFS characterization indicated
culations. The intact Au25 (SR)18 /CeO2 -rod catalyst had no activity that Au nanoclusters with residual thiolates was more positively
due to the lack of exposed Au atoms for CO adsorption. As the thi- charged than the bare Au due to electron transfer from Au to S.
olate ligands were partially removed, starting from the interface Therefore, the presence of thiolates on Au nanoclusters not only
between Au nanoclusters and CeO2 at 150 ◦ C and above, CO was reduced the oxidation of benzyl alcohol to benzaldehyde by elec-
favorably adsorbed on exposed Au atoms and then reacted with tron withdrawal but also suppressed the esterification reaction
the lattice oxygen of CeO2 to generate CO2 . As ligands removal pro- by site isolation effect. It confirmed that the catalytic proper-
gressed to the top surface sites of Au nanoclusters, CO and O2 were ties of metal nanoclusters can be tuned through modulating their
also co-adsorbed on exposed Au atoms and then led to CO oxi- electronic structure and the steric environment by chemical mod-
dation. Isotopic labeling and pulse experiments showed that CO ification.
oxidation proceeded predominantly at the interfaces between the Subsequently, the water-soluble glutathione protected gold
ligands, Au, and CeO2 via Mar-van Krevelen mechanism by utiliz- nanoclusters supported on M3 Al-layered double hydroxide (LDH,
ing the lattice oxygen of the oxide support (Fig. 4). Activation of M = Mg, Ni, Co) were prepared, calcined and applied for the
CO and O2 at the exposed Au atoms made minor contributions to oxidation of alcohols without basic additives [73]. AuNCs/Mg3 Al-
CO oxidation via Langmuir-Hinshelwood mechanism (Fig. 4). The LDH-0.23 (where 0.23 is the gold content, similar hereafter) showed
mechanism of CO oxidation by Au25 (SR)18 /CeO2 -rod catalysts was far higher properties in the oxidation of 1-phenylethanol than
proposed (Scheme 3). It showed that the thiolate ligand acted as Au/Mg3 Al-LDH by the classical deposition-precipitation method.
a double-edged sword for gold nanoclusters in CO oxidation: On The activity of AuNCs/Ni3 Al-LDH-0.22 and AuNCs/Co3 Al-LDH-0.20
one hand they retain the structural integrity of Au nanoclusters, could be further enhanced by the stronger synergy between
but they also block the catalytic sites for CO oxidation. The results Au and the supports containing transition metal cations [73,74].
clearly provided the fundamental understanding of how the ligands Phosphine-capped Au nanoclusters as the precursor for preparing
affected the catalysis of Au nanoclusters in CO oxidation. ultra-small and uniform Au nanocatalysts were also reported [75].

Benzyl alcohol oxidation. Au nanoclusters are also an efficient Styrene oxidation. The size evolution of ultra-small Au nanoclus-
catalyst for the selective oxidation of alcohols. For example, poly(N- ters provided a good opportunity for the selective oxidation of

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
6 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Scheme 3. CO Oxidation Mechanism on Intact, Partially and Fully Dethiolated Au25 (SR)18 /CeO2 Rod Catalysts. Reproduced with permission from [70]. Copyright 2014
American Chemical Society.

Scheme 4. The proposed mechanism for the activation of O2 by Au:PVP. Reproduced with permission from ref [71]. Copyright 2009 American Chemical Society.

Fig. 5. Selectivity and conversion during benzyl alcohol oxidation over various supported Au25 NCs. Reproduced with permission from ref [72]. Copyright 2014 American
Chemical Society.

styrene, which is an important chemical transformation process. n = 25, 38, 144) in styrene oxidation under three different oxidation
Turner et al. [76] found that triphenylphosphane stabilized Au55 conditions: (a) tert-butyl hydroperoxide (TBHP) as the oxidant, (b)
(not well-defined) clusters were an efficient and robust catalyst TBHP as the initiator and O2 as the oxidant, and (c) O2 as the oxidant.
for the selective oxidation of styrene by O2 . The particle size of The activity of three systems showed in the order: (a) > (b) > (c).
Au55 /BN and Au55 /SiO2 remained unchanged (∼1.4 nm), and both Using O2 as the oxidant, Au25 (SR)18 showed the highest activity
catalysts achieved about 20% conversion of styrene and 80% selec- among the different sizes and it was related to its particularly effi-
tivity to benzaldehyde. In particular, only Au nanoparticles with cient activation of O2 . No drastic size effect was found when TBHP
less than 2 nm in size were found to have activity, while tradi- was used as the oxidant because of its facile activation by Au, while
tional supported gold nanoparticles larger than 2 nm were inactive, O2 activation is harder. In comparison with SiO2 , HAP had the role
indicating a sharp size threshold in catalytic activity. The observed to stabilize Au nanoclusters and then promoted the activity. The
activity originated from the modified electronic structure of Au calcined catalysts exhibited higher activity than the counterparts
entities and the weak adsorption of styrene on Au nanoclusters. without calcination. The proposed mechanism for the selective oxi-
Zhu et al. [77,78] investigated the catalytic activity of dation of styrene catalyzed by Au25 (SR)18 is illustrated in Fig. 6
Aun (SR)m /SiO2 and Aun (SR)m /HAP nanoclusters (R = CH2 CH2 Ph, [77].

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 7

Fig. 6. Proposed mechanism of selective oxidation of styrene catalyzed by a Au25 (SR)18 cluster. The thiolate ligands are not shown. (Dark gray: Au atoms of the core, light
gray: Au atoms of the shell). Reproduced with permission from ref [77]. Copyright 2010 Wiley-VCH.

Although it is confirmed that small amounts of P and Cl had the understanding of the ligand effect on the catalytic activity and
no negative effect on the activity of supported Au55 nanoclus- the design of Au catalysts with high performance.
ters [76], thiolate ligands were generally considered to restrict the
accessibility of reactants to Au atoms and thereby deactivate Au Cyclohexane oxidation. Selective oxidation of cyclohexane to cyclo-
nanoclusters [77,79]. Therefore, Liu et al. [79] removed the thiolate hexanone and cyclohexanol is one of the most important reactions,
ligands through calcination at 300 ◦ C and studied the activity of the as the resulting compounds serve as the essential intermediates in
as-obtained Au25 /HAP. The ligand-off Au nanoclusters preserved the nylon-related chemical industry. Liu et al. explored the cat-
the cluster size with a flattened morphology. 100% conversion of alytic performances of Au10 (SG)10 , Au18 (SG)14 , Au25 (SG)18 , and
styrene and 92% yield of styrene oxide were obtained. It was noted Au39 (SG)24 (where SG represents glutathionate) nanocluster sup-
that the high performance was attributed to the choice of solvent ported on hydroxyapatite(HAP) as the precursors for the aerobic
(benzene) and oxidant (TBHP). oxidation of cyclohexane [82]. TEM revealed no appreciable aggre-
Thermal removal of ligands is a destructive approach, which gation of Au nanoclusters after the calcination due to the strong
may lead to the sintering of Au nanoclusters. Two strategies for mild interaction between Au and PO4 3− moieties of HAP. Dramatic size-
removal of thiolate ligands were developed recently, i.e. reductive dependent activity of Au nanoclusters was observed, in which Au39
deprotection and oxidative stripping [80,81]. Mercaptopropyl- had the highest activity with 14.9% of cyclohexane conversion and
modified SBA-15 was employed as the support to obtain the 99% selectivity to cyclohexanone and cyclohexanol (Fig. 8) [82]. The
supported Au25 (SCH2 CH2 Ph)18 and Au144 (SCH2 CH2 Ph)60 catalysts. activity of supported Au nanoclusters was related with the support
Controlled removal of thiolate ligands was accomplished by treat- material and the treatment temperature [83].
ing the above catalysts with the different concentration NaBH4
solution (Fig. 7) [80]. Characterizations showed that the Au size Hydrogenation
remained almost no change with the efficient removal of lig- Selective hydrogenation is one of the most important reactions
ands. 100% conversion and 100% selectivity for benzaldehyde were heavily applied in industry. The reaction is generally catalyzed by
achieved in the oxidation of styrene over such catalysts, in com- VIII group metals such as Ni, Ru, Pd and Pt [1]. The pioneering work
parison with the performances of the untreated catalysts with 30% by Bond and co-workers revealed that supported gold nanoparti-
conversion and 100% selectivity for benzaldehyde [80]. cles exhibited excellent selectivity in the gas-phase hydrogenation
Another strategy for ligand removal has been developed by of 1, 3-butadiene to butane [84]. Despite the impressive results,
treatment with TBHP or H2 O2 as the oxidant [81]. The thiolate lig- gold nanoparticle catalysts for selective hydrogenation was mostly
ands were removed after oxidation, while the size and intergity of ignored in subsequent years until the supported gold nanoparti-
Au nanoclusters had almost no change. The as-prepared Au clusters cles with the size of 3–5 nm was found to promote the selective
exhibited an enhanced activity in the selective oxidaiton of styrene. hydrogenation of carbonyl or nitro groups in the presence of other
Of note, 4-mercaptobenzoic acid was not removed by this method reducible functional groups at high conversion [85,86]. Nowadays
[81]. gold nanoparticles [87] and Aun (SR)m nanoclusters [15] have also
Both the reduction and oxidation methods allow the efficient been demonstrated to be capable of catalyzing the selective hydro-
removal of ligands from the metal nanoclusters, which promotes genation reactions.

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
8 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Fig. 7. Illustration of enhanced catalytic activity in selective oxidation reaction with supported ultrasmall organothiolate-protected AuNP catalysts by mild chemical stripping
of their surface ligands. Reproduced with permission from ref [80]. Copyright 2014 Wiley-VCH.

Scheme 6. The proposed mechanism of the chemoselective hydrogenation of ␣,␤-


unsaturated ketone to unsaturated alcohol catalyzed by Au25 (SR)18 nanoparticles.
The thiolate ligands are not shown. Dark grey: Au atoms of the core; light gray:
Au atoms of the shell. Reproduced with permission from ref [91]. Copyright 2010
Wiley-VCH.

Fig. 8. TOF values as a function of the cluster size n in the selective oxidation of ters were also prepared and found to show high activity and nearly
cyclohexane. Reproduced with permission from ref [82]. Copyright 2011 American complete selectivity in the hydrogenation of ␣, ␤-unsaturated alde-
Chemical Society.
hydes under mild conditions [93]. These studies suggested that
the complete selectivity for unsaturated alcohols can be attained
Hydrogenation of carbonyl compounds. Chemoselective hydrogena- by employing ligand-on gold nanoclusters as the catalyst. Fur-
tion of unsaturated carbonyl compounds while maintaining the thermore, the mechanism of 100% selective hydrogenation of ␣,
reducible C C group is an important kind of reaction because the ␤-unsaturated ketones to unsaturated alcohols on Au25 (SR)18 have
resulting unsaturated alcohols are valuable chemical intermediates been theoretically unraveled by first-principles density functional
in the production of fragrances, pharmaceuticals and agrichemicals calculations [94].
[88] (Scheme 5). Although gold catalysts exhibited high selec- When the aldehyde group and nitro group co-exist in
tivity for unsaturated alcohols, 100% chemoselectivity cannot be the same substrate, the chemoselective hydrogenation of nitro
achieved by using the conventional gold nanoparticles and bare group was always preferred over conventional gold nanopar-
gold nanoclusters catalyst [86,89]. ticles and bare gold nanocluster catalysts [86,87,89]. However,
Zhu et al. [90–92] obtained 100% selectivity for unsaturated Li et al. [95] reported that water-soluble Aun (SR)m (where
alcohols in the hydrogenation of a range of ␣, ␤-unsaturated SG represents glutathionate) nanoclusters could achieve 100%
ketones (except crotonaldehyde) over Au25 (SR)18 catalysts. For selective hydrogenation of aldehyde group with moderate con-
instance, using supported Au25 (SCH2 CH2 Ph)18 as the catalyst, ben- version in the homogeneous chemoselective hydrogenation of
zalacetone in a mixture of toluene and ethanol was stirred at 0 ◦ C for 4-nitrobenzaldehyde in water. A drastic size dependence and lig-
3 h under 1 atm H2 , and the complete selectivity was obtained with and steric effect were observed. Larger gold nanoclusters or the less
40% conversion of benzalzcetone [91]. The catalyst showed only bulky ligand protected gold nanoclusters are found to have higher
a slight decrease in activity with 100% selectivity for unsaturated catalytic activity (Fig. 9).
alcohols after six circles. The proposed mechanism for the reac- Furthermore, using hydrophobic thiolate ligand-on gold nan-
tion process is illustrated in Scheme 6 [91]. The electron-deficient oclusters supported on oxides like Au99 (SPh)42 /CeO2 as the
Au12 shell is considered to provide sites for H2 adsorption and dis- catalyst, about 93% conversion of 4-nitrobenzaldehyde and 100%
sociation. After dissociation, the H atom is adsorbed in a bridge selectivity for 4-nitrobenzyl alcohol were attained in the presence
position bonding to two low-coordinated gold atom. The electron- of pyridine using water as the solvent and H2 as the hydro-
rich Au13 core favors adsorption of C O via the interaction with the gen source [96]. The catalyst also showed excellent durability
Au atom and the O in carbonyl group, with partial electron transfer (no loss of catalytic reactivity after three circles) and good ver-
from Au13 to the O atom. Subsequently, the weakly nucleophilic satility for a range of substrates with aldehyde and nitro groups
attacks the activated C O group, thereby leading to the production (Table 1). The complete chemoselective hydrogenation of the alde-
of unsaturated alcohols. hyde group is remarkable, which is in sharp contrast with the
In addition to thiolate ligated gold nanoclusters, other ligands previously reported chemoselective hydrogenation of the nitro
like tert-butyl(naphthalen-1-yl)phosphine oxide ligated gold clus- group against the aldehyde group by conventional gold nanoparti-

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 9

Scheme 5. Chemoselective reductions of unsaturated carbonyl compounds to unsaturated alcohols.

Table 1
Chemoselective Hydrogenation of a Range of Substrates with Nitro and Aldehyde Groups Using Au99 (SPh)42 /CeO2 as the Catalyst.

a
Reproduced with permission from ref [96]. Copyright 2014 American Chemical Society.

cles and bare gold nanocluster catalysts [86,89]. These new, unique wise detachment of an “Au-SR” unit. This led to the exposure of
catalytic properties were possibly related to the selective adsorp- Au13 -core to reactants and possibly provided the catalytic sites.
tion of the aldehyde group onto the surface gold atoms of Aun (SR)m Besides, Lewis acids bound on the Au13 -core possibly functioned
nanoclusters [96]. More details on the mechanism require future as the different active sites for the hydrogenation reactions [97].
experimental and theoretical works. In gold nanoparticle and nanocluster catalyzed hydrogenation
Subsequently, Li et al. [97] found that the addition of a Lewis reactions, pyridine was usually added and acted as an efficient
acid markedly enhanced the catalytic performance of Au25 (SR)18 additive for activation of reactants [96–99], but it is toxic and
(where R = CH2 CH2 Ph) for the chemoselective hydrogenation of 4- unfriendly even at low concentrations. Recently, the diphenyl-2-
nitrobenzaldehyde and its derivatives in the presence of base under pyridylphosphine ligand protected gold clusters were designed
mild conditions. Among the Lewis acids tested, Co2+ yielded the and found to have excellent properties in the chemoselective
best catalytic property. The conversion was significantly improved hydrogenation of nitrobenzyldehyde to nitrobenzal alcohol in the
to 90.1% at 50 ◦ C when Co(OAc)2 was added, while only 13.4% con- absence of pyridine [99]. It avoided the use of toxic pyridine by
version was obtained without a Lewis acid. The above catalytic introducing the pyridine group close to gold surface to assist the
system also showed good recyclability with no appreciable loss of reactant activation.
catalytic reactivity after three circles. In the presence of Lewis acid,
new species Au24 (SR)17 was generated, which originated from the
adsorption of Mz+ onto “-SR-”of Au25 (SR)18 and the resulting step- Hydrogenation of nitro compounds. Aniline and its derivatives are
important chemical intermediates for pharmaceuticals, dyes, agro-

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
10 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

olate ligated Au25 (SR)18 displayed better accessibility and activity


than those bearing a longer chain. The aromatic ligands increased
the accessibility of Au25 (SR)18 and thus led to an enhanced activity
[103]. Recent studies indicated that ligands modulated the catalytic
active-sites of gold nanoclusters and then directed the reaction
pathway for the hydrogenation of 4-nitrophenol [104]. It suggested
again that ligand engineering can be an efficient tool for tuning the
catalytic property of gold nanoclusters.
However, other studies showed that thiolate protected Au nan-
oclusters supported on oxides and HAP were inactive for the
hydrogenation of nitro group with H2 [105]. After the removal
of thiolates, both bared Au25 /HAP and Au/TiO2 displayed good
activity for the hydrogenation of nitrobenzene to aniline, in agree-
Fig. 9. Hydrogenation of 4-Nitrobenzaldehyde Catalyzed by Unsupported Aqueous ment with the earlier literature [86,89]. Especially, gold catalysts
Soluble Aun (SG)m Nanoclusters as Catalysts. Reproduced with permission from ref with well-controlled size (2.0 nm) were obtained through the load-
[95]. Copyright 2014 American Chemical Society. ing of cysteine-capped Au25 nanoclusters onto ZnAl-Hydrotalcite
(ZnAl-HT) and then the calcination of the resulting precatalysts
(300–500 ◦ C) (Fig. 10). It was unprecedented that the catalyst
chemicals and polymers. They are generally produced through the exhibited excellent selectivity (>98%) to 3-vinylaniline at complete
hydrogenation of nitro compounds over metal catalysts [100]. Due conversion of 3-nitrostyrene over broad reaction duration and tem-
to good catalytic activity and unparalleled selectivity, gold nanocat- perature windows (Fig. 11) [106]. No loss of catalytic activity and
alysts have attracted intensive attention [85–87]. It was reported selectivity was found after five circles for 3-nitrostyrene hydro-
that ligand-on gold nanoclusters were catalytically active for the genation. The outstanding properties possibly depend on the small
reduction of 4-nitrophenol to 4-aminophenol by NaBH4 [101–103]. size of Au and unique nature of oxide support derived from ZnAl-HT
It also showed that Au25 (SR)18 retained the structural integrity [106].
during the catalytic process and was stable and recyclable [102].
Moreover, the nature of ligand had a significant impact on the
catalytic property of gold nanoclusters [103]. The shorter chain thi-

Fig. 10. HAADF-STEM images of the catalysts: a) Au25 /ZnAl-HT; b) Au25 /ZnAl-HT-300; c–d) Au25 /ZnAl-HT-600. Reproduced with permission from ref [106]. Copyright 2017
Wiley-VCH.

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 11

Table 2
Homocoupling of aryl iodides using the Au25 (SR)18 /CeO2 catalyst a.

Fig. 11. The conversion, selectivity, and TOF values (line) over the Au25 /ZnAl-HT
catalysts calcined at different temperatures. The TOF value was measured below 20%
conversion.Reproduced with permission from ref [106]. Copyright 2017 Wiley-VCH.

a
Adapted with permission from ref [111]. Copyright 2012 Royal Society of Chem-
Semihydrogenation of alkynes. The semihydrogenation of alkynes is istry.

a valued approach to alkenes, which are useful building blocks of


fine chemicals [107]. Gold nanoparticles were reported to be an Carbon–Carbon coupling reaction
efficient catalyst for the reaction [87,108]. Li et al. [109] inves- The carbon–carbon coupling reactions are the mostly used
tigated the catalytic properties of the spherical Au25 (SR)18 and approaches to prepare biaryl or propargylamines, which are valu-
rod-shaped Au25( PPh3 )10 (C CPh)5 X2 nanoclusters supported on able synthetic building blocks in the manufacture of biologically
oxides for sehihydrogenation of terminal alkynes to alkenes. About active compounds, natural products, and pharmaceutical organic
100% conversion of the alkynes with complete selectivity to alkene compounds [110]. Gold nanoparticle catalysts have been employed
products was achieved under mild conditions. The open triangular in the carbon–carbon coupling reactions [110]. Likewise, gold nan-
Au3 facet on the spherical Au25 (SR)18 (R = CH2 CH2 Ph) nanoclus- oclusters have good catalytic activity [111–116]. For example, Li
ters and the waist sites of the rod-shaped Au25 (PPh3 )10 (C CPh)5 X2 et al. found that the Au25 (SR)18 /CeO2 (R = CH2 CH2 Ph) nanocluster
(X = Br/Cl) nanoclusters were considered to be the active sites catalyst exhibited high activity in the homocoupling of aryl iodides
in the semihydrogenation process. The mechnism of the semi- (Table 2) [111]. It also showed good stability since only a slight
hydrogenation process was proposed [109], see Fig. 12. Besides, decrease in catalytic activity was observed after five circles (within
the semihydrogenation of the internal alkynes did not proceed 5%). The positively charged surface Au atoms of Au25 (SR)18 were
over the ligand-on nanoclusters, while they can be catalyzed by considered to be catalytic sites for activating iodobenzene, while
the ligand-off nanoclusters. Combining with FT-IR spectrocopy, a the electron-rich, redox-active Au13 kernel participated the elec-
unique activation pathway of terminal alkynes was identified. The tron transfer process in the catalytic reaction.
terminal alkynes followed a deprotonation activation pathway by a Au25 (SR)18 /CeO2 also achieved about 96% conversion
Rı́-C C-[Aun Lm ] (where L represents the ligands on the cluster) and of iodoanisole and 88% selectivity for 1-methoxy-4-(2-
it was in contrast with the case of conventional gold nanoparticles phenylethynyl)benzene in the sonogashira cross-coupling reaction
catalysts. between iodobenzene and phenylacetylene [112]. DFT calcina-

Fig. 12. Proposed mechanism for the ligand-on Au25 nanocluster catalyzed semihydrogenation of terminal alkynes to alkenes using H2 (20 bar). Left panel: Au25 (SR)18
nanocluster. Right panel: Au25 (PPh3 )10 (C CPh)5 X2 (X = Cl, Br) nanocluster was drawn according to the crystal structure of the Au25 (PPh3 )10 (C CPh)5 X2 nanocluster. Color
code: Au, green; S, yellow; C, gray; P, pink; X, cyan. Hydrogen atoms are omitted for clarity. The areas marked in orange are the Au3 active sites (left panel) and the waist
active sites (right panel) in the alkyne semihydrogenation reaction, respectively. Reproduced with permission from ref [109]. Copyright 2014 American Chemical Society.
(For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
12 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Fig. 13. (a) Top view of one of the two open facets of Au25 (SCH2 CH2 Ph)18 clusters where three external gold atoms are exposed; the other facet is on the back side. (b)
Side view of the two facets which are at top and bottom. (c) Top view and (d) side view of the co-adsorption of phenylacetylene and iodobenzene on the surface of the
Au25 (CH2 CH2 Ph)18 clusters. Au, yellow; S, blue; C, gray; H, white; I, green. Reproduced with permission from ref [112]. Copyright 2013 Elsevier.

tions suggested that iodobenzene and phenylacetylene together


adsorbed on the open gold atoms of Au12 shell: iodobenzene
adsorbed one gold atom, phenylacetylene adsorbed the second
gold atom, and the C CH group of phenylacetylene and the −I
group of iodobenzene pointed towards the third atom (Fig. 13).
The co-adsorption configuration, in which its adsorption energy is
about the sum of the individual adsorption energies, favored the
coupling between reactants and led to the product. The supported
Au25 (SR)18 was also applied in the Suzuki crossing coupling
reaction [113].
Besides, Au nanoclusters such as Au38 (SR)24 (where
R = CH2 CH2 Ph) were found to be an efficient catalyst in the
three-component coupling of an aldehyde, an alkyne and an
amine to synthesize propargylamines [114]. Oxides as the support
made no contribution to the catalytic performance of Au38 (SR)24 .
Furthermore, ligand-off Au38 /CeO2 had a lower catalytic activity
than ligand-on Au38 (SR)24 nanoclusters (Fig. 14), in comparison Fig. 14. Conversion of benzaldehyde, phenylacetylene and piperidine as a function
with the traditional view that the more bare surface atoms exhib- of reaction time over Au38 (SCH2 CH2 Ph)24 , Au38 (SCH2 CH2 Ph)24 /CeO2 , and ligand-off
Au38 /CeO2 catalysts. Reproduced with permission from ref [114]. Copyright 2016
ited higher catalytic property. It demonstrated the importance of
Royal Society of Chemistry.
the entire structure of Au38 (SR)24 nanoclusters, indicating that
the synergistic effect of the electron-deficient Au␦+ (0 < ␦+ < 1)
shell and the electron-rich Au23 core was critical for the catalytic Au25 (SCH2 CH2 Ph)18 and Au25 (S-Nap)18 clusters have the same
reaction. structural framework, although a slight expansion of the Au13
Recently, Li et al. [116] prepared an aromatic-thiolate- kernel induced by the aromatic-thiolate was observed in Au25 (S-
protected gold nanoclusters Au25 (S-Nap)18 (where S-Nap = 1- Nap)18 nanoclusters. More importantly, the catalytic activity and
naphthalenethiolate) and observed that aromatic ligands had selectivity of Au25 (SR)18 clusters (R = Nap, Ph, CH2 CH2 Ph, n-
distinct effects on the electronic and catalytic properties. Both C6 H13 ) are largely influenced by the chemical nature of thiolate
ligands in the Ullmann heterocoupling reaction. Among them,

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 13

Fig. 15. The photocatalytic mechanisms of Au25 (SR)18 /TiO2 under visible light irra-
diation. Reproduced with permission from ref [118]. Copyright 2013 American
Chemical Society. Fig. 16. Time course of water splitting over (red) Au25 –BaLa4 Ti4 O15 photocata-
lyst (0.1 wt% Au) prepared by the present method and (black) AuNP–BaLa4 Ti4 O15
photocatalyst (0.5 wt% Au) prepared by the conventional photodeposition method.
Au25 (S-Nap)18 /CeO2 gave the best catalytic performances with 91% Reproduced with permission from ref [124]. Copyright 2013 Royal Society of Chem-
istry. (For interpretation of the references to colour in this figure legend, the reader
conversion and 82% selectivity for the heterocoupling product. It
is referred to the web version of this article.)
showed the effective of ligand engineering for tailoring the catalytic
properties of nanoclusters.
unique electronic structure and framework. It indicates that visible
Photocatalysis light photocatalytic selective organic transformations proceed
Thiolate-protected gold nanoclusters possess molecule-like, efficiently over gold nanoclusters, which is also confirmed by the
discrete electronic transitions in optical properties (such as visible recent literature [122].
light absorption), which are very different from the optical prop- Besides, gold nanoclusters also have an enormous potential
erties of larger nanoparticles [117]. What is more, these properties application in clean hydrogen production [123,124]. For instance,
can be tuned through the exact atom number, composition and lig- Negishi et al. [124] investigated glutathione-protected Au25 clus-
ands, so they hold great potential for applications in the fields of ters (1.2 ± 0.3 nm) onto BaLa4 Ti4 O15 as precatalysts for water
photocatalysis, energy and chemical sensors [117–124]. Here we splitting. The photocatalytic property of as-prepared material was
discuss briefly their applications in photocatalysis. determined to be 2.6 times higher than that of BaLa4 Ti4 O15 loaded
Yu et al. [118] reported that loading of Au25 (SR)18 with larger gold nanoparticles (10–30 nm) by the conventional
(R = CH2 CH2 Ph) nanoclusters onto TiO2 gave rise to a 1.6 times photo-deposition method (Fig. 16). The enhanced activity was asso-
increase in the visible light photocatalytic activity for the degra- ciated with the increased surface area-to-mass ratio and specific
dation of methyl orange. However, Au25 (SR)18 /TiO2 offered no electronic structure of ultra-small gold clusters.
drastic increase in photocatalytic activity under UV light irradiation
compared with plain TiO2 . The enhanced photocatalytic activ- Gold-based bimetal nanoclusters
ity may be related to quantum confinement within Au25 (SR)18
nanoclusters with a HOMO-LUMO gap of Eg ∼ 1.3 eV, which is In the subnanometer regime, every atom can exert a substan-
similar to a small-band-gap semiconductor. Under the visible tial effect on the catalytic properties, thus precise control over the
light irradiation, the photogenerated electrons from Au25 (SR)18 size of a nanocluster has been proven to be an efficient approach
were transferred to the conduction band of TiO2 , thereby leading for tuning the activity and/or the selectivity in a catalytic process
to the electron-hole efficient separation. The produced hole can [8,61,62,82,95]. Besides, heteroatom substitution in homometal
further generate hydroxyl radicals, which are responsible for the nanoclusters also provides good opportunities to tune the catalytic
decomposition of the dye. Singlet oxygen (1 O2 ), generated from properties of nanoclusters at the atomic level [20]. Herein, we show
Au25 (SR)18 nanoclusters under visible light, also participated in the catalytic applications of the heteroatom-doped Au25 (SR)18 clus-
the decomposition of the dye. Two mechanisms were proposed ters.
to explain the significantly promoted photocatalytic activity of
Au25 (SR)18 /TiO2 under visible light (Fig. 15) [118].
Oxidation
Sakai et al. found that Au25 (SG)18 clusters (where SG represents
CO oxidation. Li et al. [125] investigated Ag and Cu
glutathionate) was efficiently adsorbed onto the surface of TiO2 via
doped Au25 (SR)18 (where R = CH2 CH2 Ph) nanoclusters sup-
the carboxylic group in GSH. The photogenerated electrons from
ported on CeO2 as catalysts for CO oxidation. These
Au25 (SG)18 clusters were injected into the TiO2 conduction band
catalysts exhibited the catalytic activity in the order of
with 60% internal quantum yield under visible light irradiation
Cux Au25-x (SR)18 > Au25 (SR)18 > Agx Au25-x (SR)18 . The exposed
[119]. The photocatalytic material was used in the photooxidation
metal atoms (Au, Ag and Cu) in the clusters were considered to
of phenol derivates and ferrocyanide, and reduction of Ag+ , Cu2+
be the catalytic active sites. The observed activity is related to
and dissolved oxygen [119,120].
the ability of CO adsorption and metal exchange between the
Subsequently, Chen et al. [121] prepared
icosahedral core and the staple motif of the clusters at higher
Au25 (PPh3 )10 Cl2 (SC3 H6 SiO3 )5 /TiO2 by hydrolysis of
temperature. This example demonstrates the effectiveness of
Au25 (PPh3 )10 Cl2 (SC3 H6 SiO3 )5 in the presence of P25 support.
tuning the catalytic property of gold nanoclusters by heteroatom
The as-obtained catalyst exhibited good photocatalytic properties
doping at the atomic level. Further research is needed to explore
for the selective oxidation of amines to imines under visible light.
the details of CO oxidation pathway and how the amount of doping
The turnover frequency of methylbenzylamine oxidation was
may influence the activity.
1522 h−1 , which is considerably higher than that of conventional
gold catalysts (Table 3). The enhanced activity was related to its

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
14 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

Table 3
Photocatalytic oxidation of a range of amine substrates over [Au25 ]/P25 catalyst a.

a
Adapted with permission from ref [121]. Copyright 2017 American Chemical Society.

Fig. 17. Catalytic Performances of Au25 /CNT and Pd1 Au24 /CNT for Benzyl Alcohol
Oxidation. Reproduced with permission from ref [126]. Copyright 2012 American
Chemical Society.

Benzyl alcohol oxidation. Compared with homogold Au25 nanoclus-


ters, single-palladium doped Pd1 Au24 catalysts [126] displayed Fig. 18. Catalytic performances of Au34 Ag28 (PhC C)34 supported on XC-72 in the
hydrolytic oxidation of triethylsilane before and after thermal treatment at dif-
higher activity in the benzyl alcohol oxidation reaction (Fig. 17).
ferent temperatures. Reproduced with permission from ref [133]. Copyright 2016
The former gave a 22% conversion of benzyl alcohol, while the lat- American Chemical Society.
ter offered a significantly higher conversion (74%). TEM imaging
showed that both clusters had a similar size in the range 1.1–1.2 nm.
Styrene oxidation. It was reported that the heteroatom doped gold
The remarkably enhanced activity was ascribed to the modulation
nanoclusters as the catalyst efficiently facilitated the transfor-
of the electronic structure by electron transfer from Pd to Au [126]
mation of styrene oxidation and/or tuned the selectivity for the
(Fig. 18).
products [129–131]. The Pt1 Au24 (SR)18 /TiO2 catalyst was found
Cd1 Au24 (SR)18 nanoclusters were prepared by the metal
to exhibit higher catalytic performance than the homogold Au25
exchange method and characterized by x-ray single crystal diffrac-
nanocluster in the selective oxidation of styrene [129]. The con-
tion [55]. The Cd1 Au24 (SR)18 nanoclusters inherits the same
version of styrene and the selectivity for benzaldehyde were
core-shell structure as that of Au25 (SR)18 and the Cd is in the center
90.8% and 89.9% over the doped clusters catalysts, compared
of the 13-atom icosahedral core. It was found that the Cd dop-
with 58.9% styrene conversion and 54.0% benzaldehyde selec-
ing significantly enhanced the stability of the nanocluster under
tivity over the Au25 catalyst, respectively [129]. In recent work,
O2 environment [127]. More importantly, the Cd1 Au24 (SR)18 nan-
Wang et al. synthesized 70-atom Ag46 Au24 bimetallic nanoclus-
oclusters supported on CNT exhibited higher activity than the
ters, which consisted of a 40-atom bimetallic Ag2 @Au18 @Ag20 core
corresponding homogold catalyst in the selective oxidation of
protected by a Ag24 Au6 (SR)32 shell [130]. The bimetallic nanoclus-
benzyl alcohol (63% and 33% conversion of benzyl alcohol, respec-
ters showed higher selectivity for benzaldehyde (96.3%) than that
tively), but both catalysts exhibited similar selectivity for the
of homogold Au25 (SR)18 nanoclusters (66.4% only), but with sim-
benzaldehyde product [127]. Besides, the Cu doping into Au nan-
ilar conversion of styrene in the styrene oxidation reaction. The
oclusters was reported to enhance the catalytic property in the
surface composition and structure of the bimetallic nanoclusters
selective oxidation of benzyl alcohol [128].
were responsible for the enhanced selectivity of products [130].
Although the detailed mechanism for how the heteroatom doping

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 15

affects the catlaytic properties remians unclear, the above cases as the catalysts, 83.0% and 48.5% conversion were obtained with
have certainly benefited the understanding of the structure-related 54.4% and 67.2% selectivity for the Sonogashira cross-coupling
catlaytic property of bimetallic gold nanoclusters at an unprece- product, respectively. However, Cux Au25−x (SR)18 preferred the
dented level. Ullmann homo-coupling reaction of p-iodoanisole, and so 52.4%
conversion of p-iodoanisole and only 28.3% selectivilty for the Sono-
Crotyl alcohol oxidation. Bimetallic AuPd nanoclusters supported gashira cross-coupling product were attained. Characterizations
on alumina were obtained from Au25 (SR)18 nanoclusters through and DFT calculations showed that the catalytic activity was largely
activation with mild thermal treatment or LiBH4 treatment fol- determined by the electronic effects in the bimetallic nanoclusters’
lowed by selective deposition of Pd via ascorbic acid reduction 13-atom core, and that the selectivity was dominated by the type of
[132]. Characterizations showed that AuPd nanoclusters derived atoms on the Mx Au12−x shell (M = Ag, Cu and Au) in the bimetallic
from LiBH4 -activation had a small avarage particle size of 2.5 nm nanocluster catalysts [135].
with Au core-Pd shell structure. The bimetallic AuPd nanoclusters Recently, a novel Au26 Cd5 nanocluster has been prepared using
exhibited higher catalytic activity and selectivity for crotonalde- the thiolate-protected Au25 nanocluster as a precursor by peeling
hyde than the homometallic Pd catalyst in the liquid oxidation and doping [136]. The Au26 Cd5 nanocluster catalyst was found to
of crotyl alcohol, the turnover number (TON) of the former being exhibit high catalytic activity and good reusability in the three-
about 10–20 times higher than that of the latter, while pure Au cat- component coupling reaction of alkynes, aldehydes and amines.
alysts were catalytically inert under the investigated conditions. In contrast, homogold Au25 clusters showed no catalytic activity
The promoted catalytic activity of AuPd nanoclusters was ascribed under the conditions. The high catalytic activity was related to the
to the strong electronic interaction between Au and Pd and their cooperation between doping cadmium atoms and the neighboring
small particles size [132]. It provided a novel and efficient method gold atoms in bimetallic nanoclusters [136]. These works provide
to design gold-based bimetallic nanoclusters using ligand-on gold insight into the fundamental relationship between the composi-
nanoclusters as a precursor. tions of bimetallic nanoclusters and their catalytic properties, and
will stimulate more future research on the fine tuning of nan-
Organosilane oxdation. Recently, Wang et al. obtained alkynyl- oclusters’ compositions and structures to improve their catalytic
protected intermetallic nanocluster Au34 Ag28 (PhC C)34 by co- performance.
reduction of gold and silver salts with tert-butylamine borane
complex [133]. The clusters displayed a four-concentric-shell
Ag@Au17 @Ag27 @Au17 structure, and the exterior 17 gold atoms Future perspective
were linearly coordinated with phenylalkynyl ligands to form
staple-like “PhC C−Au−C CPh” motifs. Ligand-on AuAg nanoclus- Nanoclusters with a precise number of atoms display unique
ters supported on carbon were found to exhibit far higher activity and even unexpected properties. Herein, we have illustrated the
than those with ligand-off for the hydrolytic oxidation of triethylsi- case of gold nanoclusters as the catalyst. The synthetic method-
lane (Fig. 18) [133]. It showed the important role of surface ligands ologies [26,27] could be extended to the cases of Pd, Pt and other
in promoting the catalytic reaction. Future research should focus on metals. Undoubtedly, the controlled synthesis of metal nanoclus-
the fundamental understanding of the intriguing promotion effect. ters with atomic precision will offer new exciting opportunities for
achieving fundamental insight into nanocatalysis and correlation
Selective hydrogenation of the catalytic properties with their atomic level structure.
Zhu et al. [134] found that the Pd1 Au24 (SR)18 /Fe2 O3 Due to ultra-small size, the properties of nanoclusters are very
(R = CH2 CH2 Ph) catalyst exhibited similar catalytic activity sensitive to the number of atoms in the clusters. Significant changes
compared to that of Au25 (SR)18 /Fe2 O3 in selective hydrogenation in the properties may occur when even a single atom is added or
of benzalacetone (42% conversion vs 40%). It indicated that the removed [137,138], and such catalysts are particularly useful in
central palladium atom exerted no significant effect on the cat- revealing the dependence on the structure and electronic proper-
alytic hydrogenation activity of homogold nanoclusters, while it ties of nanoclusters [139]. Future work should further explore such
contributed to their remarkably increase in the catalytic oxidation unique properties.
activity [126]. Bimetallic nanoclusters hold great potential in catalysis. Distinct
Besides, the nature of the ligand can significantly influence improvements in catalytic properties of bimetallic gold nanoclus-
the catalytic hydrogenation performance of bimetallic gold nan- ters compared to homogold ones were observed when another
oclusters. For instance, Yang et al. [128] found that the presence element is introduced [125–132,135,136]. It is an efficient method
of the small size 2-pyridylthiolate on the surface of Au13 Cu8 for fine tuning of the catalytic property of Au nanoclusters. As
nanoclusters still exposed the active sites of Au atoms to small respect to Ag or Cu doped gold nanoclusters, pure Agx Aum–x (SR)m
molecules, and then the hydrogenation of 4-nitrophenol to 4- or Cux Aun–x (SR)m clusters with a specific value of x should be sep-
aminophenol by NaBH4 efficiently took place in the presence of arated and obtained, and then the precise correlation the doping
Au13 Cu8 nanoclusters. In comparison, the larger size ligands such amount and sites in heteroatom doped nanoclusters with the cat-
as triphenylphosphine and 4-tert-butylbenzenethiol on the sur- alytic property should be explored in future work.
face of Au13 Cu2 and Au13 Cu4 nanoclusters completely blocked the Ligand not only retains the structural integrity of gold
Au active sites, so the above hydrogenation did not proceed over nanoclusters but also imposes the unnegligible effect on the cat-
Au13 Cu2 or Au13 Cu4 nanoclusters. alytic properties [65–72,93,96–98,106,109,114,116,133,140,141].
Ligand-protected gold nanoclusters have been found to be cat-
Carbon–Carbon coupling reaction alytically active for many reactions, and they even exhibit unique
Li et al. [135] found that gold-based bimetallic Mx Au25−x (SR)18 property in some cases [95–97,109,114,116,133], but the detailed
(M = Pt, Ag and Cu; R = CH2 CH2 Ph) nanoclusters (supported on role of ligand remains unclear in many reactions and should be
TiO2 ) displayed distinct activity and selectivity for the Sonogashira studied carefully in future research. Based on the reported results,
cross-coupling product in the carbon–carbon coupling reaction of ligand engineering is an efficient approach to tailoring the catalytic
p-iodoanisole and phenylacetylene. Au25 (SR)18 gave 79.5% conver- property of gold nanoclusters.
sion of p-iodoanisole and 65.7% selectivilty for the Sonogashira Besides, ligand is usually considered to be an adverse factor
cross-coupling product. Using Agx Au25−x (SR)18 and Pt1 Au24 (SR)18 in catalysis as it prevents reactants from accessing the catalytic

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
16 J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx

sites. So the removal of ligand is often required to obtain high per- [26] R. Jin, H. Qian, Z. Wu, Y. Zhu, M. Zhu, A. Mohanty, N. Garg, J. Phys. Chem. Lett.
formance on gold nanocluster catalysts. Several methods for the 1 (2010) 2903–2910.
[27] C. Zeng, Y. Chen, A. Das, R. Jin, J. Phys. Chem. Lett. 6 (2015) 2976–2986.
removal of ligand have been developed, and the size of the as- [28] H. Qian, M. Zhu, U.N. Andersen, R. Jin, J. Phys. Chem. A 113 (2009)
prepared gold catalysts was reported to be preserved in many cases 4281–4284.
[65,79–82,142]. Of note, TEM can characterize whether the size of [29] H. Qian, Y. Zhu, R. Jin, A.C.S. Nano, (2009) 3795–3803.
[30] C. Zeng, Y. Chen, G. Li, R. Jin, Chem. Mater. 26 (2014) 2635–2641.
gold nanocluster is preserved, but the precise composition (such [31] C. Liu, J. Lin, Y. Shi, G. Li, Nanoscale 7 (2015) 5987–5990.
as the number of gold atoms) of nanoclusters was not identified [32] H. Qian, R. Jin, Nano Lett. 9 (2009) 4083–4087.
[81,82,143]. Determining the gold atom number of supported gold [33] H. Qian, Y. Zhu, R. Jin, Proc. Natl. Acad. Sci. 109 (2012) 696–700.
[34] S. Kumar, R. Jin, Nanoscale 4 (2012) 4222–4227.
nanoclusters is still a challenge because of the limitations of the
[35] L.M. Tvedte, C.J. Ackerson, J. Phys. Chem. A 118 (2014) 8124–8128.
current characterization tools. Thus, characterization of the precise [36] C. Zeng, H. Qian, T. Li, G. Li, N.L. Rosi, B. Yoon, R.N. Barnett, R.L. Whetten, U.
size of supported nanoclusters should be explored in future study. Landman, R. Jin, Angew. Chem. Int. Ed. 51 (2012) 13114–13118.
[37] C. Zeng, C. Liu, Y. Pei, R. Jin, ACS Nano 7 (2013) 6138–6145.
Although ligand-on or ligand-off gold nanoclusters have been
[38] C. Zeng, T. Li, A. Das, N.L. Rosi, R. Jin, J. Am. Chem. Soc. 135 (2013)
found to exhibit high catalytic properties in many chemical reac- 10011–10013.
tions, detailed catalytic mechanisms have not been revealed yet [39] C. Zeng, C. Liu, Y. Chen, N.L. Rosi, R. Jin, J. Am. Chem. Soc. 136 (2014)
in many cases. Identifying the catalytic sites and exploring the 11922–11925.
[40] C. Zeng, Y. Chen, K. Kirschbaum, K. Appavoo, M.Y. Sfeir, R. Jin, Sci. Adv. 1
reaction mechanisms should be pursued in future research. It is (2015) e1500045.
noted that the structure of gold nanoclusters may be dynamic in [41] M. Zhu, C.M. Aikens, F.J. Hollander, G.C. Schatz, R. Jin, J. Am. Chem. Soc. 130
some harsh reactions (e.g. under high temperatures); thus, moni- (2008) 5883–5885.
[42] M.W. Heaven, A. Dass, P.S. White, K.M. Holt, R.W. Murray, J. Am. Chem. Soc.
toring the structure under the reaction conditions is important for 130 (2008) 3754–3755.
understanding the mechanism. More efforts should be focused on [43] H. Qian, M. Zhu, Z. Wu, R. Jin, Acc. Chem. Res. 45 (2012) 1470–1479.
utilizing in-situ or operando techniques, such as DRIFT-IR, XANES [44] Z. Wu, C. Gayathri, R. Gil, R. Jin, J. Am. Chem. Soc. 131 (2009) 6535–6542.
[45] Y. Shichibu, Y. Negishi, T. Tsukuda, T. Teranishi, J. Am. Chem. Soc. 127 (2005)
and EXAFS [70,144,145], to determine how the characteristics 13464–13465.
of metal nanoclusters change under working conditions. Besides, [46] Z. Wu, J. Suhan, R. Jin, J. Mater. Chem. 19 (2009) 622–626.
applying gold nanoclusters in other important reactions such as [47] Y. Negishi, W. Kurashige, Y. Niihori, T. Iwasa, K. Nobusada, Phys. Chem.
Chem. Phys. 12 (2010) 6219–6225.
propylene epoxidation should also be considered in future study.
[48] H. Yang, Y. Wang, J. Yan, X. Chen, X. Zhang, H. Häkkinen, N. Zheng, J. Am.
Chem. Soc. 136 (2014) 7197–7200.
Acknowledgements [49] J.L. Zeng, Z.J. Guan, Y. Du, Z.A. Nan, Y.M. Lin, Q.M. Wang, J. Am. Chem. Soc.
138 (2016) 7848–7851.
[50] X. Kang, S. Wang, Y. Song, S. Jin, G. Sun, H. Yu, M. Zhu, Angew. Chem. Int. Ed.
J.Z. acknowledges the financial support from the National Nat- 55 (2016) 3611–3614.
ural Science Foundation of China (21576248) and a research fund [51] S. Matsuo, S. Takano, S. Yamazoe, K. Koyasu, T. Tsukuda, ChemElectroChem 3
(2016) 1206–1211.
from the doctoral program of Zhengzhou University of Light Indus- [52] X.K. Wan, X.L. Cheng, Q. Tang, Y.Z. Han, G. Hu, D.E. Jiang, Q.M. Wang, J. Am.
try (2014BSJJ007). R.J. acknowledges the funding support from the Chem. Soc. 139 (2017) 9451–9454.
U.S. National Science Foundation (DMREF-0903225). [53] X. Kang, J. Xiang, Y. Lv, W. Du, H. Yu, S. Wang, M. Zhu, Chem. Mater. 29
(2017) 6856–6862.
[54] Q. Li, S. Wang, K. Kirschbaum, K.J. Lambright, A. Dasa, R. Jin, Chem. Commun.
References 52 (2016) 5194–5197.
[55] S. Wang, Y. Song, S. Jin, X. Liu, J. Zhang, Y. Pei, X. Meng, M. Chen, P. Li, M. Zhu,
[1] H. Ertl, F. Knözinger, J. Schüth (Eds.), Handbook of Heterogeneous Catalysis, J. Am. Chem. Soc. 137 (2015) 4018–4021.
2nd ed., Wiley-VCH, Weinheim, Germany, 2008. [56] R. Kazan, B. Zhang, T. Bürgi, Dalton Trans. 46 (2017) 7708–7713.
[2] Y. Lei, F. Mehmood, S. Lee, J. Greeley, B. Lee, S. Seifert, R.E. Winans, J.W. Elam, [57] M.S. Bootharaju, C.P. Joshi, M.R. Parida, O.F. Mohammed, O.M. Bakr, Angew.
R.J. Meyer, P.C. Redfern, D. Teschner, R. Schlögl, M.J. Pellin, L.A. Curtiss, S. Chem. Int. Ed. 55 (2016) 922–926.
Vajda, Science 328 (2010) 224–228. [58] S. Wang, H. Abroshan, C. Liu, T.Y. Luo, M. Zhu, H.J. Kim, N.L. Rosi, R. Jin, Nat.
[3] G.C. Bond, D.T. Thompson, Cat. Rev.-Sci. Eng. 41 (1999) 319–388. Commun. (2017), http://dx.doi.org/10.1038/s41467-017-00939-0.
[4] H. Lee, S.E. Habas, S. Kweskin, D. Butcher, G.A. Somorjai, P.D. Yang, Angew. [59] K.R. Krishnadas, A. Ghosh, A. Baksi, I. Chakraborty, G. Natarajan, T. Pradeep, J.
Chem. Int. Ed. 45 (2006) 7824–7828. Am. Chem. Soc. 138 (2016) 140–148.
[5] I. Lee, F. Delbecq, R. Morales, M.A. Albiter, F. Zaera, Nat. Mater. 8 (2009) [60] H.J. Freund, G. Meijer, M. Scheffler, R. Schlögl, M. Wolf, Angew. Chem. Int. Ed.
132–138. 50 (2011) 10064–10094.
[6] L.L. Chng, N. Erathodiyil, J.Y. Yin, Acc. Chem. Res. 46 (2013) 1825–1837. [61] M. Valden, X. Lai, D.W. Goodman, Science 281 (1998) 1647–1650.
[7] Q. Feng, S. Zhao, Y. Wang, J. Dong, W. Chen, D. D.H, J. Wang, Y. Yang, H. Zhu, [62] A.A. Herzing, C.J. Kiely, A.F. Carley, P. Landon, G.J. Hutching, Science 321
L. Zhu, Z. Gu, Y. Li, R. Liu, J. Yu, Y. Li, J. Am. Chem. Soc. 139 (2017) 7294–7301. (2008) 1331–1335.
[8] M. Haruta, T. Kobayashi, H. Sano, N. Yamada, Chem. Lett. 16 (1987) 405–408. [63] A. Sanchez, S. Abbet, U. Heiz, W.D. Schneider, H. Häkkinen, R.N. Barnett, U.
[9] G.J. Hutchings, J. Catal. 96 (1985) 292–295. Landman, J. Phys. Chem. A 103 (1999) 9573–9578.
[10] A. Corma, H. Garcia, Chem. Soc. Rev. 37 (2008) 2096–2126. [64] B. Yoon, H. Häkkinen, U. Landman, A.S. Wörz, J.M. Antonietti, S. Abbet, K.
[11] Y. Zhang, X. Cui, F. Shi, Y. Deng, Chem. Rev. 112 (2012) 2467–2505. Judai, U. Heiz, Science 307 (2005) 403–407.
[12] J.B. Zhao, G. Yu, K. Xin, L. Li, T. Fu, Y. Cui, H. Liu, N. Xue, L. Peng, W. Ding, [65] X. Nie, H. Qian, Q. Ge, H. Xu, R. Jin, ACS Nano 6 (2012) 6014–6022.
Appl. Catal. A-Gen. 482 (2014) 294–299. [66] Z. Wu, G. Hu, D. Jiang, D.R. Mullins, Q.F. Zhang, L.F. Allard, L.S. Wang, S.H.
[13] R.Y. Zhong, K.Q. Sun, Y.C. Hong, B.Q. Xu, ACS Catal. 4 (2014) 3982–3993. Overbury, Nano Lett. 6 (2016) 6560–6567.
[14] A. Villa, N. Dimitratos, C.E. Chan-Thaw, C. Hammond, L. Prati, G.J. Hutchings, [67] X. Nie, C. Zeng, X. Ma, H. Qian, Q. Ge, H. Xu, R. Jin, Nanoscale 5 (2013)
Acc. Chem. Res. 48 (2015) 1403–1412. 5912–5918.
[15] G. Li, R. Jin, Acc. Chem. Res. 46 (2013) 1749–1758. [68] S. Gaur, J.T. Miller, D. Stellwagen, A. Sanampudi, C.S.S.R. Kumard, J.J. Spivey,
[16] S. Yamazoe, K. Koyasu, T. Tsukuda, Acc. Chem. Res. 47 (2013) 816–824. Phys. Chem. Chem. Phys. 14 (2012) 1627–1634.
[17] C. Zeng, Y. Chen, K. Kirschbaum, K.J. Lambright, R. Jin, Science 354 (2016) [69] W. Li, Q. Ge, X. Ma, Y. Chen, M. Zhu, H. Xu, R. Jin, Nanoscale 8 (2016)
1580–1584. 2378–2385.
[18] R. Jin, C. Zeng, M. Zhou, Y. Chen, Chem. Rev. 116 (2016) 10346–10413. [70] Z.L. Wu, D.E. Jiang, A.K.P. Mann, D.R. Mullins, Z.A. Qiao, L.F. Allard, C.J. Zeng,
[19] E.C. Tyo, S. Vajda, Nat. Nanotechnol. 10 (2015) 577–588. R.C. Jin, S.H. Overbury, J. Am. Chem. Soc. 136 (2014) 6111–6122.
[20] R. Jin, K. Nobusada, Nano Res. 7 (2014) 285–300. [71] H. Tsunoyama, N. Ichikuni, H. Sakurai, T. Tsukuda, J. Am. Chem. Soc. 131
[21] C.D. Bain, E.B. Troughton, Y.T. Tao, J. Evall, G.M. Whitesides, R.G. Nuzzo, J. (2009) 7086–7093.
Am. Chem. Soc. 111 (1989) 321–335. [72] T. Yoskamtorn, S. Yamazoe, R. Takahata, J.-I. Nishigaki, A. Thivasasith, J.
[22] C.A. Mirkin, R.L. Letsinger, R.C. Mucic, J.J. Storhoff, Nature 382 (1996) Limtrakul, T. Tsukuda, ACS Catal. 4 (2014) 3696–3700.
607–609. [73] L. Li, L. Dou, H. Zhang, Nanoscale 6 (2014) 3753–3763.
[23] M. Brust, M. Walker, D. Bethell, D.J. Schiffrin, R. Whyman, J. Chem. Soc. [74] S. Wang, S. Yin, G. Chen, L. Li, H. Zhang, Catal. Sci. Technol. 6 (2016)
Chem. 0 (1994) 801–802. 4090–4104.
[24] Y. Negishi, K. Nobusada, T. Tsukuda, J. Am. Chem. Soc. 127 (2005) [75] R.H. Adnan, G.G. Andersson, M.I.J. Polson, G.F. Methad, V.B. Golovko, Catal.
5261–5270. Sci. Technol. 5 (2015) 1323–1333.
[25] M. Zhu, E. Lanni, N. Garg, M.E. Bier, R. Jin, J. Am. Chem. Soc. 130 (2008) [76] M. Turner, V.B. Golovko, O.P.H. Vaughan, P. Abdulkin, A. Berenguer-Murcia,
1138–1139. M.S. Tikhov, B.F.G. Johnson, R.M. Lambert, Nature 454 (2008) 981–983.

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009
G Model
NANTOD-639; No. of Pages 17 ARTICLE IN PRESS
J. Zhao, R. Jin / Nano Today xxx (2018) xxx–xxx 17

[77] Y. Zhu, H. Qian, R. Jin, Chem. Eur. J. 16 (2010) 11455–11462. [112] G. Li, D. Jiang, C. Liu, C. Yu, R. Jin, J. Catal. 306 (2013) 177–183.
[78] Y. Zhu, H. Qian, M. Zhu, R. Jin, Adv. Mater. 22 (2010) 1915–1920. [113] H. Abroshan, G. Li, J. Lin, H.J. Kim, R. Jin, J. Catal. 337 (2016) 72–79.
[79] Y. Liu, H. Tsunoyama, T. Akita, T. Tsukuda, Chem. Commun. 46 (2010) [114] Q. Li, A. Das, S. Wang, Y. Chen, R. Jin, Chem. Commun. 52 (2016)
550–552. 14298–14301.
[80] S. Das, A. Goswami, M. Hesari, J.F. Al-Sharab, E. Mikmeková, F. Maran, T. [115] B. Vilhanová, J. Václavík, L. Artigli, M. Ranocchiari, A. Togni, J.A. van
Asefa, Small 10 (2014) 1473–1478. Bokhoven, ACS Catal. 7 (2017) 3414–3418.
[81] B. Zhang, J. Fang, J. Li, J.J. Lau, D. Mattia, Z. Zhong, J. Xie, N. Yan, Chem. Asian J. [116] G. Li, H. Abroshan, C. Liu, S. Zhuo, Z. Li, Y. Xie, H.J. Kim, N.L. Rosi, R. Jin, ACS
11 (2016) 532–539. Nano 10 (2016) 7998–8005.
[82] Y. Liu, H. Tsunoyama, T. Akita, S. Xie, T. Tsukuda, ACS Catal. 1 (2011) 2–6. [117] R. Jin, Nanoscale 7 (2015) 1549–1565.
[83] B. Zhang, S. Kaziz, H. Li, M.G. Hevia, D. Wodka, C. Mazet, T. Bürgi, N. Barrabës, [118] C. Yu, G. Li, S. Kumar, H. Kawasaki, R. Jin, J. Phys. Chem. Lett. 4 (2013)
J. Phys. Chem. C 119 (2015) 11193–11199. 2847–2852.
[84] P.A. Bond, D.A. Sermon, G. Buchanan, P.B. Webb, J. Chem. Soc. Chem. [119] N. Sakai, T. Tatsuma, Adv. Mater. 22 (2010) 3185–3188.
Commun. 13 (1973) 444–445. [120] A. Kogo, N. Sakai, T. Tatsuma, Electrochem. Commun. 12 (2010) 996–999.
[85] P. Claus, A. Brueckner, C. Mohr, H. Hofmeister, J. Am. Chem. Soc. 22 (2000) [121] H. Chen, C. Liu, M. Wang, C. Zhang, N. Luo, Y. Wang, H. Abroshan, G. Li, F.
11430–11439. Wang, ACS Catal. 7 (2017) 3632–3638.
[86] A. Corma, P. Serna, Science 313 (2006) 332–334. [122] Z. Li, C. Liu, H. Abroshan, D.R. Kauffman, G. Li, ACS Catal. 7 (2017) 3368–3374.
[87] T. Mitsudomea, K. Kaneda, Green Chem. 15 (2013) 2636–2654. [123] Y.S. Chen, P.V. Kamat, J. Am. Chem. Soc. 136 (2014) 6075–6082.
[88] S. Nishimura, Handbook of Heterogeneous Catalytic Hydrogenation for [124] Y. Negishi, M. Mizuno, M. Hirayama, M. Omatoi, T. Takayama, A. Iwase, A.
Organic Synthesis, Wiley-VCH, New York, 2001. Kudo, Nanoscale 5 (2013) 7188–7192.
[89] L. Liu, B. Qiao, Y. Ma, J. Zhang, Y. Deng, Dalton Trans. 0 (2008) 2542–2548. [125] W. Li, C. Liu, H. Abroshan, Q. Ge, X. Yang, H. Xu, G. Li, J. Phys. Chem. C 120
[90] Y. Zhu, Z. Wu, C. Gayathri, H. Qian, R.R. Gil, R. Jin, J. Catal. 271 (2010) (2016) 10261–10267.
155–160. [126] S. Xie, H. Tsunoyama, W. Kurashige, Y. Negishi, T. Tsukuda, ACS Catal. 2
[91] Y. Zhu, H. Qian, B.A. Drake, R. Jin, Angew. Chem. Int. Ed. 49 (2010) (2012) 1519–1523.
1295–1298. [127] H. Deng, S. Wang, S. Jin, S. Yang, Y. Xu, L. Liu, J. Xiang, D. Hu, M. Zhu, Gold
[92] Y. Zhu, H. Qian, R. Jin, J. Mater. Chem. 21 (2011) 6793–6799. Bull 48 (2015) 161–167.
[93] I. Cano, A.M. Chapman, A. Urakawa, Piet W.N.M. van Leeuwen, J. Am. Chem. [128] H. Yang, Y. Wang, J. Lei, L. Shi, X. Wu, V. Makinen, S. Lin, Z. Tang, J. He, H.
Soc. 136 (2014) 2520–2528. Häkkinen, L. Zheng, N. Zheng, J. Am. Chem. Soc. 135 (2013) 9568–9571.
[94] R. Ouyang, D. Jiang, ACS Catal. 5 (2015) 6624–6629. [129] H. Qian, D. Jiang, G. Li, C. Gayathri, A. Das, R.R. Gil, R. Jin, J. Am. Chem. Soc.
[95] G. Li, D. Jiang, S. Kumar, Y. Chen, R. Jin, ACS Catal. 4 (2014) 2463–2469. 134 (2012) 16159–16162.
[96] G. Li, C. Zeng, R. Jin, J. Am. Chem. Soc. 136 (2014) 3673–3679. [130] S. Wang, S. Jin, S. Yang, S. Chen, Y. Song, J. Zhang, M. Zhu, Sci. Adv. 1 (2015)
[97] G. Li, H. Abroshan, Y. Chen, R. Jin, H.J. Kim, J. Am. Chem. Soc. 137 (2015) e1500441.
14295–14304. [131] G. Li, R. Jin, Catal. Today 278 (2016) 187–191.
[98] M. Yan, T. Jin, Y. Ishikawa, T. Minato, T. Fujita, L. Chen, M. Bao, N. Asao, M. [132] K.E. Lee, A. Shivhare, Y. Hu, R.W.J. Scott, Catal. Today 280 (2017) 259–265.
Chen, Y. Yamamoto, J. Am. Chem. Soc. 134 (2012) 17536–17542. [133] Y. Wang, X.K. Wan, L. Ren, H. Su, G. Li, S. Malola, S. Lin, Z. Tang, H. Hakkinen,
[99] C. Liu, H. Abroshan, C. Yan, G. Li, M. Haruta, ACS Catal. 6 (2016) 92–99. B.K. Teo, Q.M. Wang, N. Zheng, J. Am. Chem. Soc. 138 (2016) 3278–3281.
[100] H. Blaser, H. Steiner, M. Studer, ChemCatChem 1 (2009) 210–221. [134] E. Qian, Y. Barry, R. Zhu, Acta Phy. −Chem. Sin. 27 (2011) 513–519.
[101] H. Yamamoto, H. Yano, H. Kouchi, Y. Obora, R. Arakawa, H. Kawasaki, [135] Z. Li, X. Yang, C. Liu, J. Wang, G. Li, Prog. Nat. Sci-Mater. 26 (2016) 477–482.
Nanoscale 4 (2012) 4148–4154. [136] M. Li, S.k. Tian, Z. Wu, Chin. J. Chem. 35 (2017) 567–571.
[102] A. Shivhare, S.J. Ambrose, H. Zhang, R.W. Purves, R.W.J. Scott, Chem. [137] M. Zhu, H. Qian, R. Jin, J. Phys. Chem. Lett. 1 (2010) 1003–1007.
Commun. 49 (2013) 276–278. [138] Q. Li, T.Y. Luo, M.G. Taylor, S. Wang, X. Zhu, Y. Song, G. Mpourmpakis, N.L.
[103] J. Li, R.R. Nasaruddin, Y. Feng, J. Yang, N. Yan, J. Xie, Chem. Eur. J. 22 (2016) Rosi, R. Jin, Sci. Adv. 3 (2017) e1603193.
14816–14820. [139] S. Zhao, A. Das, H. Zhang, R.X. Jin, Y.B. Song, R. Jin, Prog. Nat. Sci.-Mater. 25
[104] R.R. Nasaruddin, T. Chen, J. Li, N. Goswami, J. Zhang, N. Yan, J. Xie, (2016) 483–486.
ChemCatChem (2017), http://dx.doi.org/10.1002/cctc.201701472. [140] P. Liu, R. Qin, G. Fu, N. Zheng, J. Am. Chem. Soc. 139 (2017) 2122–2131.
[105] J. Fang, J. Li, B. Zhang, X. Yuan, H. Asakura, T. Tanaka, K. Teramura, J. Xie, N. [141] C. Lavenn, A. Demessence, A. Tuel, Catal. Today 235 (2014) 72–78.
Yan, Nanoscale 7 (2015) 6325–6333. [142] J. Fang, B. Zhang, Q. Yao, Y. Yang, J. Xie, N. Yan, Coord. Chem. Rev. 322 (2016)
[106] Y. Tan, X.Y. Liu, L. Zhang, A. Wang, L. Li, X. Pan, S. Miao, M. Haruta, H. Wei, H. 1–29.
Wang, F. Wang, X. Wang, T. Zhang, Angew. Chem. Int. Ed. 56 (2017) [143] G. Ma, A. Binder, M. Chi, C. Liu, R. Jin, D. Jiang, J. Fan, S. Dai, Chem. Commun.
2709–2713. 48 (2012) 11413–11415.
[107] C. Oger, L. Balas, T. Durand, J.M. Galano, Chem. Rev. 13 (2013) 1313–1350. [144] E. Bayram, J.C. Linehan, J.L. Fulton, J.A.S. Roberts, N.K. Szymczak, T.D.
[108] S.S. Li, X. Liu, Y.M. Liu, H.Y. He, K.N. Fan, Y. Cao, Chem. Commun. 50 (2014) Smurthwaite, S. Özkar, M. Balasubramanian, R.G. Finke, J. Am. Chem. Soc.
5626–5628. 133 (2011) 18889–18902.
[109] G. Li, R. Jin, J. Am. Chem. Soc. 136 (2014) 11347–11354. [145] G. Malta, S.A. Kondrat, S.J. Freakley, C.J. Davies, L. Lu, S. Dawson, A. Thetford,
[110] G. Li, R. Jin, Nanotechnol. Rev. 2 (2013) 529–545. E.K. Gibson, D.J. Morgan, W. Jones, P.P. Wells, P. Johnston, C. Richard, A.
[111] G. Li, C. Liu, Y. Lei, R. Jin, Chem. Commun. 48 (2012) 12005–12007. Catlow, C.J. Kiely, G.J. Hutchings, Science 355 (2017) 1399–1403.

Please cite this article in press as: J. Zhao, R. Jin, Heterogeneous catalysis by gold and gold-based bimetal nanoclusters, Nano Today
(2018), https://doi.org/10.1016/j.nantod.2017.12.009

You might also like