au25SR - Captainc8nr02973c

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 77

Nanoscale

View Article Online


REVIEW View Journal | View Issue

Au25(SR)18: the captain of the great nanocluster


Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Cite this: Nanoscale, 2018, 10, 10758


ship
Xi Kang, Hanbao Chong and Manzhou Zhu *

Noble metal nanoclusters are in the intermediate state between discrete atoms and plasmonic nano-
particles and are of significance due to their atomically accurate structures, intriguing properties, and
great potential for applications in various fields. In addition, the size-dependent properties of nanoclusters
construct a platform for thoroughly researching the structure (composition)-property correlations, which
is favorable for obtaining novel nanomaterials with enhanced physicochemical properties. Thus far, more
than 100 species of nanoclusters (mono-metallic Au or Ag nanoclusters, and bi- or tri-metallic alloy
nanoclusters) with crystal structures have been reported. Among these nanoclusters, Au25(SR)18—the
brightest molecular star in the nanocluster field—is capable of revealing the past developments and pro-
specting the future of the nanoclusters. Since being successfully synthesized (in 1998, with a 20-year
history) and structurally determined (in 2008, with a 10-year history), Au25(SR)18 has stimulated the interest
of chemists as well as material scientists, due to the early discovery, easy preparation, high stability, and
easy functionalization and application of this molecular star. In this review, the preparation methods,
crystal structures, physicochemical properties, and practical applications of Au25(SR)18 are summarized.
The properties of Au25(SR)18 range from optics and chirality to magnetism and electrochemistry, and the
property-oriented applications include catalysis, chemical imaging, sensing, biological labeling, biomedi-
cine and beyond. Furthermore, the research progress on the Ag-based M25(SR)18 counterpart (i.e.,
Ag25(SR)18) is included in this review due to its homologous composition, construction and optical
absorption to its gold-counterpart Au25(SR)18. Moreover, the alloying methods, metal-exchange sites and
property alternations based on the templated Au25(SR)18 are highlighted. Finally, some perspectives and
Received 12th April 2018, challenges for the future research of the Au25(SR)18 nanocluster are proposed (also holding true for all
Accepted 21st May 2018
members in the nanocluster field). This review is directed toward the broader scientific community inter-
DOI: 10.1039/c8nr02973c ested in the metal nanocluster field, and hopefully opens up new horizons for scientists studying nano-
rsc.li/nanoscale materials. This review is based on the publications available up to March 2018.

1 Introduction morphology (or composition) correlated physicochemical pro-


perties have also been extensively researched for investigating
The research on metal nanoparticles has always been a hot more enhanced properties of these nanomaterials.4,5,7,8,11,13–16
issue in the academic frontier, due to their attractive catalytic, However, the atomically accurate structure–property relation-
optical, magnetic, and electrochemical properties.1–12 The ship of nanoparticles has not been thoroughly grasped, mainly
because of two general issues: the polydisperse size as well as
the uncertain surface structure of nanoparticles.18,19 An
Department of Chemistry and Center for Atomic Engineering of Advanced Materials,
atomic-level understanding of such relationships requires
Institute of Physical Science and Information Technology and AnHui Province Key
Laboratory of Chemistry for Inorganic/Organic Hybrid Functionalized Materials,
precise molecular entities such as model nanosystems, and
Anhui University, Hefei, Anhui 230601, P. R. China. E-mail: zmz@ahu.edu.cn precise molecular tools.18–24 On this basis, metal nanoclusters

Xi Kang earned his M.S. in Chemistry from Anhui University and Hanbao Chong is a lecturer at the Institute of Physical Science
is currently a Ph.D. candidate under the supervision of Prof. and Information Technology at Anhui University. He obtained his
Manzhou Zhu. His research interests include the structure and B.S. in Environmental Science (2011) and M.S. in Organic
property evolution of metal nanoclusters. Chemistry (2014) from Anhui University. His research interest is
the catalytic application of gold nanoparticles.

10758 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

with ultra-small sizes (the metallic core always <2 nm in dia-


meter) provide an exciting opportunity to investigate the struc-
ture–property correlations at the atomic level owing to the
monodisperse size and accurately characterized structures of
nanoclusters.20–31 Furthermore, nanoclusters fill the gap
between discrete atoms and plasmonic nanoparticles, and the
quantum size effect as well as the discrete electronic states of
nanoclusters pose more unprecedented opportunities for
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

guiding the preparation of multi-functional nanomaterials


with wider applications.18–21,23,25,30–41
Since the first atomically precise Au nanocluster protected by
thiolates (i.e., Au102(SR)44) was reported, the development of Au
nanoclusters appears to have a rapidly increasing tendency.42
To date, more than 40 species of mono-thiolated Au nano-
clusters with atomically precise structure have been reported,
such as Au18, Au20, Au21, Au23, Au24, Au25, Au28, Au30, Au34,
Au36, Au38, Au40, Au42, Au43, Au44, Au49, Au52, Au60, Au92, Au102,
Au103, Au130, Au133, Au146, Au246, Au279 and so on.42–83 In
addition, several monodisperse Au nanoclusters also have been
synthesized, but with no crystal structures.84–90 Among these Scheme 1 Summary of the Au nanoclusters ( protected by thiolates)
nanoclusters, Au25(SR)18 and its several derivatives hold a very with atomically precise structures (based on the publications available
special place in the nanocluster field (Scheme 1) because of the up to February 2018, redrawn from ref. 42–83). Note that due to space,
not all thiolated Au nanoclusters are shown in this scheme. Colour
early discovery, easy preparation, high stability, and thus can
codes: orange spheres, Au; red spheres, S; the carbon and hydrogen
easy functionalization and application.51,52,91–93 The construc- atoms are not shown for clarity.
tion of Au25(SR)18 exhibits an icosahedral Au13 kernel, which is
further wrapped by six Au2(SR)3 staple motifs (Scheme 2).51,52
Through many years of efforts and developments, the physico-
chemical properties of Au25(SR)18 has been fully researched,
including catalytic, optical, electrochemical, chiral, and mag-
netic properties.91–96 In addition, as depicted in Scheme 2,
Au25(SR)18 and its several functionalized derivatives have stimu-
lated great interest not only in theoretical research, but also in
practical applications.25–27,30 For instance, the available crystal
structure of the Au25(SR)18 permits the correlation of its catalytic
activity with the atomically precise structure.25,30,97
Furthermore, benefiting from the high biocompatibility, good
photostability, and low toxicity, Au25 nanoclusters are highly
promising for application in the biological field, namely, cell
labeling, phototherapy, and biosensing.23,27,34,41,98–102
Previously, some overviews have touched upon the
syntheses,31,34,36,40,103–105 crystal structures,19,26,28,29,38,106–109
properties,23–27,33,110,111 and applications of metal
nanoclusters.30–32,37,41,99,104,111–115 Jin et al. summarized the
principles that permit atomically precise syntheses, new types
of crystal structures, and unique physical and chemical pro-
Scheme 2 Syntheses, properties, and applications of the Au25(SR)18
nanocluster.

Manzhou Zhu is Professor of Chemistry at Anhui University. He


received his Ph.D. in Chemistry from the University of Science and perties of atomically accurate nanoclusters (mainly gold nano-
Technology of China (USTC, Hefei, China) in 2000. He then con- clusters).20 Pradeep and co-workers summarized the measur-
ducted postdoctoral research at USTC and Carnegie Mellon ing methods, development history, and main applications
University (CMU, Pittsburgh, USA). He then joined the chemistry based on the metal nanoclusters, and Ag-based nanoclusters
faculty of Anhui University in 2010. His current research interests have also been summarized in detail in their review.21 There
include atomically precise nanoclusters, structure–property corre- are also some overviews focusing on a particular physico-
lation of nanoclusters, and applications. chemical property (such as photo-luminescence, chirality, cata-

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10759
View Article Online

Review Nanoscale

lysis etc.)23–25,27,38,110 or other special research field (for is the need for methods to prepare the Au25(SR)18 nanocluster
instance, inter-particle reaction or carbon-centred gold with high yield and purity. Extensive efforts have therefore
nanoclusters).116–118 However, these overviews just horizontally been made toward the syntheses and separation of Au25(SR)18
summarized the development of the nanocluster (i.e., reviewed nanoclusters.
almost all the species of nanoclusters and obtained the history 2.1.1 Earlier works on synthesizing Au25(SR)18 nano-
of the nanoclusters). Au25(SR)18, as the brightest molecular clusters. In the 1980s, research on self-assembling the mono-
star within the great nanocluster ship, is capable of revealing layer thiolates on the bulk gold surface reflected the beginning
the past developments and prospecting the future of metal of gold–thiol chemistry.119 Subsequently, for wider application
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

nanoclusters all by itself. However, up to the present, there has of these types of nanomaterials, an enormous amount of work
been no review focusing on the comprehensive and systematic was done to obtain ultra-stable and multi-functional
development of this star-nanocluster. Considering that (i) nanoparticles.1,120–125 After the initial establishment of the
explosive developments have been achieved in the nanocluster surface structure (composition)–property correlations, the syn-
field based on Au25 since it was discovered and (ii) the research thesis of gold–thiol nanoparticles with monodisperse size and
on the Au25(SR)18 will hopefully stimulate the further develop- atomically accurate surface structure became urgent for high-
ment of the nanoclusters as well as other nanomaterials, the purity nanoclusters.
excellent works based on the Au25(SR)18 nanocluster and its In 1994, Brust et al. reported the synthesis of gold nano-
derivatives should be re-visited via a review. particles within a two-phase liquid–liquid system, which led to
This review highlights the research achievements and more homogeneous nanoparticles relative to the previously
advances of Au25(SR)18, which cover the preparation methods, reported ones (Fig. 1a).124 Specifically, the gold source from
crystal structures, physicochemical properties (mainly on optics, the [AuCl4]− aqueous solution was transferred into the oil
chirality, electrochemistry, and magnetism), and practical appli- phase by the use of the phase-transfer reagent (tetraoctyl-
cations (mainly in catalytic and biological fields). In addition,
the marriage of Au25(SR)18 nanoclusters and other functional
nanomaterials (such as graphene, semiconductors, and metal–
organic frameworks) poses a promising approach to preparing
composite materials with intriguing properties, and is also sum-
marized. Furthermore, we summarized the inter-nanocluster
reactions since these works enable us to fully understand the
physicochemical properties of such nanoclusters. The research
processes of Ag25(SR)18 the silver-based nanocluster are also
covered because of its homologous composition, construction
and optical absorption compared to its gold-counterpart
Au25(SR)18. Moreover, considering that alloying is of importance
in the nanocluster field for controllably tailoring the properties
of nanoclusters, the alloying strategies, metal-exchange sites
and property alternations based on the templated Au25(SR)18
have been reviewed. Of note, the sequence of these sections
does not relate to the importance. Instead, each part is of sig-
nificance in researching the Au25(SR)18 nanoclusters. At the end
of this review, we would like to propose some perspectives and
challenges for the future research of the Au25(SR)18 nanocluster;
for instance, mechanisms of the structure evolution as well as
the property origination, details in biological and catalytic pro-
Fig. 1 “Brust–Schiffrin two-phase method” and its modifications. (a)
cesses, new strategies for preparing alloy nanoclusters with con- Synthetic procedure of the “Brust–Schiffrin two-phase method”.
trollable composition and enhanced properties, and so on. Note Reproduced from ref. 124 with permission from The Royal Society of
that these prospects also hold true for all members in the nano- Chemistry, copyright 1994. (b) “Brust–Schiffrin two-phase method” in
cluster field. We hope that this review will provide an updated Au nanoparticle synthesis in different thiolated-addition situations.
Reproduced from ref. 126 with permission from American Chemical
summary for the researchers interested in the related
Society, copyright 2010. (c) The mechanism for the chalcogenate-pro-
investigations. tected metal nanocluster synthesis by the “Brust–Schiffrin two-phase
method”. Reproduced from ref. 127 with permission from American
Chemical Society, copyright 2011. (d) Possible precursor species of gold
ions in the “Brust–Schiffrin two-phase method”. Reproduced from ref.
2 Preparation and purification 128 with permission from American Chemical Society, copyright 2011.
2.1 Preparation of oil phased Au25(SR)18 (e) Schematic illustrations of the steps involved in the “Brust–Schiffrin
two-phase method”. The aqueous phase and the toluene phase are
For investigating the properties of Au25(SR)18 at a deeper level drawn in blue and orange, respectively. Reproduced from ref. 129 with
and promoting the wider application of the nanocluster, there permission from The Royal Society of Chemistry, copyright 2017.

10760 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

ammonium bromide, TOAB), and then reduced with aqueous Au25(SR)18 has to be separated from the mixture on the basis
sodium borohydride (NaBH4) in the presence of dodecanethiol of several techniques, for example, liquid-phase extraction and
(C12H25SH). In these processes, they proposed that the Au(III) chromatography.132–136 In this context, the high-yield and
was firstly reduced to Au(I) induced by C12H25SH, and then high-purity synthesis of Au25(SR)18 is urgent. In 2008, the Jin
reduced into Au(0) to form the metallic kernel. Furthermore, group devised a kinetically controlled, thermodynamically
the Au(I) existed in the (Aum)(C12H25SH)n(C6H5Me) polymer selective strategy (Scheme 3a),92 which mainly had three modi-
form before the addition of reducing agent. The aforemen- fications relative to the previous “Brust–Schiffrin two-phase
tioned “Brust–Schiffrin two-phase method” is of significance method”. (i) The magnetic stirring of the solution was reduced
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

in Au colloid chemistry and has a profound effect on the later to a very low speed (∼30 rpm) at the temperature of 0 °C (ice-
synthetic strategies of Au nanoclusters. Of note, in the later bath) before the addition of HS-C2H4Ph, which resulted in the
research and modification of the “Brust–Schiffrin two-phase concentration of the size of Au(I)-SR polymeric intermediates.
method”, the thiol effect on the Au(III) salt was determined to (ii) Before the addition of NaBH4, the magnetic stirring contin-
be that of a weak reducing agent only, in which process the ued for almost 1 hour, and the solution gradually changed
RSH would not link to the generated Au(I) polymers, but just from faint yellow to clear; (iii) after the complete formation of
generate the independent RSSR dithiol (Fig. 1b–d).126–128 More the Au(I) polymer, the stirring was changed to relatively fast
recently, Booth et al. reported the significance of bromide in (∼1100 rpm) and the NaBH4 dissolved in water was immedi-
the “Brust–Schiffrin two-phase method” of thiol protected Au ately added. It should be noted that both the low temperature
nanoparticles. The species [AuBr4] is shown to be a preferable and slow stirring conditions are critical for the formation of
precursor in the synthesis as it is more resistant to the for- monodisperse Au(I)-SR polymeric intermediates as well as the
mation of Au(I) thiolate species than [AuCl4].129 In addition, final Au25(S-C2H4Ph)18. This was the first time that
four steps have been confirmed in the formation of Au nano- Au25(S-C2H4Ph)18 could be synthesized with high-yield and
particles: (i) the phase transfer process; (ii) the ion exchange -purity (cal. ∼40% yield, based on the Au source HAuCl4). The
process; (iii) the addition of alkanethiol causing a reduction to UV-vis spectrum of the crude product (without any purifi-
Au(I); (iv) the phase transfer process occurring on addition of cation) exhibited an obvious peak at 670 nm and two shoulder
NaBH4 leading to nanoparticle formation (as shown in peaks at 400 and 450 nm. The similar optical absorption
Fig. 1e).129 between the crude and pure products (with intense absorption
Murray and co-workers substituted the C12H25SH in the at 410 and 680 nm, and several shoulder peaks at 320, 350,
“Brust–Schiffrin two-phase method” and modified the syn- 560 and 800 nm) also verified the high-purity synthesis.
thetic procedure, and finally obtained the Au38(S-C2H4Ph)24 2.1.3 One-phase synthesis of the Au25(SR)18 nanocluster.
(finally unified as Au25(S-C2H4Ph)18 with the advances of In 2010, the Murray group presented a one-phase (tetrahydro-
mass-spectrometric techniques).91,130,131 The separated nano-
cluster was stable and isolable in relatively high purity, and
finally, about 200 mg quantities per reaction batch were
obtained. The rise of electrospray ionization mass spec-
trometry (ESI-MS) offered the opportunity to accurately charac-
terize the molecular weight of nanoclusters with mono-
disperse sizes.132,133 In 2007, the Murray group obtained the
ESI-MS spectra of HS-PEG (where HS-PEG represents
HS-(C2H4O)5-CH3) ligand-exchanged Au25(S-C2H4Ph)18. The
obtained Au25(S-PEG)x(S-C2H4Ph)18−x (x = 5–13) enabled ESI by
metal ion (i.e., Na+ from NaOAc) coordination with atomically
precise mass analysis.132 This work also demonstrated the −1
charge state of the prepared Au25(S-C2H4Ph)18.
It has been suggested that the surfactant plays an impor-
tant role in the synthesis of Au25(SR)18 nanocluster.124–133 Dass
and co-workers reported the surfactant-free synthesis of gold
nanoclusters in methylene chloride employing NaBH4 as a
reducing agent with the usual synthetic conditions.133
A mixture of Au25, Au38, Au102, and Au144 was obtained relative
to Au25 only in the presence of TOAB.133
2.1.2 Kinetically controlled, thermodynamically selective
synthesis of the Au25(SR)18 nanocluster. Despite the remark-
able aforementioned progress that has been made in the
Scheme 3 Schematic illustration of the formation of the
syntheses of Au25(SR)18, the synthetic procedures suffer from Au25(S-PhC2H4)18 nanocluster by (a) a kinetically controlled, thermo-
mixed productions with different nanocluster sizes, and thus dynamically selective two-phase strategy (corresponding to ref. 92) and
the yield of Au25(SR)18 is quite low.91,130–133 In addition, the (b) a one-phase in situ synthetic method (corresponding to ref. 93).

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10761
View Article Online

Review Nanoscale

furan, THF) synthetic method for (cal. ∼50% yield, based on Table 1 Solubility analysis of the [Au25(S-PhC2H4)18]−[TOA]+ nano-
the Au source in HAuCl4) preparing the monodisperse cluster (common reagents in the laboratory).
[TOA]+[Au25(SR)18]− in high yield (Scheme 3b).93 In this syn-
High Mild
thetic procedure, HAuCl4 and TOAB (TOA+Br−) with almost Nanocluster solvency solvency Non-solvency
1 : 1 of molar ratio were firstly added to the THF solution and − +
[Au25(S-PhC2H4)1x] [TOA] CH2Cl2; CH3CN; H2O; MeOH;
stirred for 15 min. HS-C2H4Ph was then added at room temp-
CH3Cl; Tol; DMSO EtOH; Hex;
erature in a 5-fold molar ratio. The solution was stirred for at DMF; THF Et2O; PE
least 12 h until it was completely colorless (in this process,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

the Au(I)SR polymeric intermediates were completely syn-


thesized). After that, the 5-fold molar ratio of NaBH4 (com-
pared to the Au moles) dissolved in ice-cold water was then Au25(S-PhC2H4)18 with “−1” charge state, we also summarized
rapidly added to the THF solution. The reaction was allowed the solubility analysis of this nanocluster in Table 1.
to quietly stir for more than 48 h, and then the obtained solu- 2.1.4 Au25(SR)18 nanoclusters protected by other thiol
tion was filtered and evaporated to get the crude product. The ligands. Apart from HS-C2H4Ph, oil-phase Au25 nanoclusters
toluene extract was washed several times with copious could also be capped by other thiol ligands.93 Two methods
amounts methanol to get the pure [TOA+][Au25(S-C2H4Ph)18−]. are mainly adopted: (i) in situ synthesis based on the above-
They also synthesized several other species of Au25 nano- mentioned synthetic methods with a variety of thiol ligands;
clusters protected by different thiols (i.e., HS-(CH2)5CH3, (ii) the ligand-exchange method from a templated Au25(SR)18.
HS-CH2CH(CH3)2, and HS-(CH2)11CH3), and all of these Functional Au25(SR)18 nanoclusters could be obtained by tai-
[Au25(SR)18]− nanoclusters in high yields using this synthetic loring the capped thiol ligands.
procedure. More importantly, without the addition of Compared with the ligand-exchange method, it is easier to
Oct4N+Br− in the synthesis, the obtained Au25(SR)18 is in the obtain the Au25(SR)18 nanoclusters with the mono-thiol protec-
oxidized state (that is, [Au25(SR)18]0, as opposed to the nega- tion by in situ synthesis. With the one-phase synthetic method,
tively charged state in [Au25(SR)18]−). Because this section the Murray group has obtained the mono-disperse Au25(SR)18
mainly focuses on the synthetic procedure, the comparisons nanoclusters protected by different thiol ligands (HS-
of Au25(SR)18 with different charge states are discussed in (CH2)5CH3, HS-CH2CH(CH3)2, and HS-(CH2)11CH3 ligands,
detail elsewhere (chapter 2.3). summarized in the One-phase synthesis section).93 Also, using
It should be noted that the as-prepared Au25(SR)18 nano- para-substituted thiophenols (HSPh-X) did not yield stable
clusters are actually contained in a mixture of Au25(SR)18 nano- Au25(SR)18 in this synthetic method, with or without TOAB.93
clusters, small-sized Aum(SR)n complexes, redundant thiol, and In 2011, using the chiral thiols PET* (where PET* rep-
some large-sized nanoclusters.92,93 The purification of resents chirally modified phenylethylthiolate -SCH2C*H-(CH3)
Au25(SR)18 is necessary and of great significance for the sub- Ph at the 2-position), we prepared the R- or S-Au25(PET*)18
sequent crystallization and property investigation. In the kine- nanoclusters with our previously reported two-phase
tically controlled, thermodynamically selective method for method.137 These optically active Au25(PET*)18 nanoclusters
synthesizing the Au25(S-PhC2H4)18 nanocluster,92 a large are close analogues of the optically non-active
amount of methanol is used to wash away the redundant salt, Au25(S-C2H4Ph)18 (Fig. 2a). Based on the atomically precise
thiol. This process needs to be repeated until the impurity is Au25(S-C2H4Ph)18, we explicitly revealed that the ligands and
all washed away. Acetonitrile is then used to extract the surface gold atoms of Au25(PET*)18 play a critical role in
Au25(S-PhC2H4)18 nanocluster from the nanocluster mixture. effecting the circular dichroism (CD) responses from the nano-
The obtained [Au25(S-PhC2H4)18]−[TOA]+ is pure enough for the clusters.137 Also with the chiral thiol ligands, another chiral
further crystallization and characterization. Au25 nanocluster protected by (R/S)-2-amino-3-phenylpropane-
In the one-phase synthetic method,93 after the reaction was 1-thiol (appt* for short, see the UV-vis and mass spectra in
completed, the product solution was gravity filtered to remove Fig. 2b) was obtained.138 Furthermore, we also synthesized the
the insoluble component. Then, the filtered solution was roto- 2-(naphthalen-2-yl)ethanethiolate (NAPS for short) protected
vapped to remove the tetrahydrofuran solvent. After this, Au25 nanocluster (Fig. 2c), which exhibited enhanced photo-
toluene was used to dissolve the product and the obtained luminescence relative to Au25(S-C2H4Ph)18 (about 6.5-fold
solution was washed several times with nanopure water. The enhancement).139
toluene solution was subsequently rotovapped to dryness and The Maran group prepared a series of Au25(S-CnH2n+1)18
washed thoroughly with methanol to remove the excess thiol clusters in which n = 2, 4, 6, 8, 10, 12, 14, 16, 18 (Fig. 2g). On
and TOAB, leaving the pure [Au25(S-PhC2H4)18]−[TOA]+. Of the basis of these Au25 nanoclusters, they studied how elec-
note, the purification and separation of other oil-phase trons tunnel through the 3D monolayers on the Au25(SR)18
Au25(SR)18 nanoclusters are almost similar to the nanocluster surface.140 They further obtained the crystal struc-
[Au25(S-PhC2H4)18]−[TOA]+. In addition, the solubility analysis ture of Au25(S-C2H5)18, which had electron nuclear double reso-
of the Au25(SR)18 nanocluster is of significance in the separ- nance.141 In addition, the linear polymer behavior was found
ation and purification processes. Considering that the afore- in the crystal structure of Au25(S-Bu)18 (where S-Bu represents
mentioned two synthetic methods all generate the n-butanethiolates).142 Note that Häkkinen and co-workers

10762 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 2 Optical absorption spectra and mass spectra of Au25 nanoclusters protected by different ligands. (a) UV-vis and MALDI-TOF-MS spectra of
the R- and S-Au25 nanoclusters protected by PET* ligands. Reproduced from ref. 137 with permission from American Chemical Society, copyright
2011. (b) UV-vis absorption and LDI mass spectra of the R- and S-Au25 nanoclusters protected by appt* ligands. Reproduced from ref. 138 with per-
mission from The Royal Society of Chemistry, copyright 2013. (c) UV-vis absorption of [Au25(S-PhC2H4Ph)18]− and [Au25(NAPS)18]1−, and the
MALDI-TOF mass spectrum of the [Au25(NAPS)18]1− nanocluster. Reproduced from ref. 139 with permission from The Royal Society of Chemistry,
copyright 2014. (d) Positive-mode MALDI-TOF mass spectrum of the [Au25(S-nC5H11)18]1− nanocluster. Reproduced from ref. 145 with permission
from American Chemical Society, copyright 2017. (e) UV-vis absorption spectra and MALDI mass spectra of the Au25(SBB)18 nanocluster in the posi-
tive ion mode. Reproduced from ref. 146 with permission from American Chemical Society, copyright 2014. (f ) Instrumental setup where the
monomer, dimer and trimer signals of Au25(SR)18 are generated, and the optimized aggregation mode of the Au25-Au25 dimer. Reproduced from ref.
144 with permission from The Royal Society of Chemistry, copyright 2016. (g) UV-vis absorption, MALDI-TOF mass, and 1H NMR spectra of mono-
disperse [Au25(S–Cn)18]0 (n = 2, 4, 6, 8, 10, 12, 14, 16, 18) and different sites of H in the Au25 structure which correspond to the 1H NMR spectra.
Reproduced from ref. 140 with permission from American Chemical Society, copyright 2014. (h) Left: comparison of UV-vis absorption spectra of
Au25 in the crude product and pure crystals; middle: ESI-MS spectra of the Au25(S-PhC2H4)18 nanocluster; right: UV-vis spectra of the crude products
of Au25(SR)18 nanoclusters with different capped ligands. Reproduced from ref. 152 with permission from The Royal Society of Chemistry, copyright
2009. (i) Left: Comparison of UV-vis absorption of the Au25 nanoclusters with different ligands; right: MALDI-TOF mass spectrum of [Au25(SNap)18]1−
nanocluster. Reproduced from ref. 154 with permission from the American Chemical Society, copyright 2016.

theoretically researched the assembled action and packing Au25(S-C2H4Ph)18 using ion mobility mass spectrometry
mode of Au25(S-C2H4Ph)18 in the crystal (to be dimers), and (Fig. 2f ).143,144 Very recently, the Maran group successfully
the Pradeep group detected the dimer formation of obtained the crystal structure of Au25(S-nC5H11)18 (as well as

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10763
View Article Online

Review Nanoscale

three other known Au25(SR)18 nanoclusters; see the mass spec- more thiolates) can be synthesized through the ligand-
trum of Au25(S-nC5H11)18 in Fig. 2d) using the electrocrystalli- exchange method.
zation (emphasized in section 3.1).145 In 2007, the Murray group carried out the ligand-exchange
In 2014, Au25(SBB)18 (where SBB represents 4-(t-butyl)benzyl process on the Au25(S-C2H4Ph)18 with several different thiol
mercaptan) was synthesized by the Pradeep group (Fig. 2e).146 ligands, such as HS-Ph, HS-C6H13, HS-PEG-biotin (biotiny-
In addition, the specific host–guest interaction between the lated), and HS-PhCOOH ligands.157 Even HS-Ph can be
β-cyclodextrin (CD) and the ligands anchored on the nano- exchanged on the surface of the Au25 nanocluster. However,
cluster was exploited to obtain the Au25(SBB)18@CDn (n = 1–4) only one HS-PEG-biotin can be exchanged on the Au25
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

composite nanomaterials, which displayed enhanced stability (Fig. 3a). These results suggest that just a part of the
relative to the nanocluster only.146 Interestingly, the relatively -S-C2H4Ph ligands could be ligand-exchanged to generate the
unstable Au25(SBB)18 was altered into Au29(SBB)24S with the bi-thiolate co-protected Au25(S-C2H4Ph)18−x(SR)x nanoclusters.
process of thermal treatment followed by rapid cooling.147 It should be noted that the mass peaks of
In 2016, Ishida et al. reported the synthesis and characteriz- (Cs@Au25(S-C2H4Ph)18)2+, (Cs2@Au25(S-C2H4Ph)18)2+, and
2+
ation of Au25 nanoclusters fully protected by cationic thiolates (Cs3@Au25(S-C2H4Ph)18) have been detected on the pure
(i.e., Au25(SR+)18, where SR+ represents 11-mercaptoundecyl)- Au25(S-C2H4Ph)18, which reveals the existence of −1, 0, +1
N,N,N-trimethylammonium chloride, S-(CH2)11N(CH3)3+·Cl−:SR+, charge states of Au25(S-C2H4Ph)18, respectively.157 They sub-
which filled the gaps of cationized Au nanomaterials in the sequently reported the intact mass peaks of the
nanocluster field.148 Au25(S-C2H4Ph)18 nanocluster and its CH3(OCH2CH2)ySH (y =
Interestingly, Au25(SPhNH2)17, with fewer thiol ligands com- 1, 5) ligand-exchange products (also partial exchange) in the
pared with other common Au25(SR)18 nanoclusters, was matrix-assisted laser desorption/ionization time of flight mass,
reported by Demessence and co-workers.149 The molecular using the trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenyli-
formula was determined by ESI-MS. It should be noted that dene]malononitrile (DCTB) as the matrix.158 Firstly, the intact
the electrically neutral Au25(SPhNH2)17 has 8e free electrons mass peaks with DCTB had a clear distinction compared to
(25 − 17 = 8), which is the same as the Au25(SR)18 with a nega- the previously reported Au nanocluster peaks with a large
tive valence state of −1. amount of fragment signals.130,159–168 More importantly,
It is worth noting that the in situ synthetic procedure could MALDI-TOF-MS results suggested that up to 12 CH3(OCH2CH2)
produce the multi-thiolate co-protected Au25(SR)18 in the pres- SH (PEG1 for short) could be exchanged on the surface of the
ence of different thiol ligands. Hassinen et al. demonstrated a Au25 nanocluster, resulting in the Au25(S-PhC2H4)6(PEG1)12
simple one-pot procedure to synthesize fluorescent Au25 nano- nanocluster. In contrast, all of the S-PhC2H4 ligands on the
clusters carrying controlled amounts of bulky calix[4] arene Au25 surface can be substituted by S-C6H13, and thus generate
functionalities.150 In this work, the bulk thiol ligands have the mono-thiolated Au25(S-C6H13)18 nanocluster (Fig. 3b).158 In
been successfully introduced into the ultra-small, atomically 2010, bi-thiolate co-protected Au25(SR)18 was obtained via the
precise nanoclusters.150,151 ligand-exchange method on the Au25(S-C2H4Ph)18 nanocluster
Wu et al. reported a one-pot method for synthesizing atom- with electron-withdrawing substituents (HSPh-p-X; X = Br or
ically monodisperse Au25 nanoclusters stabilized with various NO3), and 1H nuclear magnetic resonance was used to monitor
functionalized thiols, such as HS-C11H22-OH and the ligand-exchange processes.169 In the same year, they
HS-C11H22OC(O)C(CH3)2Br (Fig. 2h).152 Among these nano- exchanged the Au25(S-C2H4Ph)18 with a dithiol (i.e., toluene-
clusters, Au25(S-MI)18 is of particular interest as it bears a 3,4-dithiol, CH3Ph(SH)2), and the ESI and MALDI results
polymer initiator for the atom-transfer radical polymerization suggested the generation of Au25(CH3Ph(S-)2)x(CH3Ph(S-)
reaction and thus allows for preparing Au nanoparticle– (SH))y(S-C2H4Ph)18−2x–y nanoclusters, which could be called a
polymer composite architectures.153 tri-thiolate co-protected Au25(SR)18 (Fig. 3c).170 Using the
There are also several functional Au25 nanoclusters pro- MALDI-TOF, the statistical aspects of ligand populations in
tected by the single type of thiol ligands, for example, mixed monolayer Au25(SR)18 nanoclusters have been fully ana-
Au25(S-Nap)18 (where S-Nap represents 1-naphthalenethiolate) lyzed, which demonstrates that the S-C2H4Ph host ligands are
with enhanced catalytic properties and photoresponsive more easily substituted by S-C6H13 relative to S-Ph ligands
Au25(S-Az)18 (where S-Az represents azobenzene derivative (Fig. 3d).171
thiols; the UV-vis and mass spectra of Au25(S-Az)18 are shown For obtaining the Au25 nanoclusters with more enhanced
in Fig. 2i).154–156 The function-directed syntheses of these Au25 physicochemical properties, several functionalized ligands
nanoclusters are further summarized in the sections on pro- have been exchanged on the Au25 surface by substituting the
perties and application. parent ligands. The Bürgi group reported the ligand exchange
The ligand-exchange method for the templated Au25(SR)18 on Au25(S-C2H4Ph)18 with two chiral ligands R/S-BINAS and
is not only capable of obtaining the Au25 nanoclusters with NILC/NIDC (where the R/S-BINAS represents R/S-1,1′-
novel capped ligands, but also provides a platform for binaphthyl-2,2′-dithiol; the NILC/NIDC represents
thoroughly analyzing the intermolecular reaction process on N-isobutyryl-L-cysteine/N-isobutyryl-D-cysteine).172 Resultantly,
the nanocluster surface. In addition, unlike the in situ syn- the exchange products with significant chiral activity were
thetic procedure, bi-thiolate co-protected Au25(SR)18 (even obtained. After this, they also analyzed the reaction products

10764 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 3 Number of exchanged ligands on the Au25(SR)18 nanocluster determined by mass spectroscopy. (a) ESI-MS spectra to confirm the number of
exchanged HS-Ph, HS-C6H13 and HS-PEG-biotin ligands on the surface of the Au25(S-PhC2H4)18 nanocluster. Reproduced from ref. 157 with per-
mission from American Chemical Society, copyright 2007. (b) MALDI-TOF-MS spectra to confirm the number of exchanged PEG1 and S-C6H13
ligands on the surface of Au25(S-PhC2H4)18 nanocluster. Reproduced from ref. 158 with permission from American Chemical Society, copyright
2008. (c) ESI-QQQ mass and MALDI mass spectra to detect the number of exchanged ligands including monothiol and dithiol. Reproduced from ref.
170 with permission from American Chemical Society, copyright 2010. (d) Monolayer ligands distributions of the ligand-exchange products
Au25(S-PhC2H4)18−x(SC6)x and Au25(S-PhC2H4)18−x(S-Ph)x as observed by MALDI-MS spectra. Reproduced from ref. 171 with permission from
American Chemical Society, copyright 2008.

using MALDI-TOF mass spectrometry.173 Lee and co-workers nance), and mass spectrometry measurements illustrated that
firstly synthesized the HS-C6H12 protected Au25(S-C6H12)18 spiropyran bound to the nanoclusters isomerizes in a revers-
nanocluster. Then, the pyrene-labeled Au25 nanocluster was ible fashion when exposed to UV and visible light. It should
synthesized by the exchange of thiolated pyrene (HS-py) onto also be noted that the photoswitchable fluorescence modu-
prepared Au25(S-C6H12)18, resulting in the lation on the nanoclusters was achieved accompanied by this
Au25(S-C6H12)17(S-Py)1 nanocluster.174 Importantly, this work reversible fashion, which is beneficial for developing novel
unambiguously showed that the obtained probes for super-resolution imaging.176
Au25(S-C6H12)17(S-Py)1 nanocluster can work as an electron Due to the atomically accurate structures, the property com-
donor with electrochemical and optical measurements. parison of these Au25 nanoclusters sheds light on the full
Klajn and co-workers carried out the ligand-exchange reac- understanding of the surface structure–property
tion on the Au25(SR)18 nanocluster to synthesize a spiropyran correlations in the nanocluster field, which is analyzed in
molecular switch.175 The optical, NMR (nuclear magnetic reso- detail in section 4.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10765
View Article Online

Review Nanoscale

The aforementioned ligand-exchange results reveal that this


is always an incomplete substitution process (also likely to be a
step-wise process), and thus, each ligand at the different sites of
the Au25(SR)18 nanoclusters should reflect different activity in
this exchange process. Gaining clear insight into this process is
of importance for the controllable synthesis of target nano-
clusters with different structures (compositions) and properties.
The Ackerson group reacted the Au25(S-C6H12)18 nanocluster
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

with a 5-fold molar excess of p-bromobenzenethiol (pBBT), and


the ligand-exchange process was stopped after reacting for
7 minutes.177 The crystal structure of the resulting
Au25(S-C6H12)16(pBBT)2 illustrated that the two ligand-exchange
positions were on the symmetric sites and bonded to the most
solvent-exposed Au atom in the structure (Fig. 4).177 For the first
time, the ligand exchange process was structurally resolved on
the Au25(SR)18 nanocluster, and more importantly, it unambigu-
Fig. 5 Positive-ion MALDI mass spectra monitored at different times
ously revealed that the ligand exchange on the Au25(SR)18 nano-
after the mixing of Au25(S-PhC2H4)18 ( purple) and Au25(S-tBu)18 (orange)
cluster was a step-wise process. In 2015, the Aikens group theo- nanoclusters with a ratio of 1:1 in CH2Cl2. Reproduced from ref. 180
retically investigated the kinetics of the ligand-exchange process with permission from American Chemical Society, copyright 2017.
based on the structure of Au25(SR)18 (where SR represents S-CH3
in this work), and three possible ligand exchange sites were pro-
posed.178 Based on the results of proton nuclear magnetic reso- Accordingly, the Au–Se bond is more covalent and has a
nance, the Murray group reported the second-order rate con- higher bond energy than the Au–S bond, and thus, Se-based
stants of the associative ligand exchange of Au25(S-C6H12)18 ligands may further increase the stability of relevant
nanoclusters with different charge states and concluded that the nanoclusters.96,127,182–185 Pei and co-workers theoretically com-
electron depletion retarded the ligand exchange.179 Very recently, pared the Au nanoclusters protected by S- and Se-based
Bürgi and co-workers investigated the factors that influence the ligands, and found that almost all of the selenolate-protected
ligand exchange process in the reaction of the Au25(SR)18 nano- Au nanoclusters exhibited higher thermal stability than those
cluster with excess different thiol ligands, and demonstrated protected by thiolate.186
that the thiolate monolayer of the Au25 nanoclusters have a Negishi et al. firstly reported the in situ synthesis of the
dynamic nature.180 Also, in this work, MALDI mass spectroscopy Au25(SeC8H17)18 nanocluster by reducing the Au(I)-SeR precur-
was performed to monitor the ligand distribution on the Au25 sor (Scheme 4a).187 The negative-ion ESI spectrum revealed the
nanocluster after the mixing of Au25(S-PhC2H4)18 and Au25(S-
tBu)18 nanoclusters (Fig. 5). Pengo et al. analyzed the ligand
exchange reaction sites on the neutral Au25(S-C2H4Ph)18 nano-
cluster by using 4-fluorobenzylthiol and a series of substituted
arylthiols, and the selectivity for the inner and outer positions of
the dimeric staples of the cluster could be modulated by using
incoming thiols with different structures.181
2.1.5 Syntheses of Se or Te ligand protected Au25(SePh/
TePh)18 nanoclusters. In comparison to the S atom, the atomic
radius and electronegativity of Se are closer to those of Au.

Fig. 4 Ligand-exchange (crystal structure of Au25(S-PhC2H4)16( pBBT)2)


of pBBT on the surface Au25(S-PhC2H4)18 nanocluster. Color code: Scheme 4 Schematic illustration of the formation of the Au25(Se-R)18
orange, Au; yellow, S; gray, C; red, Br. Reproduced from ref. 177 with nanocluster through (a) in situ synthesis (corresponding to ref. 187) and
permission from American Chemical Society, copyright 2014. (b) ligand-exchange methods (corresponding to ref. 192).

10766 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

−1 charge state of Au25(SeC8H17)18. After this work, the bi-


metallic Au25−xCux(SeC8H17)18 was also synthesized by the
direct synthetic method.188 It should be noted that the
obtained Au25−xCux(SeC8H17)18 exhibited enhanced stability
relative to the analogous nanoclusters containing thiolate
ligands.188,189 The higher stability of the selenolate-protected
Au25 nanoclusters relative to the thiolated-protected ones
was further confirmed by the Negishi group by comparing
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

the degradation in solution, thermal dissolution, and laser


fragmentation of synthesized [Au25(SC8H17)18]− and
− 190
[Au25(SeC8H17)18] nanoclusters. Pohjolainen et al. theoreti-
cally investigated the role of both the anchor atom and the
organic ligand on the Au25 nanoclusters protected by thiolates,
selenolates, and tellurolates.191 They found that the changes in
the anchor atom and R group had a pronounced effect on the
optical spectra, which could be attributed to the change in
molecular orbital splitting and geometry changes of Au25
nanoclusters.191
Similar to the Au25 nanoclusters protected by other thiol
ligands (summarized in section 2.1.4.), the selenolate- or tell-
urolate-protected Au25 nanoclusters could be obtained by the
ligand-exchange strategy from thiolated-protected Au25 nano-
cluster. In 2012, we obtained the selenophenolate-capped Au25
by ligand exchange of the Au25(S-C2H4Ph)18 with a small
amount of HSe-Ph ligand (Scheme 4b).192 It is worth noting
that the selenol ligand is capable of completely substituting
the thiol ligands on the Au25 surface, which is in sharp
contrast to the ligand-exchange process with thiolate
ligands.150,157,158,174,175 Also, with the ligand-exchange
method, the Negishi group obtained the Fig. 6 The degree of ligand-exchange is dependent on the HOMO–
Au25(TePh)n(SC8H17)18−n bi-ligand co-protected Au25 LUMO (the highest occupied molecular orbital and lowest unoccupied
nanocluster.193 molecular orbital) transition gap of (a, b) Au25(SR)18−x(SeR)x and (c, d)
The atomically accurate structure of Au25(SePh)18 was Au25(SR)18−x(TeR)x monitored by DPV curves or optical absorption
spectra. Reproduced from ref. 195 with permission from American
obtained in 2014, which was synthesized by the in situ syn-
Chemical Society, copyright 2016.
thetic method.194 The Au25(SePh)18 showed noticeable differ-
ences from the previously reported Au25(S-C2H4Ph)18 counter-
part, albeit, both of them share the icosahedral Au13 kernel
and six Au2(SeR)3 or Au2(SR)3 staple motifs. In addition, dis- nanoclusters benefit from high biocompatibility, good photo-
tinct differences in the electronic structures and optical, cata- stability, and low toxicity, and their use in cell labeling, photo-
lytic and electrochemical properties were revealed.195 Recently, therapy, and biosensing is thus highly
the Negishi group achieved the stepwise ligand exchange promising.98,112,114,115,197–205 High-purity and -yield synthesis
process from Au25(S-C2H4Ph)18 to Au25(SePh)x(S-C2H4Ph)18−x of these water-soluble nanoclusters is fundamental for accu-
(where x could be separated from 1 to 18).195 Combining the rately learning about the interactions between the nano-
ESI, differential pulse voltammetry (DPV), optical absorption clusters and bio-receptors.
spectroscopy measurements and the crystal structure of an 2.2.1 Syntheses and separation of Au25(SG)18 (where SG
intermediate product, the preferential exchange sites, as well represents glutathione). In earlier work, the Whetten group
as the ligand-exchange effects, on the electronic structure of synthesized a large amount of glutathione (GSH for short)
Au25, have been fully researched. Specifically, the HOMO– capped Au nanoclusters with molecular weights of 4.3, 5.6, 8.2
LUMO transition gap decreases with the increasing degree of and 10.4 kDa.161,162 Among them, the nanocluster with the
ligand-exchange (Fig. 6).195 molecular weight of 10.4 kDa was separated by polyacrylamide
gel electrophoresis (PAGE) and is the most researched. Due to
2.2 Preparation and separation of water-phase Au25 the undeveloped technology on the molecular mass detection,
nanoclusters this nanocluster was initially assigned as Au28(SG)16.206,207 The
Au25(SG)18, with same numbers of Au atoms and capped Tsukuda group performed PAGE on the mixture of Aun(SG)m
ligands as Au25(SR)18, is believed to possess analogous struc- nanoclusters and a series of relatively pure products was separ-
tures to the oil-phase cluster.196 Significantly, water-soluble Au ated and assigned to Au10(SG)10, Au11(SG)11, Au12(SG)12,

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10767
View Article Online

Review Nanoscale

Au15(SG)13, Au18(SG)14, Au22(SG)16, Au22(SG)17, Au25(SG)18,


Au29(SG)20, Au33(SG)22, Au35(SG)22, Au38(SG)24, and Au39(SG)24,
respectively (Fig. 7). Significantly, the nanocluster with
10.4 kDa molecular weight was identified as Au25(SG)18.135,136
Of note, the crystal structures of Au18, Au25 and Au38 protected
by thiolated ligands have been reported, which indicate the
same number of capped ligands as SG protected
ones.43,44,51,52,64,66
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

For these reported Au25(SG)18, the low purity is a major


issue because the target Au25(SG)18 should be separated from
the Aun(SG)m mixture with PAGE;135,136,161,162,206,207 high-
purity synthesis remains challenging at that stage. In 2005,
Shichibu et al. reported the large-scale synthesis of Au25(SG)18
from the phosphine-stabilized Au11 nanocluster using the
ligand exchange method.208 Of note, in 2014, Hutchison and
co-workers found that the ligand-exchange from Au11 to
Au25(SG)18 could be achieved on Au11(PPh)7Cl3. As for
[Au11(PPh3)8Cl2]Cl, the products yielded only small nano-
clusters relative to Au25(SG)18 (Fig. 8).209 The purity of the
obtained Au25(SG)18 was significantly improved as compared
with the previous in situ synthetic strategy.208,209 They also
found that the Aun(SG)m (n < 25) nanoclusters were completely
oxidized to Au(I)-SG complexes, while Aun(SG)m (n > 25) nano-
clusters were etched into Au25(SG)18 by free GSH molecules.
This conclusion opened up the possibility for the selective syn-
thesis of Au25(SG)18 on a large scale.135
Xie and co-workers reported the slow reduction method to
Fig. 8 Results of the ligand exchange of the Au11(PPh)7Cl3 and
synthesize Au25(SG)18.210 Importantly, NaBH4, the reducing
[Au11(PPh3)8Cl2]Cl with GSH. Reproduced from ref. 209 with permission
agent in previous studies (see ref. 135, 136, 161, 162, 206, 207 from American Chemical Society, copyright 2014.
and 211) was adjusted in CO for slowing down the reduction
rate. Furthermore, the correlation between the pH of the solu-
tion and the size of the obtained nanoclusters was obtained.
Specifically, the products would be Au10–12(SG)10–12,
Au15(SG)13, Au18(SG)14, and Au25(SG)18 when the pH of the
solution was regulated to be 7, 9, 10, and 11, respectively
(Fig. 9). It should be noted that the gram-scale production was
achieved with the NaOH-mediated NaBH4 reduction
method.212 By balancing the reduction rate (controlling the pH
of the solution as well as reducing the CO), they also achieved
the gram-scale production of several other water-soluble Au25
nanoclusters, such as Au25 protected by HS-CnH2n-COOH (n =
2, 7, 10).212 Considering that the synthesis of water-soluble
Au25 nanoclusters typically consists of a pair of reversible reac-
tions (i.e., a fast reduction-growth reaction and a slow size-
focusing reaction), they introduced the high-temperature syn-
thesis method to more efficiently synthesize the Au25 nano-
clusters (with >95% yield), and discussed the temperature
effects on the synthesis of Au25 nanoclusters.213
2.2.2 Synthesis of water-soluble Au25 nanoclusters pro-
tected by other ligands. Synthesizing the Au nanoclusters with
the same metal number, but different ligands (Scheme 5), not
only leads to multi-functional nanoclusters with similar struc-
tures, but also allows insight into the structure–property corre-
Fig. 7 PAGE results of the mixture Aun(SG)m nanoclusters and the
corresponding low-resolution ESI MS as well as the Au(4f ) XPS spectra.
lation. The Xie group synthesized a fluorescent Au25 nanocluster
Reproduced from ref. 136 with permission from American Chemical protected by a common protein (i.e., bovine serum albumin,
Society, copyright 2005. BSA), and the quantum yield (QY) of BSA–Au25 was about 6% at

10768 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

λem of 640 nm;214 noted that this fluorescent nanocluster has


been widely used in the ion detection and bioimaging fields
(vide infra in section 5.1). After this work, they also prepared the
BSA-protected Au4, Au8, Au10, Au13, and Au25 nanoclusters by
just tailoring the conformation of proteins.215 Note that the
BSA–Au25 nanocluster presents a long fluorescence lifetime
and thus results in polarization behaviour, which can be
explored for time-gated detection in microscopy and
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

imaging.216
In 2012, Yuan et al. reported the fast synthesis of
Au25(Cys)18 (where Cys represents cysteine) nanoclusters by
using a protection–deprotection strategy.217 They introduced a
surfactant molecule (cetyltrimethylammonium bromide,
CTAB) to protect Cys-Au(I) complexes during the reduction.
The surfactant CTAB was then removed from the surface of the
acquired CTAB-capped Au25(Cys)18 nanomolecule (dissolved in
toluene), and the water-soluble Au25(Cys)18 was obtained. Note
that this protection–deprotection method can synthesize high-
purity Au25(Cys)18 within 10 min.
The water-soluble Au25(Capt)18 (where Capt represents (2S)-
1-[(2S)-2-methyl-3-sulfanylpropanoyl]pyrrolidine-2-carboxylic
acid, captopril) nanocluster was prepared by the Jin group.218
The Au25(Capt)18 exhibited high thermal stability compared
with Au25(SG)18 because of the enhanced ligand stability of
captopril compared with GSH. In addition, this nanocluster
Fig. 9 Production of the water-soluble Au10–12, Au15, Au18, and Au25
nanoclusters capped by GSH ligands by adjusting the pH of the reaction
and other chiral ligand-capped Au25 nanoclusters showed dis-
solution in the CO-reduction method, and the corresponding UV-vis tinct chiral properties.
and ESI mass spectra. Reproduced from ref. 210 with permission from Shivhare et al. reported the synthesis of 11-mercaptoun-
American Chemical Society, copyright 2013. decanoic acid (11-MUA for short) and 16-mercaptohexadecanoic
acid (16-MUA for short) protected Au25 nanoclusters using a
NaBH4 purification strategy, and the yield of the synthesis was
15%.219
The ligand-exchange method has also been applied to syn-
thesize water-soluble Au25 nanoclusters protected by functio-
nalized ligands. For instance, the Pradeep group exchanged
the Au25(SG)18 with a water-soluble 3-mercapto-2-butanol (MB)
ligand, and the excitation spectrum of the MB-exchanged Au25
was entirely different from the acetyl and formyl exchanged
Au25 products.220

2.3 Mechanistic insights into the formation of Au25


nanoclusters
It is hoped that understanding the mechanism of the for-
mation process of Au25 nanoclusters will facilitate the syn-
thesis of atomically accurate nanoclusters with desirable sizes,
compositions, and functions.105 Research on the mechanism
has attracted wide attention since the production of thiolate-
capped Au nanoparticles by Brust and co-
workers.87,124–128,136,221–226 However, the synthetic procedure
of the oil-phase Au25 nanocluster is hard to track because of
the fast and complex synthetic environment. Thanks to the
slow formation ratio of the water-soluble Au25 nanoclusters,
researchers can understand the formation of Au25 (involving
Scheme 5 Water-soluble Au25 nanoclusters protected by different
the reduction-growth and the size-focusing processes) on a
ligands (ref. 136 for GSH, ref. 214 for BSA, ref. 217 for Cys, ref. 218 for deeper level.221–223 In this context, we mainly focus on the for-
Capt, ref. 219 for 11-MUA, ref. 222 for m-MBA, ref. 223 for p-MBA). mation of the water-soluble Au25 nanocluster in this section.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10769
View Article Online

Review Nanoscale

The Xie group monitored the whole process of the for- stage, in which the intermediate Aun(SR)m nanoclusters were
mation of Au25(Cys)18 nanocluster (reducing the Au(I)-Cys com- size-focused into the monodisperse Au25(m-MBA)18 nano-
plexes by gaseous CO), and this was the first time that the clusters.222 More importantly, the evolution of the free electron
growth mechanism of atomically precise nanoclusters was number was determined, from 0e for Au(I)-MBA and some
detected.221 In the reaction solution, distinct color changes small nanoclusters (Au number ≤10) to 2e (Au number from
(colorless → yellow → orange → brown → red-brown) were 11 to 15), 4e (Au number from 15 to 18), 6e (Au number from
detected by optical absorption and ESI-MS measurements, 19 to 23), 10e (Au number of 29, finally decomposed), and
which allowed the formation of Au25(Cys)18 to be reconstructed finally to 8e (Au25(m-MBA)18 nanocluster) (Fig. 10). This is in
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

from several key intermediates. Specifically, the formation of good agreement with the suggestion that atomically precise Au
Au25(Cys)18 underwent three identifiable stages: (i) the nanoclusters might be produced via the kinetically controlled
reduction of Au(I)-Cys complexes to generate Au10–15 nano- formation of intermediate nanoclusters, followed by the size
clusters; (ii) the nanoclusters growth from Au10–15 to Au16–25 focusing of these nanoclusters from various intermediate
nanoclusters; (iii) the transformation or size-focusing of the sizes.227,228
polydisperse Au16–25 to the monodisperse Au25(Cys)18 Very recently, the Xie group systematically investigated the
nanocluster.221 size growth mechanism at the molecular level by monitoring
Considering that the Au25(m-MBA)18 (where m-MBA rep- the reaction from the Au25( p-MBA)18 (where p-MBA represents
resents 3-mercaptobenzoic acid) nanocluster is more tolerant para-mercaptobenzoic acid) to a relatively larger Au44( p-mba)26
of fragmentation in mass spectrometry analysis relative to nanocluster.223 Through analyzing 35 intermediate species in
Au25(Cys)18, Xie and co-workers monitored the growth mecha- this transformation process, two different size-evolution pro-
nism of Au25 nanocluster again by replacing the ligand of Cys cesses were detected: monotonic LaMer growth and volcano-
with m-MBA.222 Through the results of ESI-MS and UV-vis shaped aggregative growth. Additionally, the evolution of the
measurements, they mapped out all the intermediate nano- free electron number in the formation of Au25(3-MBA)18 was
cluster species from Au(I)-MBA complexes to Au25(m-MBA)18 further extended to the following: 8e → 10e → 12e → 14e →
nanoclusters, thereby gaining new insights into the mecha- 16e → 18e → 20e.222,223
nism involved in the formation of the Au25 nanocluster In contrast, Whetten and co-workers researched the photo-
(Fig. 10). Similar to the formation process with Au25(Cys)18,221 dissociation of the Au25( p-MBA)18 nanocluster under the ultra-
two stages were detected: (i) the reduction-assisted growth violet irradiation.229 Activation, including collisional- and elec-
stage, in which the Au(I)-SR complexes were reduced to form tron-based methods, produced relatively few fragment ions
the intermediate Aun(SR)m nanoclusters; (ii) the size evolution from the intact Au25( p-MBA)18, and even a single ultraviolet
pulse (at λ = 193 nm) would cause extensive fragmentation of
the positively charged Au25 nanocluster.229
Ackerson and co-workers found that the presence of oxygen
in thiolate-protected gold nanocluster synthesis influenced the
product distribution.230 Specifically, the radicals (from O2 or
radical initiators such as 4-hydroxy-TEMPO) are necessary
components for the synthesis of the water-soluble Au25 nano-
cluster, as well as the etching process of aqueous colloidal
gold by thiols.230

2.4 Au25 nanoclusters with different charge states


The Au25(SR)181− nanocluster, containing 25 gold atoms as
well as 18 capped ligands, is an 8e− system that corresponds to
a noble gas-like 1S21P6 superatom electron configuration. The
electron configuration of Au25(SR)18 can be modified as
1S21P6, 1S21P5, and 1S21P4 in the Au25(SR)181−, Au25(SR)180,
and Au25(SR)181+ nanoclusters, respectively. Note that the
aforementioned Au25(SR)18 prepared by two-phase methods
are almost in the negative charge state, and in these synthetic
procedures the TOAB not only serves as a phase transfer agent,
but also as a counterion to balance the charge.91,92,132
Au25(SR)18 nanoclusters with different charge states possess
Fig. 10 Time evolution of UV-vis absorption and ESI-MS spectra of the diverse properties including magnetism, optical absorption,
reactants during the synthesis of [Au25(m-MBA)18]− (Left). Normalized
catalytic activity and stability, which can be rationalized in
ESI-MS spectral intensity profiles of the complexes or the nanoclusters
throughout the synthesis, and the final UV-vis and mass spectra of the
terms of the different superatom electron configurations. In
Au25 nanocluster (Right). Reproduced from ref. 222 with permission this section, the syntheses of Au25(SR)18 with different charge
from American Chemical Society, copyright 2014. states are summarized, and the physicochemical properties of

10770 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

these nanoclusters are briefly compared (see the section 4 for using H2O2.234 Considering that the count of the free electrons
more detailed analyses). of Au25 with neutral state is 7 (that is, 25 − 18 = 7), which
In 2007, the Tsukuda group revealed that the Brust method reveals the existence of lone pair electrons, they monitored the
yielded [Au25(SC6H13)18]x nanoclusters (x = −1, 0, +1) with a electron paramagnetic resonance (EPR) measurement on Ag25
distribution in charge states.231 [Au25(SC6H13)18]1+ and with neutral state. EPR quantification of the Au250 indicated
[Au25(SC6H13)18]1− can be prepared by chemical oxidation and the existence of one unpaired spin per nanocluster.
reduction from [Au25(SC6H13)18]0, respectively. The ESI mass Furthermore, the magnetic Au250 and non-magnetic Au251−
spectrometric approach was employed in this work to deter- could be mutually converted using H2O2 as the oxidant or
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

mine the chemical compositions and charge states of these NaBH4 as the reducer. The intrinsic magnetism in the
Au25 nanoclusters (Fig. 11a); all of the crystal structures of [Au25(S-C2H4Ph)18]0 nanocluster and the reversible switching
Au25 with different charge states have been solved.51,52,232,233 render these nanoclusters useful as paramagnetic probes.234
In 2008, the Jin group reported the Au25(S-C2H4Ph)18 with The Jin group probed the structure, the charge state, and
neutral charge state by oxidizing the [Au25(S-C2H4Ph)18]1−.232 surface thiolate ligand distribution of the water-soluble
In this state, the [Au25(S-C2H4Ph)18]1− lost one electron in the Au25(SG)18 nanocluster by means of NMR in combination with
presence of air and was converted into [Au25(S-C2H4Ph)18]0, mass spectrometry.235 LDI (laser desorption ionization) mass
and then exhibited chemical inertness. In addition, several spectra and NMR results revealed the −1 charge state of the
differences in the optical absorption were observed by compar- Au25(SG)18 nanocluster. Additionally, they demonstrated that
ing the Au25 with the neutral or reductive state (Fig. 11b). all Au25(SR)18 nanoclusters capped by different types of thio-
Specifically, the 400 nm peak became more prominent while late ligands should adopt a common two-shell structure,
the 450 nm peak decreased. Concurrently, the shoulder band which was caused by the particular stability of the core–shell
at 800 nm (which is the characteristic of the Au25− nanocluster) configuration of Au25(SR)18. Based on the anionic or neutral
disappeared and a new, small shoulder band emerged at Au25, the Jin group analyzed the charge state effects on ultra-
630 nm; after this work, they also oxidized the Au25− into Au250 fast electron relaxation dynamics.236 For both nanoclusters
(Au25 with −1 or 0 charge states), photoexcitation occurred in
two nondegenerate states near the HOMO–LUMO gap, which
were derived from the core orbitals. However, a large difference
in the lifetime of the core excitations was observed.
[Au25(SR)18]− exhibited a decay rate more than 1000 times
slower than the neutral one.236
In 2008, Murray and co-workers measured the rate constant
and activation energy barrier for electron self-exchanges of the
Au25(S-C2H4Ph)18 nanoclusters with −1 and 0 charge states.237
The α-methylene (on the thiolate ligand) proton resonances of
electrolytically prepared CD2Cl2 solutions of the Au250 and
Au25− nanoclusters exhibited characteristic chemical shifts
and line-shapes. Also, they found that the barrier energy was
larger than the calculated estimate of the outer-sphere re-
organization energy, implying the presence of a significant
inner-sphere reorganization energy, which was further con-
firmed by the differences in the solid Raman Au–S bond
stretching energies of Au250 and Au25− nanoclusters.
Maran and co-workers synthesized the Au25(S-C2H4Ph)18
with −1 and 0 charge states,93 and oxidized them (with bis
( pentafluorobenzoyl) peroxide) into Au25 with +1 charge
state.238 NMR results for these three nanoclusters nicely
matched the corresponding structures, which indicated the
presence of two different ligand populations in the capping
monolayer. In addition, the similar NMR patterns of Au25
nanoclusters with +1 and −1 charge states showed that the
[Au25(S-C2H4Ph)18]+ was formed as a diamagnetic species. In
Fig. 11 Comparison of Au25(SR)18 nanoclusters with different charge 2013, they synthesized the Au25 nanocluster with neutral
states. (a) Optical spectra and ESI mass spectra of Au25 with −1, 0, and charge state, and then studied its reduction or oxidation to a
+1 charge states. Reproduced from ref. 231 with permission from
series of charge states, −2, −1, +1, +2, and +3, under cyclic vol-
American Chemical Society, copyright 2007. (b) Comparison the UV-vis
optical absorption of Au25− and its oxidation product (Au250).
tammetry (CV).239 Also, the DPV results provided evidence for
Reproduced from ref. 232 with permission from American Chemical the different charge states of the Au25(SR)18 nanocluster, which
Society, copyright 2008. validated the existence of the −2, −1, 0, +1, +2, +3 charge

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10771
View Article Online

Review Nanoscale

hydrogen source to reduce the 2-nitrobenzonitrile into


2-amniobenzamide, and further return Au25 to the negative
state. It should be noted that the Au25 nanoclusters would not
structurally change in the recycling system, but just function
as an electron bridge to consistently transfer the electrons.
It is well-known that the physicochemical properties are
determined by the structures of the nanoclusters. Accordingly,
the Au25 nanoclusters with different charge states should have
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

different electronic structures as well as diverse properties. In


particular, Tofanelli et al. reported the Jahn–Teller effects
based on the crystal structures of Au25(SR)18 with −1, 0, +1
Fig. 12 DPV spectrum, providing evidence for the different charge charge states, in which the correlations between oxidation
states (−2, −1, 0, +1, +2, +3 charge states) of the Au25(SR)18 nanocluster. state, structure, and magnetism were described by first-order
Reproduced from ref. 239 with permission from American Chemical Jahn–Teller distortions.233,243 The noble gas-like configuration
Society, copyright 2013. (1S21P6) underlies the [Au25(SR)18]1− diamagnetism and com-
paratively high thermal stability. As for [Au25(SR)18]0 with the
1S21P5 superatom electron configuration, an unpaired 1p elec-
states of the Au25(SR)18 nanocluster (Fig. 12).239 Concerning tron is the origin of the paramagnetism and the stability of
the lifetime of the various oxidation states, the order is [Au25(SR)18]0 is not as good as [Au25(SR)18]1−. Larger distortions
Au25(SR)182− (4 ms) < Au25(SR)183+ (20 ms) < Au25(SR)182+ (0.33 s) were observed in [Au25(SR)18]1+ compared with Au25 nano-
(Fig. 12). Jin and co-workers also evidenced the different clusters with 0 or −1 charge states, caused by the unoccupied
charges of the Au25(SR)18 nanocluster in the DPV results and P orbital.233 Note that similar Jahn–Teller distortions were also
investigated the UV-vis absorption of these differently charged observed in single Pt and Pd doped Au25(SR)18 nanoclusters.244
Au25(SR)18 nanoclusters.274 Standard heterogeneous electron- Very recently, the Ackerson group compared the stability of
transfer (ET) rate constants, as well as the reorganization ener- [Au25(SR)18]1−/0/1+ nanoclusters with both shell-closing aspects
gies, were determined for each electrode process.239 It should and colloidal stability aspects.245 They proposed that the
be noted that the reduction to form Au252− and the oxidation decomposition of the nanocluster might follow either the
to form Au252+ and Au253+ are chemically irreversible. fusion or fission pathway, which could be determined by the
Significantly, they unambiguously pointed out that the Au25+ is solvent polarity.245
a diamagnetic species. In addition, the splitting of the HOMO
manifold in the Au25+ was severe, which suggested that the
usefulness of the superatom interpretation is limited to some 3 Atomically precise structure
extent.239 The stability of the Au25(SCnH2n+1)18 nanoclusters, in
which the number of carbon atoms varies from 2 to 12, with Despite the long period of research on the nanoparticles, the
different charge states (−2, −1, 0, +1, +2 charge states) has surface structures, including organic stabilizers and the in-
been investigated by Maran and co-workers by the use of CV organic–organic interfaces, remain poorly understood. On this
measurements.240 issue, without the information of the atomically precise
Accompanied by the single-electron redox reactions, the surface structures, the structure–property correlations cannot
phenylethanethiol-protected Au25−, Au250, and Au25+ nano- be obtained, not to mention the deep understanding of the
clusters can be reciprocally transformed with maintained con- origin of these properties. Moreover, the formation mecha-
struction. In 2012, we reported intermolecular electron trans- nism of nanoparticles with different shapes remains unclear
fer between 2,2,6,6-tetramethylpiperidin-1-oxoammonium without the accurate structures of the intermediate products
tetrafluoroborate (TEMPO+BF4−) and the as well as the final products. Metal nanoclusters, with the

[Au25(S-C2H4Ph)18] TOA nanocluster.241 Specifically, Au25−
+
advantages of monodisperse size and accurate structure, may
can work as an electron donor and be transformed into Au25+ be able to solve the aforementioned three fundamental issues
in the electron transfer reaction (occurs in a quantitative at the atomic level; the metallic kernel structures (i.e., the
manner with TEMPO+BF4−), as revealed by UV-vis, 1H NMR, arrangement of metals), as well as the metal–organic interfaces
and ESR analyses. In another work, Au25 nanoclusters with (including the arrangement of capped ligands and their
different charge states were employed as electron transfer cata- linking mode to the metallic kernel) should be determined. To
lysts to induce an intramolecular cascade reaction under obtain the atomically precise structures of such nanoclusters,
ambient conditions and gave rise to high conversion (87%) the most reliable approach is single-crystal X-ray diffraction
and selectivity (96%).242 Specifically, 2-nitrobenzonitrile is (SC-XRD).
capable of capturing an electron from Au25−, which forms a Since the crystal structure of Au25(SR)18 was obtained, there
free radical and further initiates the free radical reaction. has been an obviously notable leap in the scientific research
Electron–spin-resonance (ESR) spectra were employed to on the nanocluster. This includes the fundamental under-
capture the signals of free radicals. Excess NaBH4 provided the standing of the physicochemical properties correlating the

10772 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

structures and the formation mechanism from small metallic formed after approximately 7 days and appeared as dark
complexes to large nanoclusters, which highlights the impor- square crystals.
tance of the atomically precise structure in the nanomaterial In 2014, Maran et al. reported the synthesis and crystal
range. In this section, the approaches of crystallization, the structure of Au25(S-C2H5)18 in the neutral state.141 Black crys-
structural anatomies, and the structural comparisons of tals of this nanocluster were obtained by vapor diffusion of dii-
Au25(SR)18 nanoclusters with different capped ligands and sopropyl ether into a toluene solution of Au25(S-C2H5)18. Using
charge state are summarized. the same method, the crystal structure of Au25(S-nBu)18 was
also obtained.142
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

3.1 Approaches to obtaining the crystals of Au25(SR)18 We reported the crystal structure of [Au25(SePh)18]−(TOA)+,
nanoclusters which was obtained by the liquid diffusion method with the
A single crystal with high quality is the basis for determining CH2Cl2–ethanol system at room temperature;194 the crystals
the structure of a nanocluster with high reliability. To this were generated after 2–3 days. As for the Se@S or Te@S co-pro-
end, several approaches have been adopted. In 2008, tected Au25 nanocluster, Negishi and co-workers obtained the
the Murray group obtained the crystal structure of crystals using the gas phase diffusion method.195 Specifically,
[Au25(S-C2H4Ph)18]−(TOA)+.51 Firstly, they synthesized the 10 mg of the nanoclusters were dissolved in toluene (500 μL)
[Au25(S-C2H4Ph)18]−(TOA)+ with the modification of the Brust in an inner vial and ethanol was used as the anti-solvent in the
method (that is, using dichloromethane as a solvent and 1:1 outer vial. Needle-like crystals were obtained within 2–6 days.
thiol mixture of HS-C2H4Ph and HS-Ph).124 Of note, they found Very recently, Maran et al. reported an electrocrystallization
that the desired crystal would not form without HS-Ph. After strategy to prepare the crystals of nanoclusters in large quan-
purification, methanol was added to the crude mixture and tities and very high quality (Fig. 13).145 They found that the
the crystals of [Au25(S-C2H4Ph)18]−(TOA)+ were generated from electrocrystallization was specifically valid for obtaining the
the black oil after several days. crystals of Au25(SR)18 nanoclusters. With this strategy, they
At the same time, the Jin group also obtained the atomic- determined the atomic structures of [Au25(S-nC5H11)18]− and
ally precise structures of [Au25(S-C2H4Ph)18]−(TOA)+.52 It previously reported [Au25(S-C2H4Ph)18]0 nanoclusters.
should be noted that the crystals were generated from the pure
products in this work, relative to the aforementioned 3.2 Approaches to predicting the structures of Au25
one.51,52,92 Specifically, crystallization was performed by nanoclusters
adding 4 volumes of ethanol to 1 volume of a toluene solution
of the [Au25(S-C2H4Ph)18]−(TOA)+ nanocluster (20 mg mL−1 in Indeed, the most credible way to obtain the atomically precise
toluene), followed by centrifugation (∼5000 rpm for 5 min) to structure of a nanocluster is by analyzing the crystal data
remove the excess undissolved nanocluster. The saturated obtained from the SC-XRD. However, determination of the
supernatant was transferred to a vial, and the solution was atomic structure with a single crystal is a challenge, especially
allowed to stand on the bench under ambient conditions. in the early stage of the nanocluster research. In addition, it
After standing overnight, needle-like crystals (typically,
∼0.1 mm in diameter and 2–3 mm in length) were formed.
Finally, a fragment of one such crystal was used in X-ray struc-
tural determination and then the crystallographic data of
the [Au25(S-C2H4Ph)18]−(TOA)+ nanocluster was collected.
With this crystallization method, the crystal structure of
[Au25(S-C2H4Ph)18]0 was also obtained.232
In 2016, the Ackerson group successfully obtained the
atomically precise structure of the [Au25(S-C2H4Ph)18]+[PF6]−
nanocluster.233 It should be noted that the Au25+ was obtained
by bulk electrolysis from the twice-crystallized Au25−, and in
this process, the color of the solution was altered from orange
(or yellow) to green. The [Au25(S-C2H4Ph)18]+[PF6]− should be
crystallized shortly after the preparation, as the compound
appears to be unstable in solution for short periods of time.
The crystals can form in the saturated toluene/ethanol solution
by slow cooling.
The Ackerson group determined the crystal structure of
Au25(S-C2H4Ph)16( pBBT)2 in the neutral state (the synthetic
process has been discussed in section 2.1.4.).177 Crystals were
Fig. 13 Different shapes of crystals of [Au25(S-nC5H11)18]0, [Au25(S-
formed by dissolving Au25(S-C2H4Ph)16( pBBT)2 in toluene with nC4H9)18]0 or [Au25(S-C2H4Ph)18]0 nanoclusters covering the electrode,
ethanol added as the anti-solvent. The liquid diffusion system prepared by the electrocrystallization strategy. Reproduced from ref. 145
was then allowed to slowly crystallize at −20 °C. Crystals with permission from American Chemical Society, copyright 2017.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10773
View Article Online

Review Nanoscale

seems impossible to obtain the crystals of the nanoclusters researched the energetic preference of staple formation for the
capped by flexible ligands (GSH, for example), at least with the [Au25Cl18]− nanocluster by DFT-based global minimization.250
present technical means. In this context, other ways to predict Tlahuice-Flores et al. addressed how capped ligands modu-
the accurate structures of nanoclusters have also been lated the structures of a massive amount of the experimentally
performed. known anionic Au25(SR)18 nanoclusters.251 Specifically, low-
In 2007, Kojima and co-workers predicted the structure of polarity R groups do not significantly disturb the framework of
Au25(SG)18 through the 197Au Mössbauer spectroscopy.246 The Au25(SR)18, and the inversion symmetry (Ci) of the crystalline
proposed structure of Au25 has a Au7 metallic core capped by a state is retained. However, with the p-SPhX ligands, increasing
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Au12(SG)12 cage-like structure and two Au3(SG)3 rings. The distortion was observed in the framework of Au25( p-SPhX)18
core–shell construction gives rise to the high stability of the with X = H, Cl, NO2 and COOH. Accordingly, the inversion
Au25(SG)18 nanocluster against thiolate etching.246 symmetry of the Au25(SR)18 nanocluster was destroyed. In
Theoretical prediction is an effective strategy for obtaining addition, linking two –CH3 substituent groups at two positions
the optimized structures of nanoclusters. The combination of of HS-C2H4Ph ligands retained the framework as well as the
density functional theory (DFT) calculations and global optim- inversion symmetry; however, the –NH2 groups would destroy
ization algorithms has proven to be a powerful tool for predict- the symmetry. It is worth noting that the more distorted struc-
ing the construction of nanoclusters. This can be used for the tures resulted in the significantly reduced HOMO–LUMO gaps
determination of the optimized structures by comparing the and the optical absorption spectra were affected
computed photoelectron spectra with the experimental accordingly.251
spectra.108 In 2008 (the same time the crystal structure of It is interesting that from the MALDI-MS spectrum of the
Au25(SR)18 was reported), Akola et al. explored the optimized Au25(SR)18 nanocluster, the Au21(SR)14 fragment could be
structure of Au25(SR)18 nanocluster by DFT calculations.247 The identified (Fig. 14a). Liu et al. firstly calculated the optimal
proposed structure was based on a compact icosahedral Au13 structure of Au21(SR)14, and then predicted the fragmentation
kernel protected by six Au2(SR)3 staple motifs, which was iden- mechanism using the structures of Au25(SR)18 and
tical to the real structure of the experimentally determined Au21(SR)14.252 Finally, a stepwise fragmentation pathway was
Au25(SR)18.51,52 Compared with other predicted structures, the proposed (Fig. 14b), which contained (i) the transformation of
finally adopted structure exhibited obviously superior struc- Au25(SR)18, (ii) the detachment of [Au(SR)]x (x = 1–4), and (iii)
tural robustness and an eight-electron shell of delocalized the formation of Au21(SR)14 and a cyclic [Au(SR)]4.
Au(6s) electrons, which induced the high stability of this
structure.247
Aikens obtained the optimal structure of [Au25(SPh)18]− and
found that the [Au25(SPh)18]− geometric structure had S6 sym-
metry.248 In addition, the electronic structure of [Au25(SPh)18]−
was affected by a splitting of the superatom Pz orbital from the
set of Px and Py orbitals, which led to a double peak, similar to
the characteristic double peak of [Au25(S-C2H4Ph)18]−.
Furthermore, the structures of [Au25(SPhX)18]− (X = para-F, Cl,
Br, CH3, OCH3) were predicted. para substituents shifted the
HOMO and LUMO orbital energies, but the HOMO–LUMO gap
remained constant.248
In 2011, Dass and co-workers performed ligand exchange
on Au25(S-C2H4Ph)18 with HS-(CH2)n-SH, where n = 2, 3, 4, 5,
and 6.249 The most likely structures of the resulting bi-thiol
protected Au25 nanoclusters were predicted by DFT calcu-
lations. By analyzing the experimental and theoretical results,
the following were revealed: (i) Propanedithiol and butane-
dithiol had optimal chain lengths for interstaple binding to
the surface of the Au25 nanocluster, which induced more than
six interstaple bindings. (ii) Pentanedithiol and hexanedithiol
could also result in these outcomes to a lesser extent. (iii) The
chain length of ethanedithiol was too short for bidentate
binding.249
Jiang et al. investigated the possibility of replacing all the
Fig. 14 (a) MALDI-TOF-MS spectrum of the Au25(S-PhC2H4)18 nano-
thiolates in the Au25(SR)18 nanocluster with halogens with
cluster (data recorded by X. K.). (b) Sequential release of the [Au(SR)]x
maintained geometry and electronic structure.250 By the use of (x = 1–4) unit from Au25(SR)18−, theoretically predicted by DFT calculations.
the DFT calculations, they found that such halogen analogs of Reproduced from ref. 252 with permission from American Chemical
thiolated gold nanoclusters were highly likely. They also Society, copyright 2013.

10774 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

Furthermore, the continued fragmentation from Au21(SR)14 to traction effect of Au25(SePh)18 would induce changes in the
Au17(SR)10 was also observed, which had the same stepwise electronic properties of both the Au and Se in the
fragmentation mechanism as the fragmentation from nanocluster.259
Au25(SR)18 to Au21(SR)14.252 The Tsukuda group investigated bond stiffness in
X-ray absorption spectroscopy (XAS), including X-ray Au25(SR)18, Au38(SR)24, and Au144(SR)60 nanoclusters by EXAFS
absorption near edge structure (XANES) and extended X-ray spectroscopy.260 According to the EXAFS results, they revealed
absorption fine structure (EXAFS), are desirable tools for the the following hierarchy in the nanocluster structures: (i) The
structural prediction of Au nanoclusters and have previously long Au–Au bonds on the Au13 icosahedral core are more flex-
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

been found very useful for the study of a variety of ible than the bonds in the bulk Au. (ii) The short Au–Au bonds
nanomaterials.253–257 In 2011, Zhang and co-workers reported linking the Au in the kernel of icosahedron and the Au on the
the temperature- and solvation-dependent EXAFS study of the surface of icosahedron are stiffer than those in the bulk Au,
[Au25(S-C2H4Ph)18]− nanocluster.258 Firstly, they found that the and these bonds form a cyclic structural backbone with rigid
different peaks (or signals) in the EXAFS originated from Au– Au-SR oligomeric staple motifs.
Au or Au–S bonds centered at different sites or energies.
Furthermore, the structural changes experienced by the
Au25(SR)18 nanocluster in response to low temperature and
3.3 Comparison of the structures of Au25(SR)18 with different
different solvation environments were illustrated by the EXAFS
charge states and capped ligands
results. Moreover, they also pointed out the existence of the
interactions between the icosahedral Au13 kernel and six It is generally accepted that the small changes in composition
Au2(SR)3 staple motifs.258 In another work, they analyzed the will induce observable alternations in the configuration of
structural changes of the Au25(SePh)18 nanocluster at different nanomaterials. Understanding the composition dependent
temperatures.259 First of all, the Au L3-edge spectra and its structures in the nanocluster field will not only promote the
EXAFS fit of the Au25(SePh)18 nanocluster at different tempera- research on the origin of the properties, but the functional
tures were obtained (Fig. 15). Through analyzing the data, they nanoclusters can hopefully be controllably prepared using the
demonstrated that the icosahedral Au13 kernel of Au25(SePh)18 correlations/guidance. As for the Au25(SR)18 nanoclusters with
remained almost unchanged at low temperature (50 K) while the different capped ligands, notable differences in the struc-
aurophilic interactions on the surface were significantly longer tures have been observed. Meanwhile, different charge states
in distance compared with the thiolate-protected counterpart also have a significant impact on the structures of Au25(SR)18.
—Au25(SR)18. Furthermore, the Au–Au framework of The Au25(SR)18 nanocluster has an icosahedral Au13 core,
Au25(SePh)18 showed a significant contraction at the low temp- and this core is further capped by six pairs of Au2(SR)3 staple
erature, which was not found in the Au25(SR)18. Moreover, motifs to constitute the overall core–shell configuration
XANES was performed to demonstrate that the thermal con- (Fig. 16a–c). These six Au2(SR)3 staple motifs are all in the “V-
shape” semi-ring configuration (Fig. 16b). It should be noted
that the icosahedron possesses 20 triangular faces. Except for
the 12 faces on the icosahedral Au13 kernel capped by the

Fig. 15 XAS results of Au25(SR)18 and Au25(SePh)18 nanoclusters. Au L3- Fig. 16 Structural anatomy of the Au25(SR)18 nanocluster. (a)
edge (a) k-space spectra and (b) FT R-space spectra of Au25(S-R)18 and Icosahedral Au13 core. (b) Six pairs of Au2(SR)3 staple motifs. (c) Core–
Au25(Se-R)18 nanoclusters monitored at 50 K. (c) Au L3-edge and (d) Se shell structure of Au25(SR)18 with ball and stick style. (d) Core–shell
K-edge multi-shell EXAFS fit of Au25(SeR)18 nanocluster at 50 or 300 K. structure of Au25(SR)18 with spacefill mode. (e) The overall structure of
Reproduced from ref. 259 with permission from American Chemical Au25(SR)18. Color legend: green/violet/orange sphere, Au; red sphere, S;
Society, copyright 2014. grey sphere, C. For clarity, the hydrogen atoms are not shown.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10775
View Article Online

Review Nanoscale

staple motifs, there are also eight uncapped Au3 faces of the Ackerson and co-workers compared the structures of
icosahedron (Fig. 16d). Au25(S-C2H4Ph)18 with −1, 0 and +1 charge states in detail.233
Since the crystal structure of the [Au25(S-PhC2H4)18]+(PF6)− The continuous symmetry measurement (CSM) revealed that
nanocluster was reported,233 there have been enough models the icosahedral Au13 kernel in Au251− was more regular than
to analyse and compare the differences in atomically precise Au25 with 0 or +1 charge states, and the CSM values were
structures of Au25 with −0, 0, and +1 charge states. In 2008, 0.067, 0.201 and 0.524 (the value of the ideal icosahedron was
the Jin group analysed the differences in the structures of set to 0) for the nanocluster with −1, 0 and +1 charge states,
Au25(S-C2H4Ph)18 with −1 and 0 charge states.232 Note that respectively. As shown in Fig. 18a, the bond lengths on the
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

several differences have been discovered that are worth men-


tioning, which compare the structure of Au25 with 1- or 0
charge states, despite the close resemblance of these two nano-
clusters (Fig. 17). Firstly, no counterion matching the neutral
charge state of this nanocluster was found in the unit cell of
the crystal structure of Au250 (in comparison with Au25−).
Furthermore, the 12 Au atoms in the staple motifs are situated
around the three mutually perpendicular 2-fold axes of the ico-
sahedral Au13 core. In addition, apparent distortions have
been found in the structure of Au25− compared with Au250: (i)
Viewing along the σh plane in the x–y direction of Au25−, one
of the sulfur atoms in the Au2(SR)3 staple motif is bent down-
ward, while the inverted sulfur is bent upward. Similar distor-
tions have also been observed in the σh planes along the −z
and x–z directions in the Au25−. However for the neutral Au25,
the two Au2(SR)3 staple motifs are essentially coplanar with
the symmetry plane (σh) of the icosahedral Au13 core, regard-
less of being in the x–y, y–z and x–z planes; (ii) the ligand
orientations are totally different in the Au25− or Au250 nano-
clusters. It is suggested that these structural distortions of
Au25− come from the negative charge state, and might also be
because of the solid-state effect from the bulky TOA+ counter-
ion in the unit cell.232

Fig. 18 Structural/packing-mode comparison of Au25 with different


charge states or capped ligands. (a) A heat map of the edge lengths on
the icosahedral Au13 kernel of Au25(SR)181−/0/1+. Reproduced from ref.
233 with permission from The Royal Society of Chemistry, copyright
Fig. 17 Comparison of the crystal structures of Au25(S-PhC2H4)18 with 2016. (b, c) Self-assembled Au25(S-Bu)180 nanoclusters via single Au–Au
(a, c) anionic and (b, d) neutral charge states. Color legend: purple bonds to compose a linear polymer. (d) Un-assembled Au25(S-Et)180
sphere, Au; orange sphere, S. Reproduced from ref. 232 with permission nanoclusters. Reproduced from ref. 142 with permission from American
from American Chemical Society, copyright 2008. Chemical Society, copyright 2014.

10776 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

surface of the Au13 core vary over a range of 0.3, 0.4 and 0.7 Å
(also compared with the ideal icosahedron) for Au25 with −1, 0
and +1 charge states, respectively.233 As for the geometric dis-
tortions in the bonds of core-surface Au to the motif Au, the
CSM values were 3.407, 3.879, and 4.45, for 1, 0, and +1,
respectively. Consequently, the symmetry became lower as the
oxidation state was increased. Finally, the longest bonds (or
the weakest bonds) in the Au13 icosahedral core were 3 Å,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

3.1 Å, and 3.3 Å for Au25 with −1, 0, and +1 charge states,
respectively. These differences in the bond length (or bong
energy) might be linked to the thermal stability of the
Au25(SR)18 nanoclusters with different charge states.233
The Maran group found that the paramagnetic [Au25(S-
nBu)18]0 could self-assemble into a linear polymer of nano-
clusters connected via single Au–Au bonds and stabilized by
the proper orientation of nanoclusters as well as the interdigi-
tation of S-nBu ligands (Fig. 18b).142 Note that the inter-
molecular Au–Au bond distance is 3.152 Å, almost similar to
the average Au–Au bond distances of the core–shell Au—motif
Au (Fig. 18c). However, as for the Au25 capped by the S-Et or
S-C2H4Ph, no interdigitation was observed, which implies that
Scheme 6 Different types of Au–Au or Au–S bonds in the structure of
too short or too long ligands (with high steric hindrance) are the Au25 nanocluster.
not suitable for stabilizing the 1D polymer composed of nano-
clusters (see Fig. 18d for the crystal packing mode of the un-
assembled Au25 protected by S–Et ligands).142 [Au25(SR)18]− nanocluster (2.301 Å), which is due to the larger
Apart from the Au25(SR)18, the crystal structure of Au25 pro- covalent radius of the Se atom compared with the S atom (Se
tected by SePh was also obtained.194 SC-XRD results show that covalent radius of 1.20 Å versus S of 1.05 Å). However, the
the structure of [Au25(SePh)18]− indeed resembles that of average Au–Se–Au angle in the Au2(SePh)3 staple motif (94.89°)
[Au25(S-C2H4Ph)18]−.51,52,194 Apparent distortions have been is much smaller than the Au–S–Au angle (102.677°). In the icosa-
observed on the sulfur atoms in the Au2(SR)3 staple motif along hedral Au13 kernel, the average Au–Au bond distance is 2.797 Å
the σh plane when comparing Au25(SR)18 nanoclusters with in the [Au25(SePh)18]−, which is much longer than the average
different charge states.232 Interestingly, comparing the anionic bond length in the [Au25(S-C2H4Ph)18]− (2.775 Å). All in all, the
Au25 capped by S-PhC2H4 or Se-Ph ligands, we found that the icosahedral Au13 kernel exterior ligand shell as well as the
distortion of the Se atoms in the Au2(SePh)3 staple motifs within exterior ligand shell became more expanded due to the alterna-
the σh in the x–y, y–z and x–z planes was more obvious than tion of the capped ligands from SR to SeR. It has been suggested
those in Au2(SR)3 staple motifs (Fig. 19).194 In addition, the that the changes originated from the differences in the distri-
average bond length of Au–Se (2.434 Å) in the –Se–Au–Se–Au–Se– bution of the electron density and molecular orbital energies of
“V-shape” semi-ring is much longer than that in the Au25 nanoclusters protected by different capped ligands.194
Generally, the different capped ligands and charge states
have a significant impact on the structures of Au25(SR)18 nano-
clusters, such as bond length, bond angle, distortion degree
and the packing mode in the unit cell. As shown in Scheme 6,
there are mainly four types of Au–Au or Au–S bonds (a: core
Au—core surface Au; b: core surface Au—core surface Au; c: core
surface Au—motif Au; d: motif Au—motif S). The bond lengths
in each Au25(SR)18 nanocluster with different capped ligands
and charge states are summarized in Table 2 (corresponding
to Scheme 6; here we just summarize the Au25 structures deter-
mined by SC-XRD).

4 Structure-dependent properties
Fig. 19 Comparison of the crystal structures of [Au25(SePh)18]− (left)
The ultra-small size of nanoclusters (almost less than 2 nm in
and [Au25(S-PhC2H4)18]− (right) nanoclusters. Reproduced from ref. 194 diameter of the metallic core) places them in the gap between
with permission from The Royal Society of Chemistry, copyright 2014. the small metallic complexes and plasmonic nanoparticles. With

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10777
View Article Online

Review Nanoscale

Table 2 Bond lengths in each Au25 nanocluster (corresponding to the bond modes in Scheme 6)

Nanocluster Bond a (Å) Bond b (Å) Bond c (Å) Bond d (Å) Ref.
− +
[Au25(S-PhC2H4)18] (TOA) 2.759–2.789 2.789–2.974 3.010–3.275 2.300–2.333 51
Avg. 2.774 Avg. 2.918 Avg. 3.158 Avg. 2.311
[Au25(S-PhC2H4)18]−(TOA)+ 2.761–2.789 2.795–2.971 3.016–3.238 2.227–2.317 52
Avg. 2.775 Avg. 2.924 Avg. 3.157 Avg. 2.303
[Au25(S-C2H5)18]0 2.785–2.801 2.777–3.100 3.068–3.244 2.275–2.316 141
Avg. 2.793 Avg. 2.939 Avg. 3.154 Avg. 2.300
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

[Au25(S-nBu)18]0 2.771–2.795 2.771–3.061 3.040–3.233 2.292–2.312 142


Avg. 2.784 Avg. 2.929 Avg. 3.148 Avg. 2.304
[Au25(S-PhC2H4)18]0 2.773–2.796 2.762–3.204 3.071–3.234 2.245–2.319 145
Avg. 2.784 Avg. 2.934 Avg. 3.135 Avg. 2.296
[Au25(S-nC5H11)18]0 2.772–2.827 2.752–3.088 3.027–3.246 2.275–2.402 145
Avg. 2.789 Avg. 2.934 Avg. 3.142 Avg. 2.308
[Au25(S-PhC2H4)16(pBBT)2]0 2.779–2.801 2.770–3.098 3.063–3.222 2.254–2.322 177
Avg. 2.789 Avg. 2.936 Avg. 3.131 Avg. 2.293
[Au25(SePh)18]−(TOA)+ 2.779–2.811 2.833–3.040 3.033–3.231 2.399–2.493 194
Avg. 2.797 Avg. 2.942 Avg. 3.128 Avg. 2.431
[Au25(S-PhC2H4)18−x(SePh)x]− 2.763–2.792 2.793–2.976 3.012–3.245 2.305–2.334 195
Avg. 2.776 Avg. 2.920 Avg. 3.156 Avg. 2.319
[Au25(S-C3H7)18]0 2.776–2.792 2.774–3.040 3.072–3.206 2.291–2.310 376
Avg. 2.784 Avg. 2.928 Avg. 3.140 Avg. 2.301

strong quantum size effects as well as the discrete electronic started the gold colloid chemistry).261 When the size of Au
states, nanoclusters are capable of exhibiting a variety of fascinat- nanoparticles ranges from 5 nm to 20 nm of the metallic
ing physicochemical properties such as multiple absorption kernel, the surface plasmon resonance (SPR) bands appear in
bands, photoluminescence (PL), electrochemiluminescence, chir- the optical absorption at 520 nm. With the reduction of the
ality, nonlinear optical properties, ultrafast electron dynamics nanoparticle size, the absorption signal gradually blue-shifts
and magnetism. More importantly, slight changes in the struc- and disappears.27,262 Owing to the discrete energy levels of
ture (such as increasing/decreasing an atom, tailoring the small-sized nanoparticles (with size <2 nm, so-called nano-
capped ligands, or just changing the charge state) of the nano- clusters), other optical absorption signals emerge in the UV-
cluster result in significant alternations of properties, which visible region. Simultaneously, intriguing optical properties
empower the nanocluster with great potential for applications in such as luminescence and nonlinear properties have been
several fields. Additionally, compared with nanoparticles, nano- observed with these ultra-small nanoclusters. More impor-
clusters provide a practical platform for researching the struc- tantly, the crystal structures of these nanomaterials promote
ture–property correlations due to their monodisperse size and the research on the mechanism of these optical properties.
the atomically precise structures. In this context, for thoroughly The optical absorption, luminescence, and other optical pro-
understanding the origins and the mechanisms of these intri- perties are summarized and discussed in this section.
guing properties, fundamental research on the relationship 4.1.1 Optical absorption.. With the atomically precise
between the structures and properties is a prerequisite. structure of Au25(SR)18 and the corresponding optical absorp-
Regarding Au25, it has been mentioned above that the tion, DFT calculation is capable of researching the discrete
changes in charge states or the capped ligands can induce sig- energy levels of the nanocluster (that is, ascribing each absorp-
nificant differences in the structure and these differences will tion to the specific species of transitions in the electronic
definitely be represented in the physicochemical properties of structure). In 2008, for the first time, Aikens and co-workers
Au25 nanoclusters. So far, significant effort has been made in performed the time-dependent density functional theory
comparing, analysing, and tailoring the physicochemical pro- (TD-DFT) calculations on the structure of [Au25(SR)18]−(TOA)+
perties of Au25 nanoclusters. Also, several influential con- (Fig. 20).52 [Au25(SR)18]−(TOA)+ in the solution state exhibits
clusions have been proposed in view of the mechanisms of intense absorption at 400, 450 and 675 nm, and several
nanocluster properties. In this section, the physicochemical shoulder peaks at 350, 560 and 800 nm. The major absorp-
properties of Au25 nanoclusters are overviewed, including tions correspond to 1.82 eV (675 nm, Fig. 19a), 2.75 eV
optical, chiral, magnetic and electrochemical properties, etc. (450 nm, Fig. 20b), and 3.10 eV (400 nm, Fig. 20c) in the spec-
Furthermore, the correlations between structures (compo- trum on the energy scale, respectively.52 From the results of
sitions) and properties and how to tailor the properties with TD-DFT calculations, they found that both the sp and d bands
this relationship have also been covered. were quantized. Specifically, the UV-vis signal at 675 nm (1.82
eV, Fig. 20a) corresponds to the gap between the HOMO and
4.1 Optical properties of Au25 nanoclusters LUMO transitions, which is essentially an interband transition
The optical properties of gold nanoparticles have attracted che- (sp ← sp). As for the peak of 400 nm (3.10 eV, Fig. 19c), the
mists from the very early period of nano-science (since Faraday transition mode is an intraband transition (d ← sp), arising

10778 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 21 (a) Simulated optical absorption spectra of Au25(S-CH3)18− with


and without spin–orbit coupling. Reproduced from ref. 264 with per-
mission from American Chemical Society, copyright 2014. (b) TD-DFT
calculated optical absorptions of [Au25(SPhX)18]− nanoclusters.
Reproduced from ref. 248 with permission from American Chemical
Society, copyright 2010.

Aikens, in 2010, also predicted the geometric and electronic


structures of [Au25(SPhX)18]− (X = H, F, Cl, Br, CH3, and OCH3)
using the TD-DFT calculations.248 Firstly, six sets of three
π-stacked phenyl groups have been found in the structures of
[Au25(SPhX)18]− nanoclusters, which induced the idealized S6
Fig. 20 Kohn–Sham orbital level diagram and peak assignments of the symmetry of these nanoclusters. Furthermore, unlike the
absorption spectra of the [Au25(S-PhC2H4)18]− nanocluster. Reproduced optical absorption spectrum of [Au25(SH)]−, a double peak for
from ref. 52 with permission from American Chemical Society, copyright
the HOMO–LUMO transition as well as other differences in
2008.
the spectra of [Au25(SPhX)18]− nanoclusters were observed
because of the splitting of superatom P and D orbitals. More
importantly, the HOMO–LUMO gap almost remained constant
from the other occupied HOMO to the n orbitals (these orbi- despite the large shifts in the HOMO and LUMO orbital ener-
tals belong to the d band). The peak at 450 nm (2.75 eV, gies (Fig. 21b). This behavior has been previously observed in
Fig. 20b) is composed of the mixed types of transitions (con- electrochemical measurements.265
taining the interband sp ← sp transition and the intraband d Comparisons of the optical absorptions of
← sp transition).52 In a subsequent work, Aikens carried out Au25(S-C2H4Ph)18 with different charge states (mainly −1, 0, +1
more detailed analyses of the Au25(SR)18 system.263 charge states) have been previously reported.231–233,266 The
It should be noted that the 800 nm optical absorption is comparison of the optical absorptions of these three nano-
one of the important symbols of Au25(SR)18 with −1 charge clusters are exhibited in Fig. 22 and Table 3 (by dissolving
state. In 2014, Jiang et al. found the splitting of the supera- X-ray quality single crystals of each charge state of Au25 nano-
tomic 1P HOMO set by TD-DFT calculations, which was
caused by the spin–orbit interaction.264 Significantly, the split-
ting gives rise to the tail band at about 800 nm in the experi-
mental spectrum, which was missing in the previous DFT cal-
culations (Fig. 21a).264
It was suggested that the HOMO–LUMO transition was almost
solely composed of the atomic orbital contributions from the
icosahedral Au13 core. Considering that the optical absorption at
675 nm represents the HOMO–LUMO transition, this character-
istic peak is entirely induced by the transition from the electronic
and geometric structure of the icosahedral Au13 core (rather than
the 12 exterior Au atoms in the staple motifs). Structurally, pre-
vious work has demonstrated that the superatom-like structure of
the Au25(SR)18 nanocluster is due to the following: (i) the two-
lobe distribution resembling the atomic p orbital in the HOMO;
(ii) the d-like character in the LUMO.52,142,232,233 In this context,
the formal count of the valence electrons of 8e in the
[Au25(SR)18]− is well understood, also the 7e and 6e valence elec- Fig. 22 Absorption spectra of Au25(S-PhC2H4)18 nanoclusters with
trons of [Au25(SR)18]0 and [Au25(SR)18]+, respectively. charge states of −1, 0, and +1 (data recorded by X. K.).

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10779
View Article Online

Review Nanoscale

Table 3 Comparison of the main peaks in the optical absorptions of


Au25(S-PhC2H4)18 with charge states of −1, 0, and +1 (corresponding to
Fig. 22)

Nanoclusters Peak a (nm) Peak b (nm) Peak c (nm)

Au251− 405 445 675 & 780


Au250 400 455 630 & 685
Au251+ 385 470 660
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

clusters). The increasing HOMO–LUMO transitions were


observed from [Au25(S-C2H4Ph)18]− to [Au25(S-C2H4Ph)18]0,
then to [Au25(S-C2H4Ph)18]+. Relative to [Au25(S-C2H4Ph)18]−, a
slightly decreased energy gap (approximately 0.3 eV) was
observed for the 1P and 1De transitions in [Au25(S-C2H4Ph)18]0.
In contrast, the energy gap increased by about 0.1 eV in
[Au25(S-C2H4Ph)18]+ compared with the −1 charge state.233 The
decrease (corresponding to the Au25 going from −1 charge
state to the neutral charge state) occurred because one of the Fig. 23 (a, b) The temperature-dependent optical absorptions of the
Au25(S-C6H13)18 nanocluster. (c) Variation of the absorption maximum
1P orbitals increased in energy, but was still occupied by one
with temperature for Au25(S-C6H13)18. (d) Plots of oscillator strength
electron.52,233 The alternation of Au25 from charge state of −1 versus temperature for Au25(S-C6H13)18. Reproduced from ref. 267 with
to +1 would remove two electrons, resulting in a larger split- permission from American Chemical Society, copyright 2011.
ting of the 1P orbitals, and simultaneously, the highest energy
1P orbital would become unoccupied.233 In this context, the
increase in the transition gap (corresponding to the Au25 with Recently, the Xie group observed two abnormal characteristic
charge state going from −1 to +1) comes from the decrease in optical absorptions at 780 and 980 nm for Au25 protected by
the energy gap of the 1P to 1Dt, and further, resulting from the negatively charged thiolate ligands and then collocated with
splitting of the superatomic D orbital degeneracy. Ackerson positively or neutrally charged thiolate ligands (Fig. 24).269 They
et al. found that the average Au–S bond length between the
ligands and Au atoms on the surface of the icosahedral Au13
core was shorter in Au25 with the charge state of +1 compared
to those with charge states of 0 and −1 (Fig. 18a).233 Thus, the
increase in the energy gap for the ligand band to 1De in
[Au25(S-C2H4Ph)18]+ might arise from the electron density
attracting process of the electron deficient core from the
ligand shell. Theoretical optical absorption spectra have also
been obtained to validate the systematic blue shift of the first
absorption peak of Au25(SR)18 from the charge state of −1 to 0,
then to +1.233
Ramakrishna and co-workers monitored the temperature-
dependent optical absorptions of the Au25(S-C6H13)18 nano-
cluster from 303 K to 90 K.267 The absorptions of Au25(SR)18
with the appearance of the vibronic structure became sharper
and showed significant enhancement accompanied by the
decrease in the temperature (Fig. 23a and b). Furthermore, a
slight red-shift to high energy of the maximum absorption was
observed (Fig. 23c). Moreover, the oscillator strengths of the
transitions increased with the decrease in temperature
(Fig. 23d). The results have been explained on the basis of the
electron–phonon interactions involving the core-gold electrons
or excitons with semi-ring gold phonons.267
Sementa et al. investigated a ligand-enhanced optical
absorption on the Au25(SR)18 nanocluster via TD-DFT calcu-
lations.268 By tailoring the ligands’ steric hindrance and elec-
Fig. 24 Illustrations, UV-vis absorptions, and ESI mass spectra of the
tronic conjugating features, notable enhancements in optical mono/bi-thiolate protected Au25. Arrows in the UV-vis spectra indicate
absorptions and effective electron delocalization were simul- the novel absorptions at 780 and 980 nm. Reproduced from ref. 269
taneously achieved.268 with permission from The Royal Society of Chemistry, copyright 2016.

10780 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

proposed that the abnormal absorptions were induced by the


structural distortions of the Au25(SR)18, which originated from
the charge anisotropy on the nanocluster surface.269
We compared the optical absorption of Au25 nanoclusters
protected by SePh, S-C2H4Ph and Se-C2H4Ph.194 The optical
absorption of the Au25(Se-C2H4Ph)18 showed absorption peaks
at 400, 470 and 683 nm, similar to the Au25(S-C2H4Ph)18, with
just a little red-shift in the HOMO–LUMO peak (675 nm of
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Au25(S-C2H4Ph)18 versus 683 nm of Au25(Se-C2H4Ph)18). Thus,


the changes in anchor atoms (S and Se) with the same carbon
tails of the ligands would slightly affect the energy level posi-
tions and optical absorption peaks.194 In contrast, compared
with the optical absorption of [Au25(S-C2H4Ph)18]−, Au25 pro-
tected by SePh exhibited more intense peaks at 430, 500 and
725 nm. Accordingly, it is obvious that the HOMO–LUMO gap
of Au25(SePh)18 is much smaller than that of Au25(S-C2H4Ph)18.
Fig. 25 (a) Luminescence spectra of the Au25(SG)18 nanocluster in D2O,
In this context, the ligand conjugation effect is more impor- recorded using Si (dotted line) and Ge (dashed line) photodetectors.
tant than the S/Se stereo and electron effects in this system.194 Reproduced from ref. 206 with permission from American Chemical
Negishi and co-workers systematically performed the ligand Society, copyright 2002. (b) Left: Photographs of BSA (1) powder and (2)
exchange on the Au25(S-C2H4Ph)18 nanocluster with HSePh or aqueous solution, and BSA–Au nanocluster (3) and aqueous solution,
and (4) powder under visible or UV light. Right: Optical absorption
(TePh)2 ligands.195 The broadened and red-shifted optical
(dashed lines) and photoemission (solid lines) spectra of aqueous solu-
absorption was observed, accompanied by the increasing sub- tions of BSA (blue) and BSA–Au nanoclusters (red). Reproduced from ref.
stitution of the capped SR ligands by SeR or TeR ligands 214 with permission from American Chemical Society, copyright 2009.
(Fig. 6c and d).195 (c) Comparison of the fluorescence spectra of Au25(Capt)18 and
Compared with the oil-phase Au25(SR)18 nanocluster, water- Au25(SG)18 nanoclusters. Reproduced from ref. 218 with permission from
The Royal Society of Chemistry, copyright 2012. (d) Photoexcitation and
soluble Au25 nanoclusters capped by GSH, Cys, or MHA
emission spectra of compounds Au25(SG)18, Au25-MB, Au25-SGAN, and
ligands exhibited almost the same optical Au25-SGFN nanoclusters. Reproduced from ref. 220 with permission
absorptions.212,213,217–219 However, the Au25 nanocluster pro- from American Chemical Society, copyright 2008.
tected by BSA revealed almost no absorption in the UV-vis
spectrum, which might be due to the novel effect of the metal–
ligand interactions as well as the optical influence from the
large steric hindrance of the BSA ligands.214 Xie et al. developed a “green” method to synthesize the BSA
4.1.2 Photoluminescence (PL). PL, one of the most fasci- capped Au25 nanocluster with red emission.214 The obtained
nating properties of nanomaterials, has attracted wide BSA–Au25 nanocluster was highly stable in both the solution
research interest as a powerful tool in the chemical sensing state and in the solid state. MALDI-MS spectra provided evi-
and bioimaging fields.37,270–273 Among these fluorescent nano- dence for the molecular weight of 71 kDa of the nanocluster,
materials, nanoclusters have emerged as a prominent member which was much larger than the BSA ligand only (i.e.,
with the advantages of monodisperse size and the atomically 66 kDa).214 Significantly, the BSA–Au25 nanocluster displayed
accurate structure, and thus the mechanism of the PL could be strong luminescence at 640 nm, with PL QY of 6%
reasonably proposed. In this section, the PL of water-soluble (Fig. 25b).214 Raut and co-workers studied the steady state and
and organic-soluble Au25 nanoclusters are discussed time-resolved fluorescence properties of BSA–Au25, including
separately. polarization behavior in different solvents: glycerol, propylene
4.1.2.1 PL of water-soluble Au25 nanoclusters. In general, glycol and water.216 First of all, by a comparison of the emis-
water-soluble gold nanoclusters often exhibit relatively sions from the BSA–Au25 nanocluster and rhodamine B, they
stronger PL than their organic-soluble counterparts.274 In reconfirmed that the PL QY of BSA–Au25 is stronger than 6%
2002, Whetten and co-workers obtained the PL spectra of in water.216 In addition, they found that the fluorescence life-
the (SG)18 nanocluster (firstly defined as Au28(SG)16).206 A time of BSA–Au25 is long and heterogeneous with an average
broad PL spectrum extending from the visible to the infrared value of 1.84 μs. In glycerol at −20 °C, the anisotropy of BSA–
spectral range (2.0–0.8 eV) was observed, which contained Au25 is high, reaching a value of 0.35. Importantly, the exci-
two bands centered at 1.5 and 1.15 eV (Fig. 25a), indicating tation anisotropy strongly depends on the excitation wave-
that radiative transitions between the ground state and two lengths, which indicates a significant overlap of the different
distinctively different excited states have taken place. The transition moments. Furthermore, in water, the anisotropy
total PL quantum yield (QY) of Au25(SG)18 was relatively low decay reveals a correlation time below 0.2 μs; however, the
compared with other fluorescent nanomaterials (such as measured correlation time is longer when in propylene glycol,
lanthanide nanoparticles, organic dyes and quantum- and the initial anisotropy has been found to depend on the
dots).206,275–278 excitation wavelength.216

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10781
View Article Online

Review Nanoscale

In 2012, Jin and co-workers reported the one-pot synthesis ligands to the metal core.274 In this context, three strategies for
of the Au25(Capt)18 nanocluster.218 Compared with the enhancing the PL intensity based on Au25 nanoclusters were
Au25(SG)18 nanocluster, Au25(Capt)18 exhibited stronger emis- proposed: (i) increasing the electron donation capability of the
sion in the same luminescence region (centered at ∼700 nm, capped ligands; (ii) increasing the electropositivity of the
as shown in Fig. 25c).218 metallic core if the nanocluster core can sustain multiple
The Pradeep group exploited the ligand-exchange method charge states; (iii) employing ligands with electron-rich atoms
to functionalize the Au25(SG)18 nanocluster with MB, N-acetyl- and groups.274 For instance, when the GSH ligands on the
and N-formyl-glutathione (NAGSH and NFGSH, respectively) Au25 surface were ligand-exchanged with a long-chain peptide
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

ligands. 220 The PL spectra of Au25(SG)18, Au25-MB, Au25-SGAN nucleic acid (PNA, with many electron-rich N atoms and O
and Au25-SGFN nanoclusters were recorded at room tempera- atoms), a 1.8 times enhancement of the PL intensity was
ture (Fig. 25d). First of all, the excitation spectrum of the Au25- observed (Fig. 26c).274
MB nanocluster exhibited a single peak at about 527 nm (2.3 Using dithiols, the Wang group improved the QY of Au
eV), which showed a 20 nm blue shift compared with three nanoclusters.279–281 By reacting the Au25(S-C2H4Ph)18 with
other nanoclusters (double peaks at 2.15 and 2.30 eV). In dithiolate, the near IR luminescence of Au25 intensified;
addition, compared with the Au25(SG)18 nanocluster, the emis- however, in this process, the optical absorption diminished.280
sion peak of Au25-SGAN or Au25-SGFN was almost maintained; In the opposite process, a gradual decrease in the near IR
however, Au25-MB displayed a 30 nm blue shift. In this luminescence was observed when the Aun(dithiolate)m nano-
context, it appears that one can tune the photoluminescence clusters were ligand-exchanged with mono-thiolate (i.e.,
of water-soluble Au25 nanoclusters, though slightly, by choos- HS-C2H4Ph).280
ing proper water-soluble ligands. Wen et al. observed the quantum confined Stark effect in
4.1.2.2 PL of organic-soluble Au25 nanoclusters. The Jin BSA protected Au8 and Au25 nanoclusters.282 Specifically, the
group analyzed the ligand effect on the basis of atomically PL emission was red-shifted with increasing the pH of the
monodisperse Au25(SR)18 nanoclusters.274 First of all, the PL solution (Fig. 27a). A Stark shift plateau with 30 meV offset
intensities of Au25(S-C2H4Ph)18 with different charges are was observed (upon an increase in pH) in both absorption and
different, following the trend −1 < 0 < +1 < +2 (Fig. 26a). The fluorescence of BSA–Au25 nanoclusters, and the lifetime
surface ligands (-SR) play a major role in enhancing the fluo- measurements confirmed that the plateau of the BSA–Au25
rescence of Au25(SR)18 nanoclusters (Fig. 26b). There are two nanocluster was due to the screening effect of the semi-rings
different ways in which the surface ligands influence the fluo- in the structure.282 Two different Stark shifts of 79 and 52 meV
rescence: (i) affecting the charge transfer from the ligands to were exhibited in the dual fluorescent bands of the BSA–Au25
the metal nanocluster core (LMNCT) via Au–S bonds; (ii) tailor- nanocluster. More significantly, the Stark shifts in both BSA–
ing the direct donation of delocalized electrons of electron-
rich atoms or groups (on the substituent groups) of the

Fig. 27 PL emission or the intensity of Au25(SR)18 in solution with


different pH values or temperatures. (a) Fluorescence spectra of the
BSA-protected Au25 nanocluster at different pH values. Reproduced
from ref. 282 with permission from American Chemical Society, copy-
Fig. 26 (a) PL spectra of Au25(S-PhC2H4)18 nanoclusters with different right 2013. (b, c) PL intensity of Au25(S-PhC2H4)18 at different tempera-
charge states. (b) PL spectra of Au25 nanoclusters capped by different tures. (d) Temperature-dependent and time-resolved PL emission of
ligands. (c) Electron-rich ligand-induced PL enhancement based on the Au25(S-PhC2H4)18 obtained at several sample temperatures. Reproduced
Au25 nanocluster. Reproduced from ref. 274 with permission from from ref. 283 with permission from American Chemical Society, copy-
American Chemical Society, copyright 2010. right 2014.

10782 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

Au25 and BSA–Au8 had significant linear polar components Au25(SR)18 (R = H, CH3, CH2CH3 or CH2CH2CH3) upon photo-
due to their asymmetric structure; this could be potentially excitation using DFT and TD-DFT calculations.284 In the opti-
useful in probing local electric fields and, furthermore, can mized excited states, the large geometric relaxations (up to
also be used as a pH-sensor in the biological systems.282 0.33 Å) have remarkable impacts on the frontier orbitals of the
Green et al. monitored the PL intensity of the Au25(SR)18 Au25(SR)18 nanocluster.284 Specifically, the Stokes shift of
nanocluster accompanied by the temperature increase (from Au25(SH)18 is 0.49 eV. When capped by the ligands with longer
4.5 K to 200 K).283 Experimentally, the PL intensity of the Au25 carbon tails, the Au25(SR)18 exhibited the larger Stokes shift.
nanocluster increased sharply by 70% when the temperature Vibrational frequencies in the range 75–80 cm−1 were calcu-
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

increased from 4.5 to 45 K, and was then maintained from 45 lated for the nuclear motion involved in the excited-state
to 65 K, but the PL was finally quenched when the temperature nuclear relaxation, which is in great agreement with the
was raised above 65 K (Fig. 27b and c). The PL enhancement vibrational beating observed in time-resolved spectroscopy
from 4.5 to 45 K implies that a nanocluster with ultra-small experiments.284 Besides, [Au25(SH)18]− exhibited several excited
size might have negative expansion coefficients. The quench- states at about 0.8, 1.15, and 1.25 eV, which are theoretically
ing PL above 65 K originated from the coupling to low-fre- responsible for the emission observed experimentally from
quency vibrations associated with the ligand shell that passi- 1.15 to 1.55 eV. Only the core-based orbitals induced the
vated the nanoclusters. Additionally, in the low-temperature excited states; that is, charge-transfer states or other Au2(SR)3
region (from 4.5 to 40 K), the average PL decay rate of staple motif- or ligand-based states are not responsible for the
Au25(SR)18 increased sharply, accompanying the increasing excited states.284
sample temperature. However, the change in average decay Using the magneto-photoluminescence (MPL) spectroscopy,
appeared to be saturated at sample temperatures greater than Green et al. characterized the electronic relaxation dynamics
40 K (Fig. 27d).283 Two different vibrational modes were identi- and near-infrared emission of the neutral Au25(SR)18 nano-
fied: 200 cm−1 for Au(I)–S stretching and 90 cm−1 for Au(0)– cluster.285 Two main PL peaks at 1.78 and 1.98 eV were
Au(I) stretching. All of the PL spectra of the Au25 nanocluster detected, which were identified with the Lande g-factors of
were correlated to the relative branching ratios of the emission 1.05 ± 0.04 and 1.7 ± 0.1, respectively, and the latter value
components, confirming the decreased recombination matches the emission from a quartet state. Magnetic circular
emission associated with strong electron-vibration coupling polarized photoluminescence spectra of neutral Au25(SR)18
and high emission yields for low emission energies at low revealed that both the individual component peaks and the
temperature.283 global PL (sensitive to the magnitude) increased significantly
In 2014, we obtained the Au25 nanocluster protected by from 2.5 to 10 T, and then flattened out from 10 to 17.5 T.285
S-Nap, which exhibited enhanced PL relative to When the capping ligands of [Au25(SePh)18]− were changed
Au25(S-C2H4Ph)18 (about 6.5-fold enhancement, as shown in to S-C2H4Ph, an approximately 2-fold enhancement in the PL
Fig. 28a).139 Furthermore, band I at 740 nm and band II at intensity was observed.194 Considering that the fluorescence
680 nm were observed in the PL spectrum of the Au25(S-Nap)18 intensities of the nanoclusters are largely influenced by the
nanocluster. Because of the same framework of Au25 protected charge transfer from the surface ligands to the metallic kernel
by different ligands, different band II for Au25 protected by (LMCT process), we therefore propose that the charge transfer
S-C2H4Ph and S-Nap implied that the surface ligands played a across the S–Au bond is larger than that across the Se–Au
major role in the origin of the fluorescence. In this work, we bond. A 25 nm red-shift in the PL spectrum was observed,
also found that the PL intensities of Au25(SR)18 with different compared to the [Au25(SePh)18]− with [Au25(S-C2H4Ph)18], and
charges followed the order of −1 < 0 < +1 (Fig. 28b).139 this red-shift corresponded to the enlargement of the HOMO–
Considering that the mechanism of PL in the nanocluster LUMO transitions calculated from the optical absorption.194
remained unclear, Aikens and co-workers discussed the geo- Significantly, the Au25 nanocluster protected by BSA showed
metric and electronic structural changes of the anionic significantly higher PL compared with its organic-soluble
counterparts (that is, Au25(SR)18 nanoclusters).214 Xie et al.
reported the in situ synthesis of BSA–Au25 with red emissions
(λem max = 640 nm) and high QY (about 6%). It should also be
noted that the QY is relatively small (0.5%) when the reactive
temperature is 100 °C. Wen and co-workers subsequently
researched the fluorescence dynamics in BSA-protected Au25
nanoclusters by time-resolved PL and transient absorption
techniques covering the picosecond to microsecond scales.286
They demonstrated that the red PL of the BSA–Au25 nano-
cluster contained both the prompt fluorescence and thermally
activated delayed fluorescence, and the latter was the main
Fig. 28 (a) Comparison of the PL spectra of Au25 protected by
S-PhC2H4 OR S-Nap ligands. (b) PL spectra of Au25(S-Nap)18 nano-
contribution. Furthermore, the absorption band centered at
clusters with −1, 0, +1 charge states. Reproduced from ref. 139 with per- 2.34 eV corresponded to the HOMO–LUMO transition from
mission from The Royal Society of Chemistry, copyright 2014. the capped BSA ligands to the metallic core. An effective relax-

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10783
View Article Online

Review Nanoscale

ation pathway was shown, which existed from the higher


excited state to the LUMO state.286
4.1.3 Electrochemiluminescence (ECL). ECL, also called
electrogenerated chemiluminescence, is a process whereby
species generated at the electrode undergo a high-energy elec-
tron-transfer reaction to form an excited state that emits
light.287,288 Ding and co-workers employed ECL for the Au
nanoclusters and used the ECL signals to probe the organic
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

and inorganic species.289–292 In 2012, they investigated the


near-infrared (NIR) ECL of Au25(S-C2H4Ph)18 and observed the
NIR-ECL emission in both annihilation and co-reactant
paths.289 Specifically, spooling spectroscopy was employed
during the ECL evolution and devolution processes. Together
with the NIR-PL spectroscopy, fluorescence generation mecha-
nisms were proposed. The ECL centered at 893 nm mainly ori-
ginated from the electronic relaxation of the excited state to
the ground state of the [Au25(S-C2H4Ph)18]− nanocluster. In
addition, other NIR-PL emissions were also matched: (i) the
strong emission at 857 nm matched the HOMO–LUMO tran-
sition gap; (ii) intermediate emissions at 719 and 820 nm were
attributed to the excited states higher than the HOMO–LUMO
gap; (iii) the weak emission at 1080 nm matched the outer
semi-rings.289
Ding and co-workers found that the [Au25(S-C2H4Ph)18]−
nanocluster showed no appreciable ECL light emission under
annihilation conditions (where the nanoclusters were electro-
chemically pumped to their various oxidized and reduced
forms), and the presumable reason was the short lifetime of
Fig. 29 Spooling ECL spectra of 0.1 μM Au25 in the neutral state in the
the electrogenerated intermediates and the low activity with presence of (a) 12.5 mM, (b) 50.0 mM, and (c) 200.0 mM of TPrA and 0.1
Au252− and Au252+.290 However, with tri-n-propylamine (TPrA) M TBAP as the support electrolyte. The applied potentials are labelled
or benzoyl peroxide (BPO) as the co-reactant, relatively high near the spectra. Insets: ECL evolution (left) and devolution (right) in a
and long-lived emissions at 950 and 890 nm in the NIR region better format for clarifying the wavelengths of the ECL emission.
were observed, corresponding to the reduction and oxidation Reproduced from ref. 292 with permission from Wiley-VCH, copyright
2014.
intermediates electrogenerated from TPrA and BPO.
Interestingly, the local concentration of the excited states
determined the ECL intensity, and the intensity can also be
tailored by the applied potential and the concentration of co- Based on the BSA–Au25 nanocluster, Fang et al. investigated
reactants. Contrastively, the TPrA system was more efficient the ECL properties using triethylamine (TEA) as the co-reac-
than the BPO system and the Au25−/BPO system was more tant.293 This BSA–Au25/TEA system was applied in Pb2+ detec-
efficient than the Au25+/BPO system.290 Also, the NIR ECL tion. Zhu and co-workers discussed the possible ECL mecha-
intensity of the Au25+/TPrA system depending on the TPrA con- nism on the basis of the BSA–Au25 nanocluster.294 The
centration and working electrode potential has been systemati- effective electron transfer was from the conduction-band of
cally researched.291 the excited ITO (where ITO represents indium tin oxide) to
Ding and co-workers also constructed the Latimer-type BSA–Au25 nanoclusters. More importantly, the dopamine
diagram, which showed the relationship between the redox could be efficiently detected by this ECL system.294
potentials of the Au25 nanocluster ground and excited 4.1.4 Non-linear optical properties. Au nanoclusters in the
states.292 The spooling ECL spectra of Au250 in the presence quantum size regime (about 1–3 nm of the metallic kernel)
of TPrA with different concentrations were measured show size-dependent optical properties and functions, and
(Fig. 29). In the presence of TPrA as a co-reactant, the emis- therefore, nonlinear properties in this size regime are worthy
sions at 860, 865 and 960 nm in the NIR region were of investigation.295–298 The non-linear optical properties of
observed, which were emitted by Au25+*, Au250*, and Au25+*, nanoclusters include two-photon absorption (TPA), two-
respectively. The ECL efficiency was up to 103%, as high as photon fluorescence (TPF), and second- or third-harmonic
that of the classical [Ru(bpy)3]2+/TPrA system, and the ECL generation (SHG/THG). Regarding the TPA properties, the TPA
efficiency depended on the kinetics of the reactions between cross-sections of small thiolate-protected Au nanoclusters have
the Au25 with different charge states and the electrogenerated been investigated to be much larger than that of typical small
TPrA co-reactant.292 organic molecules; additionally, the TPA cross-sections per Au

10784 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

atom are also significantly larger than that of Au nano- cluster gap was 8 nm, strong PL intensity was observed with
particles. Due to these intriguing non-linear properties, nano- significant emission enhancement at 730 nm and moderate
clusters are prominent in the high-resolution multiphoton enhancement at 820 nm. Compared with the Au25 nanocluster
imaging and optical limiting applications.299–305 With the in- in solution, a notably larger TPA cross-section was determined
depth research on the non-linear optical characteristics of in the film state (106 GM in the film), which was correlated
atomically precise nanoclusters, the correlations between with the strong dipole coupling of the Au25 nanoclusters.301
structure and non-linear properties will hopefully be eluci- Recently, Goodson and co-workers optically interrogated
dated and there should be more optical applications. the mono-disperse Au25(SR)18 nanocluster utilizing two-
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Goodson and co-workers firstly revealed the noticeable TPA photon-excited fluorescence (TPEF) near-field scanning optical
properties of the Au25(SR)18 nanoclusters.300 The efficient TPA microscopy (NSOM).302 Significantly, a 5-fold enhancement in
of Au25(SR)18 exhibited an emission peak at 830 nm (Fig. 30a the resolution (i.e., 30 nm point resolution) of the aperture-
and b), with a cross-section of 2700 GM at 1290 nm (1 GM = based TPEF NSOM imaging was achieved. Besides, even being
10−50 cm4 s per photon). They observed TPA cross-sections in separated by several tens of nanometers, the Au25(SR)18 nano-
the NIR region of half-million to a few million GMs on the clusters can also be excited and interrogated. The enhanced
Au25(SR)18 nanoclusters, which hopefully made them efficient TPA cross-section, which was induced by the few-atom local
two-photon absorbers in several fields of application, such as field effect and local field-induced microscopic cascade, is
optical power limiting and nanolithography. Besides, with the potential for disease diagnostics, cancer cell therapy, mole-
increase in the nanocluster size, the cross-sections also cular computers, high-density data storage, etc.302
increased from 1.1 to 4.0 nm (Fig. 30c). However, the cross- Research on the non-linear properties from small-sized
section per gold atom decreased significantly (Fig. 30d).300 nanoclusters to large nanoclusters, then to the plasmonic
They also used the Au25(S-C6H13)18 nanocluster as a model to nanoparticles give us more opportunities for thoroughly ana-
investigate the linear and non-linear optical properties of the lyzing the evolution from small-sized molecules to relative
nanocluster films.301 The nanocluster films were fabricated by large nanoparticles. Philip et al. investigated the evolution of
loading the Au25(SR)18 nanoclusters on different carriers, such non-linear optical properties with the increasing of the nano-
as polyvinylpyrrolidone (PVP), polystyrene (PS), polyethylene cluster size from the non-plasmonic regime to the plasmonic
glycol (PEG), and so on. After exploration, the Au25@PS was regime.295 Strong optical power-limiting behavior was found
adopted as it produced the best-quality film that is suitable for on the Au25(SR)18, Au38(SR)24, and Au144(SR)60 nanoclusters.
optical measurements. Optical absorption illustrated that the More importantly, the Au144(SR)60 nanocluster is the boundary
intercluster distance was significantly enlarged when forming for the nonlinear transmission properties between the mole-
the films and no aggregation occurred in this process.301 cular and plasmonic (Fig. 31).295
Owing to the intercluster dipole coupling, the TPA properties The first hyperpolarizabilities of Au nanoclusters were
as well as the PL were enhanced. Significantly, when the inter- measured by Antoine and co-workers, using a hyper-Rayleigh
scattering technique.296 The first hyperpolarizability β for the

Fig. 30 (a) Two-photon emission of the Au25(SR)18 nanocluster. (b)


Power dependence suggests the two-photon emission of AU25(SR)18. (c) Fig. 31 Nonlinear transmission of (a) Au25(SR)18, (b) Au38(SR)24, (c)
Absolute TPA cross-sections of nanoclusters with different sizes. (d) TPA Au144(SR)60 and (d) Au nanocrystals with size of 4 nm (calculated from
cross section per Au atom in each nanocluster. Reproduced from ref. the z-scan data). Reproduced from ref. 295 with permission from
300 with permission from American Chemical Society, copyright 2008. American Chemical Society, copyright 2012.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10785
View Article Online

Review Nanoscale

Au25(SG)18 nanocluster was calculated as 128 × 10−30 esu, for analyzing the correlations of the ultrafast electron
which is much smaller than 509 × 10−30 esu of the Au15(SG)13 dynamics and structures of Au nanoclusters, such as femtose-
nanocluster. Besides, all of these values are significantly larger cond time-resolved fluorescence up-conversion and transient
than that of the Au nanoparticles with 10–50 nm size. That is, absorption.80,306–317 More importantly, unlike the strong
the smaller the size of the nanoclusters, the stronger the non- power-dependent electron–phonon coupling of the Au nano-
linear two-photon properties that exist.296 Knoppe et al. investi- particles, laser-power-independent electron dynamics was
gated the SHG and THG of Au nanoclusters composed of 25 or observed on the Au nanoclusters.27,80,207,306,318
38 gold atoms.303 They demonstrated that the SHG and THG The Whetten group investigated the femtosecond transient
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

of Au38(SR)24 were much stronger compared to the Au25 nano- absorption of Au nanoclusters composed of 28 gold atoms and
cluster. Specifically, the Au38(SR)24 nanocluster was active at GSH ligands (later corrected to Au25(SG)18).207 The excited-state
the wavelength of 800 nm, while almost no signal of SHG was relaxation of this nanocluster exhibited biexponential decay
found with regard to the Au25(SR)18 nanocluster. The weak with a sub-picosecond as well as a longer nanosecond decay
SHG/THG of Au25(SR)18 originated from the high symmetry of time, independent of the laser pump power. In contrast, the
its construction.303 excited-state relaxations of larger nanoparticles showed a
Due to the long wavelength emission and two-photon exci- much shorter decay, which demonstrated the molecule-like
tation of the BSA–Au25 nanocluster, it can be potentially used ultrafast dynamics behavior of the nanoclusters. Specifically,
as a probe for deep tissue imaging. Raut et al. investigated the the transient absorptions of the Au25(SG)18 nanoclusters were
two-photon luminescence behavior of the BSA–Au25 nano- composed of excited-state absorptions, which is different from
cluster, and found a quadratic relation between the excitation the strong bleaching due to SPR in the nanoparticles.27,207,306
power and emission intensity, whereas one-photon excitation Devadas et al. measured the ultrafast luminescence
showed a linear dependence.304 dynamics of Au25 nanoclusters capped by HS-C6H12 and GSH
It is now clear that the TPA cross-sections of small thiolate- and investigated the relaxation of core Au states to Au semi-
protected Au nanoclusters are much larger compared to typical ring states in the Au25 nanoclusters (Fig. 32a and b).318 The PL
small organic molecules, and the cross-sections per Au atom with low QY of the Au25 nanocluster was observed in addition
of Au nanoclusters are also significantly larger compared to Au to the NIR-PL, which was used as a probe to understand the
nanoparticles. One-photon double-resonance has been pre- excited-state relaxation of nanoclusters. The ultrafast relaxation
viously regarded as the primary cause of the high cross-sec- within the metal core was less than 200 fs, which is distinctly
tions of nanoclusters. However, because of the large system different from that of the strong core–shell coupling (1.2 ps);
size and high density of states of the nanoclusters, detailed therefore, the orbitals of icosahedral Au13 metal core and
analysis on the TPA of nanoclusters remains challenging. Very Au2(SR)3 staple motifs were separated.318 In another work from
recently, Jensen and co-workers theoretically investigated the Goodson and co-workers, the increase in emission efficiency
TPA properties of [Au25(SR)18]− based on a dampened response
theory formalism.305 As such, they proposed that the one- and
two-photon double-resonance effects were unlikely to be the
only reason for the high cross-sections of the Au25(SR)18 nano-
cluster because the calculated one- and two-photon double-
resonance effects were much smaller compared to the experi-
mentally obtained values.305 Furthermore, the symmetry
breaking of the Au25(SR)18 nanocluster slightly increased the
TPA cross-sections. In addition, the calculated TPA cross-sec-
tions per Au atom and the Kerr non-linear responses were in-
line with the proposed results from Au nanoparticles.305
4.1.5 Ultrafast electron dynamics. Metal nanoclusters
display physicochemical properties with molecule-like tran-
sitions that are different compared to larger nanoparticles. In
addition, the ultrafast relaxation processes of the nanoclusters
are not totally the same as the small molecules, making them
a new member of the organometallic light absorber class.27 In
recent years, research on the ultrafast electron dynamics based
on the atomically precise nanoclusters have been widely
studied because of their superiority regarding the understand- Fig. 32 Ultrafast luminescence decay and wavelength dependence of
ing of the electronic structure of Au nanoclusters.306–316 (a) Au25(S-C6H12)18 and (b) Au25(SG)18 nanoclusters. Reproduced from
ref. 318 with permission from American Chemical Society, copyright
Historically, ultrafast electron dynamics research based on the
2010. (c) Fluorescence lifetime comparisons for Au nanoclusters with
Au25(SR)18 nanoclusters mainly focus on the relaxation time- various sizes. (d) Transient absorption of Au25, Au55, and Au140 at 550 fs.
scales, radiative emission, and electron–phonon Reproduced from ref. 319 with permission from American Chemical
coupling.207,236,317–324 Several technologies have been adopted Society, copyright 2010.

10786 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

as well as the longer lifetimes of small-sized nanoclusters rela- S6 excited states, 2 orders of magnitude larger recovery time-
tive to the larger nanoparticles were determined (Fig. 32c and scales were found in the ground state. More importantly, the
d).319 time scales of the ground states recovery increased after the
Jin and co-workers compared the ultrafast relaxation correction of the energy gap.322
dynamics of Au25(S-C2H4Ph)18 with 0 and −1 charge states Stoll et al. investigated the superatom state-resolved
(Fig. 33).236 For Au25(SR)18 with different charge states, photo- dynamics of the anionic Au25(S-C2H4Ph)18 nanocluster and
excitation occurred in two non-degenerate states near the solved the superatom P and D states using femtosecond two-
HOMO–LUMO gap originating from the core orbitals. dimensional electronic spectroscopy.323 Hot electrons therma-
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

However, the lifetimes of the core excitations of these two lized within the manifold of the dxy superatom D states within
nanoclusters showed notable differences. The decay rate of about 200 fs have been observed. However, hot holes formed
Au25(SR)18 with reductive state was over 1000 times slower than by visible excitation showed slower electronic relaxation (about
that of the nanocluster with neutral state. They proposed that 290 fs) within the P states. In addition, the Au25(SR)18 transient
the differences mainly originated from three aspects: (i) absorption spectra mainly consisted of the excited state
different absolute energies of the transitions; (ii) enhanced absorption from the D state to higher energy states. Besides,
symmetry of Au25(SR)18 in the neutral state; (iii) the presence the 200 fs (corresponding to the electron internal conversion
of a counterion in the [Au25(SR)18]− nanocluster.236 Thomas process) resulted in an energetic blue shift in the excited state
and Knappenberger also discussed the differences in the relax- absorption signal and the delayed detection of the transient
ation dynamics of Au25 nanoclusters with different charge ground state bleaching of transitions from the HOMO−1 and
states.320 They also found that the NIR excited state absorption HOMO−2 states.323
measurements were instrumental in describing the relaxation 4.1.6 Other spectroscopies. Farrag et al. investigated the
processes of metal nanoclusters.320 infra-red (IR) spectroscopy of the dimension-growth of
Kamat and Stamplecoskie found that the Au25(SG)18 with Au25(SR)18, Au38(SR)24 and Au144(SR)60 nanoclusters (where SR
core–shell configuration displayed rapid (<1 ps) as well as represents S-C2H4Ph).325 A distinct shift in the aromatic C–H
slower relaxation (∼200 ns); however, only the slower relaxation stretching band from 3030–3100 cm−1 to below 3000 cm−1 was
was observed in the homoleptic Au10–12(SG)10–12 nano- observed in the case of the Au25(SR)18 nanocluster; however,
clusters.321 More importantly, they demonstrated that the no shifts were observed in the case of Au38(SR)24 and
excited-state behavior was dominated by the ligand-to-metal Au144(SR)60 nanoclusters.325 In other words, the C–H stretch-
charge transfer (LMCT) process.321 ing signals of the two larger nanoclusters are similar to the
Aikens and co-workers theoretically investigated the non- bare ligand. The IR shift in Au25(SR)18 was proposed to have
radiative relaxation dynamics of the anionic Au25(SH)18 nano- originated from the electronic interaction of the aromatic
cluster.322 Compared with the relaxation time scales of the S1– rings on the ligands with other aromatic rings or with the gold
kernel.325
The Raman spectra enabled the Au–S bond to be character-
ized below 400 cm−1. Bürgi and co-workers measured the
Raman spectra of [Au25(SR)18]0/−, Au38(SR)24, Au40(SR)24, and
Au144(SR)60 nanoclusters (where SR represents S-C2H4Ph) for
investigating the size-effect of Au–S bonds as well as the con-
structions of monomeric and dimeric staples.326 It was
suggested that the types and numbers of Au–S bonding units
make a significant impact on the Raman spectra. The bending
vibrations centered at 195 cm−1 for the monomeric staples
were shifted to 215 cm−1 for the dimeric staples. When capped
with the mixed ligands, the Raman signal was between these
two values. Furthermore, exchanging two monothiols for a
bithiol would completely change the Raman spectra, as the
outer staple motif modes would be changed in the process.326
Maran and co-workers experimentally and theoretically
studied the Raman vibrational spectra of Au25(SnC2n+1)18 nano-
clusters (where n = 2, 3, 4, 5, 6, 8, 10, 12, 14).327 Raman spectra
at low-frequency field revealed the flexible behavior of the
shorter chains on the nanocluster surface, while the longer
chains tended to extend into all-trans conformations, which
Fig. 33 Global analysis of transient absorption spectra of (a) Au25− and
were recorded in the relatively high-frequency field (Fig. 34).327
(b) Au250 excited at 390 nm. Transient absorption spectra of (c) Au25−
and (B) Au250 as a function of excitation wavelength and time.
Among these nanoclusters, the Au25(S-C5H11)18 nanocluster
Reproduced from ref. 236 with permission from American Chemical showed the least obvious signal in the low-frequency field due
Society, copyright 2010. to the minimal coupling state.327

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10787
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 35 (a) Au 4f spectra of the Au25(SR)18, Au38(SR)24, and Au144(SR)60


Fig. 34 (a) Raman vibrational spectra of Au25(SnC2n+1)18 nanoclusters nanoclusters as well as Au nanoparticles (compound spectra in black
(where n = 2, 3, 4, 5, 6, 8, 10, 12, 14). (b) Theoretical Raman spectra for and divided core or shell spectra in red and blue, respectively). (b)
the simplified model system Au5(SCnH2n)3 (n = 2–6). (c) Comparison of Increasing gap between Au 5d3/2 and Au 5d5/2 accompanied by the
experiment and theory for the band in the region of 300–330 cm−1. growth in size. Reproduced from ref. 330 with permission from
Reproduced from ref. 327 with permission from American Chemical American Chemical Society, copyright 2013.
Society, copyright 2016.

DFT calculations have demonstrated that the closed elec- nanocluster field can be mainly classified into the following
tron configuration ((1S)2(1P)6) of the Au25(SR)18 nanocluster, three classes: (i) asymmetric arrangements of the capped
that is, the Au25 nanocluster, can be regarded as a prototypical staple motifs (or the asymmetric assembly modes between the
superatom.247,328 Very recently, Tsukuda and co-workers kernel and the outer atoms); (ii) chiral arrangements in the
reported the photoelectron spectrum of the anionic Au25(SR)18 metallic kernels; (iii) chiral carbon tails on the protected
nanocluster under vacuum.329 The energy levels and densities ligands. For preparing the Au25(SR)18 nanocluster with chiral
of the occupied states were directly recorded for the first time. properties, only the last strategy (i.e., using the chiral ligands)
The photoelectron spectrum of Au25(SR)18 exhibited two peaks, can be adopted, due to the highly symmetrical structure of the
which correspond to the electron detachment from the nano- nanocluster. In this section, the previous reports of the chiral
cluster 1P orbitals and Au 5d orbitals of the icosahedral Au13 Au25 nanoclusters induced by the chiral ligands are
core.329 summarized.
Ohta et al. investigated the size- and structure-effects of the It is worth noting that the water-soluble GSH is a chiral
electronic states based on the X-ray photoemission spec- ligand. In the early state of the thiolated nanocluster, Whetten
troscopy (XPS) results of Au25(SR)18, Au38(SR)24, and and co-workers observed the chiroptical response (circular
Au144(SR)60 nanoclusters (where SR represents S-C12H25).330 dichroism, CD) in the Au-SG nanoclusters with formula
Accompanied by the increase in the nanocluster size, the gap weights of 4.3, 5.6, 8.2, and 10.4 kDa.161,162 However, consider-
between Au 5d3/2 and Au 5d5/2 gradually increased from the ing that the framework of the Au25 nanocluster has a symmetry
1.84 eV of Au25(SR)18, to 1.88 eV of Au38(SR)24, 2.44 eV of plane, the chiroptical response of Au25(SG)18 (also of other
Au144(SR)60, and 2.80 eV of bulk Au (Fig. 35).330 GSH protected nanoclusters) might only come from the chiral-
ligand induction. Jin and co-workers proved the achiral prop-
erty of the metallic kernel of the Au25 nanocluster.344
4.2 Chiral properties of Au25 nanoclusters Specifically, NMR was performed on the Au38(SR)24 nano-
Chirality is a universal phenomenon in nature, which refers to cluster, and different 1H signals were observed for the two
a scenario in which an object and its mirror-symmetrical geminal protons in each –CH2 on the thiolate ligands. A large
counterpart cannot be perfectly superimposed. Chirality also chemical shift (about 0.8 ppm) of R–CH2 (close to the metal
plays many important roles in scientific research, such as drug core) indicated the chiral property of the metal kernel.344
development, chiral catalysis, to name a few. Although chirality However, no chemical shift was detected in the NMR of the
has been extensively researched in the field of organic chem- Au25(SR)18 nanocluster, confirming the achiral property of the
istry, chirality in nanoscience is indeed an emerging subject. metal core in the Au25 nanocluster. Significantly, using this
Research on chiral nanoclusters has received widespread inter- method can easily identify the chirality of the nanocluster,
est because the atomically precise structure of these nano- even if the nanocluster is a racemic mixture whose chirality
materials enables the true mechanism of the cannot be determined by the CD spectroscopy (Fig. 36). In this
chirality.162,331–343 So far, the origins of the chirality in the context, they proposed that NMR spectroscopy could be a

10788 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 37 CD spectra of chiral Au25 nanoclusters. (a) Chiral Au25 nano-


clusters protected by chiral PET ligands. Reproduced from ref. 137 with
Fig. 36 Diastereotopicity in the CH2 protons of the ligands on the (a, b) permission from American Chemical Society, copyright 2011. (b) CD
chiral nanoclusters or (c, d) non-chiral nanoclusters. Reproduced from spectra of Au25 protected by different chiral ligands. Reproduced from
ref. 344 with permission from American Chemical Society, copyright ref. 218 with permission from The Royal Society of Chemistry, copyright
2011. 2012. (c) CD spectra of Au25 and its R/S-BINAS ligand-exchange pro-
ducts. Reproduced from ref. 172 with permission from American
Chemical Society, copyright 2009.

simple but useful tool for detecting the chirality of the metal Ligand-exchange is a facile method for introducing a chiral
kernels of nanoclusters based on the diastereotopicity ligand into achiral nanoclusters, thus allowing the obtained
effect.344 nanocluster to exhibit a chiroptical response. Bürgi and co-
Garzón and co-workers theoretically investigated the origins workers used ligand exchange to convert the achiral Au25(SR)18
of the chiroptical response of Au25 protected by the SG using R/S-BINAS and NILC/NIDC to obtain chiral Au25 nano-
ligand.345 The optical activity of the nanocluster originated clusters.172,173 The BINAS-Au25 nanocluster showed chiroptical
from the slight structural distortion of the core–shell configur- responses at 300, 350, 400 and 500 nm (Fig. 37c). In addition,
ation/interaction and the chiral induction of the Rcys and Rmeth the NILC/NIDC-Au25 nanocluster showed chiroptical responses
groups on the ligands.345 at 320, 360, and 410 nm.172
By introducing different chiral ligands, the chiroptical Yao and co-workers reported the magnetic circular dichro-
response of the Au25 nanocluster can be controlled. In 2011, ism (MCD) spectra of Au25(S-C2H4Ph)18 and Au25(SG)18 nano-
we designed a pair of chiral thiols (R- and S-isomers) based on clusters (Fig. 38).346 The MCD signals of these two nano-
the PET ligand.137 The obtained Au25(PET*)18 (see the synth- clusters were all at 290, 350, 450 and 650 nm (Fig. 38c and d).
eses section) revealed obvious CD responses at 275, 325, 375, The MCD responses of the Au25 nanoclusters reveal the strict
430 and 490 nm (Fig. 37a). Subsequently, Jin and co-workers “nondegeneracies” of the ground and excited states, corres-
synthesized the water-soluble Au25 nanocluster with chiral ponding to the superatom P and D orbitals in the electronic
Capt ligands.218 The chiral ligands (i.e., PET*, Capt, and SG) transitions, respectively.346 In another work, the weaker CD
gave rise to distinct chiral features (Fig. 37b). These studies all response of the silver-doped Au25−xAgx(SG)18 ( preferentially
reveal the major role of ligands in the CD responses of nano- doping to the inner core) was observed, compared with the
clusters. The effects of different scenarios that induce the chir- homo-gold Au25(SG)18 nanocluster.347 The reason for the
optical response from Au25 nanoclusters have been analysed, decreased chiroptical responses was proposed to be the
including (i) chiral solvent induction, (ii) chiral counter-ion increased geometrical isomers of the bi-metallic nanoclusters.
induction, and (iii) chiral ligand induction.138 However, no These increasing numbers of the possible configurations
chiroptical activity in the achiral Au25 nanocluster was found buffer the CD responses and average the positive and negative
when the chiral solvent or counter-ion was introduced. That is, bands of different optical isomers.347 Other works by this
only the chiral ligand could induce the chiroptical response of group have provided more insight into the preparation of
Au25(SR)18 nanoclusters.138 chiral nanoclusters as well as the chirality regulation.348,349

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10789
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 38 Absorption spectra of Au25 nanoclusters protected by (a) SG or


(b) S-PhC2H4 ligands. MCD spectra of Au25 nanoclusters protected by (c)
SG or (d) S-PhC2H4 ligands. Reproduced from ref. 346 with permission
from American Chemical Society, copyright 2012.

4.3 Electrochemical properties of Au25 nanoclusters


Due to the highly purified and mono-disperse samples of
nanoclusters consisting of the metallic cores stabilized by
monolayer ligands that prevent the agglomeration of the Fig. 39 Electrochemical properties of the Au25(S-PhC2H4)18 nano-
metallic cores in solutions, tremendous developments have cluster. (a) DPV curves at 25 °C. (b) DPV curves at −70 °C. (c) CV curves
at −70 °C. Arrow indicates solution rest potential, and “*” indicates the
been achieved based on the electrochemical observations of
signal for incompletely removed O2. Reproduced from ref. 358 with per-
the nanoclusters.95,350–357 Electrochemistry reveals the ener- mission from American Chemical Society, copyright 2004.
gies of the HOMO and LUMO (associated with the energy gap)
of metal nanoclusters.356 The HOMO–LUMO transition gap
(defined as the difference between the first oxidation and
reduction potentials) indicates the energy levels and the redox Ox3–Ox2, Ox2–Ox1, and Ox1–Re1 transitions, respectively; note
potential of the nanoclusters, and thus it is closely related to that the last one (1.6 V) represents the HOMO–LUMO gap tran-
the catalytic and photovoltaic applications. CV, DPV and sition. The same results were obtained from the CV spectra.
square-wave voltammetry (SWV) are the three most frequently The solution- or temperature-dependent electrochemical pro-
used means for investigating the electrochemical properties as perties were also investigated.358
well as the HOMO–LUMO gaps of metal nanoclusters. The Murray group also investigated electron transport on
In the early stages of protecting oil-soluble nanoclusters by the basis of the Au25 nanocluster melt, which occurred in the
using thiols, Murray, Lee and other research groups studied electron self-exchange reactions, or electron hopping between
the electrochemical properties of the Au25(S-C2H4Ph)18 the neutral Au25 and anionic Au25 nanoclusters.359 Similar
nanocluster.358–361 In 2004, Lee et al. reported the DPV and CV rates of electron and electrolyte ion transport have been
spectra of the Au25(S-C2H4Ph)18 nanocluster (Fig. 39).358 The observed. Furthermore, the electron transfer rates of the
electrochemical properties of the Au25 nanocluster dissolved in Au25(SR)18 nanocluster places it between the metal complex
the solution mixture of toluene/acetonitrile or dichloro- polyether melts and Au nanoparticle melts.359
methane were determined under the following conditions: (i) Tofanelli et al. prepared Au25(SR)18 with −1, 0, and +1
the addition of Bu4NPF6 (as an electrolyte); (ii) 0.4 mm dia- charge states with the electrochemical method.381 Specifically,
meter Pt working electrode; (iii) Ag wire quasi-reference elec- the preparation of each formal charge state of the Au25(SR)18
trode (Ag QRE); (iv) Pt wire counter electrode.358 On the basis nanocluster proceeded by the initial collection of a DPV and
of the anionic Au25(SR)18 nanocluster dissolved in the CH2Cl2 verification that the as-prepared nanoclusters showed the
at 25 °C, three main oxidized peaks at potentials of 0.1, 0.4 expected electrochemical response. Following the DPV
and 1.1 V were observed, corresponding to the Au250, Au251+ measurement, a small amount of the Au25(SR)18 nanocluster
and Au252+, respectively. In addition, only one signal with in each targeted formal charge state was prepared by bulk elec-
reductive state was found at −1.5 V, which corresponds to the trolysis.381 Bulk electrolysis was executed with two different
Au25(SR)18 with the −2 charge state. Accordingly, the formal electrolytes (TBAPF6 and TEABF4) to isolate the stability effect
potential spacings were calculated as 0.7, 0.3 and 1.6 V for the of the Au25(SR)18 nanocluster core charge from the effect of

10790 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

counterions. Because of the instability of the Au25(SR)18 nano-


cluster with charge states of −2 and +2, the stated nanoclusters
were not obtained.381
Electron-transfer dynamics enable deep insight into the col-
lective properties of nanoclusters. Lee and Kim, in 2006, quan-
tified the electron-transfer dynamics by in situ voltammetry of
the Au25(SR)18 nanocluster.360 When the Au25(SR)18 nano-
cluster surface was fabricated with various lengths of dithiols
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

(such as C5-dithiol, C6-dithiol, C8-dithiol, and C9-dithiol), the


inter-nanocluster distances were 8.0, 9.5, 12.0, and 13.3 Å,
respectively (by DFT calculations).360 The corresponding
current intensity in the SWV spectra gradually weakened
accompanied by the increase in the internanocluster dis-
tance.360 Pineda and co-workers measured the DPV and CV
spectra of the Au25 nanocluster protected by hexanethiolate.361
The HOMO–LUMO gap calculated from the electrochemical
measurements was 1.33 V, similar to the gap of Au25 protected
by S-C2H4Ph. Furthermore, the electron transfer rate constants
of the Ox1 and Ox2 of this nanocluster were 0.017 and
0.021 cm s−1, respectively, determined by the oxidation cur-
rents as well as by digital simulation from the CV spectra.361
Lee and co-workers modified the sol–gel electrode with the
anionic Au25(S-C6H13)18, and the obtained electrode exhibited
excellently mediated electrocatalytic activity, which could be
utilized for amperometric sensing of the ascorbic acid and
uric acid.362 Specifically, in the electrochemical environment,
the neutral Au25 nanocluster on the immobilized electrode was
firstly reduced to the anion Au25, and the analyte is capable of
Fig. 40 (a, b) CV spectra of the Au25 nanocluster with different concen-
oxidizing the anion nanocluster into the neutral counterpart.
trations of ascorbic acid or uric acid. (c, d) Calibration graphs for the
The current of the CV was linear with the concentration of the determination of the concentration of ascorbic acid or uric acid from
ascorbic acid and uric acid, which posed a relatively con- the voltammograms of the Au25 nanocluster. Reproduced from ref. 362
venient amperometric sensing strategy (Fig. 40a–d).362 They with permission from American Chemical Society, copyright 2011. (e, f )
also reported the ascorbic acid and dopamine sensing systems SWVs of PT-Au25 and Au25(S-C6H12)18 in CH2Cl2/toluene mixtures (v : v
in the figures). Reproduced from ref. 364 with permission from
composed by the Au25(SG)18 modified electrode.363 In 2012,
American Chemical Society, copyright 2012.
they characterized the electrochemical properties of both
organic- and water-soluble Au25 nanoclusters.364 The water-
soluble Au25 nanocluster protected by (3-mercaptopropyl)sulfo-
nate is capable of transferring into the organic phase by ion- CV features. The lifetimes of the −2, +3, +2 charged nano-
pairing the sulfonate terminal groups of the capped ligand cluster were 4 ms, 20 ms and 0.33 s, respectively, validating the
with hydrophobic counterions. The phase-transferred Au25 unstable properties of Au25 with these charge states, since the
(PT-Au25) nanocluster was stable in the organic solvent. time of the electrochemical experiment was much longer than
Compared to the DPV results of Au25 and PT-Au25, the poten- the lifetimes of the corresponding nanoclusters.239 In this
tial of the oxidized peaks was maintained; however, the context, the other properties (for example, the optical pro-
reduced peak was red-shifted from −2.15 V of water-soluble perties) of Au25 with these charge states have not yet been
Au25 to −1.85 V of PT-Au25 (Fig. 40e and f ). In addition, the obtained.239 However, homogeneous redox catalysis based on
polarity of the solvent had a significant impact on the SWV the Au25 with these charge states could be performed.365
spectra of PT-Au25, while no changes were found from the The Maran group obtained the crystals of the Au25(SR)18
SWV of the water-soluble Au25 nanocluster.364 nanocluster using the method of electrocrystallization.145 First
Maran and co-workers determined the standard hetero- of all, the solvent effect and the ligand effect on the electro-
geneous electron-transfer rate constants and reorganization chemical properties of the Au25(SR)18 nanocluster were
energies for Au25 with different charge states.239 Specifically, researched.145 As shown in Fig. 41a, large differences were
the CV and DPV were performed on the neutral detected by comparing the CV curves of the [Au25(S-n-C4H9)18]−
Au25(S-C2H4Ph)18. Except for the Au25 nanocluster with −1, 0, nanocluster in different solvents (MeCN and DCM with
+1 charge states, the −2, +2, and +3 charged Au25 nanoclusters different proportions). In detail, mild CV peaks were detected
were observed.239 However, the unusually charged Au25 nano- by increasing the proportion of DCM in the solvent mixture. In
clusters are not stable, which is manifested in the irreversible addition, different CV curves of Au25 nanoclusters protected by

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10791
View Article Online

Review Nanoscale

crete Au atoms and larger Au nanoparticles. Magnetic research


in the nanocluster field will hopefully reveal the origin of the
magnetism as well as the evolution from paramagnetic Au
atoms to diamagnetic Au nanoparticles.368–373
DFT calculations have demonstrated the 8 e− system of
anionic Au25(SR)18, which exhibits a closed electron configur-
ation (that is, a noble gas-like (1S)2(1P)6, superatom-like elec-
tron configuration).247 The electron configurations of Au25
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

could be described as (1S)2(1P)6, (1S)2(1P)5, and (1S)2(1P)4


superatom-like electron configurations, corresponding to
[Au25(SR)18]1−, [Au25(SR)18]0, and [Au25(SR)18]1+, respectively.
Accordingly, the free electron counts of these Au25 nano-
clusters are 8e, 7e, and 6e, respectively. Different free electron
counts endow these nanoclusters with different magnetism
behaviors.
The atomically precise structure of the [Au25(S-C2H4Ph)18]1−
and [Au25(S-C2H4Ph)18]0 nanoclusters provide models for corre-
Fig. 41 (a) CV curves of the [Au25(S-n-C4H9)18]− nanocluster in MeCN
lating the structure and the magnetic properties.232,234 The
(red) and after addition of 10 (black) or 30% DCM (blue). (b) CV curves of
Au25 nanoclusters capped by different ligands. Reproduced from ref. 145 electron paramagnetic resonance (EPR) signal provides evi-
with permission from American Chemical Society, copyright 2017. (c) dence for the magnetic properties of neutral Au25(S-C2H4Ph)18
DPV spectra of Au25 nanoclusters capped by different ligands. (d) nanoclusters. Specifically, EPR quantification of
Summary of the first oxidation/reduction potentials and the HOMO/ [Au25(S-C2H4Ph)18]0 in both the solution state and crystalliza-
LUMO energies of Au25 nanoclusters protected by different ligands. The
tion state suggest one unpaired spin per nanocluster. The ESR
special DPV spectra belong to the Au24Hg1 nanocluster, which will be
covered in the alloy Au25 section. Reproduced from ref. 572 with per- signal can only be detected at low temperature. When the
mission from American Chemical Society, copyright 2015. temperature increases to 100 K, a broader EPR signal is
observed. No signal can be observed when the nanocluster is
at room temperature.234 DFT calculations have been performed
different ligands demonstrated the ligand effects on the to investigate the mechanism of the magnetism (Fig. 42). An
electrochemical behaviors (Fig. 41b). Furthermore, the CV unpaired spin in the highest occupied Kohn–Sham orbital was
behaviors of the [Au25(S-n-C5H11)18]− nanocluster in different found in the neutral Au25 nanocluster, which was mainly dis-
conditions have been monitored, which guided the further tributed to the magnetism. Theoretically, the different consti-
electrocrystallization of the Au25(SR)18 nanoclusters on electro- tutions of the P-like orbital are the primary reasons for the dia-
lysis.145 Liao et al. also investigated the ligand effect on the magnetism of the anionic Au25 nanocluster, and the paramag-
electrochemical behaviors based on the Au25(SR)18 template.572 netism of the neutral Au25 nanocluster.234 Furthermore, the
Based on the DPV measurement, they observed the blue shifts neutral nanocluster has two lobes and shows distinct P-like
of the O1 and R1 peaks as well as the reduction of the HOMO– character, which also provides evidence that the Au25(SR)18
LUMO gaps in the Au25(S-C6H13)18, Au25(S-PhC2H4)18, nanocluster is a ligand-protected superatom. Moreover, the
Au25(S-C6H5)18 nanoclusters (Fig. 41c and d). cycle system between diamagnetic [Au25(S-C2H4Ph)18]1− and
Of note, the DPV spectra of the anionic Au25 protected by paramagnetic [Au25(S-C2H4Ph)18]0 has been built with the oxi-
SePh or S-C2H4Ph are totally different (Fig. 6a).194,195 (i) The dation and reduction.234 Ackerson and co-workers demon-
O1 and O2 peaks (0 and 0.3 V, respectively) of Au25(SR)18 were strated that the unoccupied P orbital of [Au25(S-C2H4Ph)18]1+
broadened and merged into a single peak in the Au25(SePh)18 make it diamagnetic;233 the Au25 with +1 and −1 charge states
centered at 0 V. (ii) The R1 peak of Au25(SR)18 at −1.6 V was sig- are diamagnetic, while the neutral Au25 is paramagnetic.232–234
nificantly red-shifted to −1.15 V for the Au25(SePh)18. (iii) The The Jin group reported the one-pot synthesis of water-
calculated HOMO–LUMO gap of Au25(SR)18 (1.6 eV) was soluble Au25(SG)18, 2 nm and 4 nm nanoparticles.374 Both the
reduced to 1.15 eV in the Au25(SePh)18 nanocluster, compared 2 nm and 4 nm Au nanoparticles exhibit paramagnetism,
to the HOMO–LUMO gap recorded from the optical-electronic while the Au25(SG)18 is diamagnetic.374 Krishna et al. deter-
spectroscopy.194,195 mined the paramagnetism of [Au25(S-C2H4Ph)18]0, the diamag-
netism of [Au25(PPh3)10(S-C12H25)Cl2]2+ and Au38(S-C12H25)24,
and the ferromagnetism of Au55(PPh3)12Cl6 using the super-
4.4 Magnetic properties of Au25 nanoclusters conducting quantum interference device (SQUID) magnet-
Unlike the diamagnetism of bulk Au, the discrete Au atoms ometer and theoretical calculations.375
exhibit paramagnetism. Since the influential publications of The Maran group revealed the paramagnetic properties of
the nondiamagnetic response in surface capped Au nano- neutral Au25 protected by the HS-C2H4 ligand.141 The magnetic
particles, magnetism in gold has triggered enormous inter- properties of this nanocluster were directly measured using
est.366,367 Au nanoclusters, being a bridge, fill the gap of dis- electron nuclear double resonance (ENDOR), through which

10792 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 43 (a) EPR spectra of the neutral Au25(SR)18 before (red curve) and
after oxidation (green curve) recorded at 6 K. Blue curves represent the
simulated spectrum of Au25. (b) Experimental (green curves) and theore-
tical (blue curves) EPR spectra of oxidized neutral Au25(SR)18 at different
temperatures. Reproduced from ref. 239 with permission from American
Chemical Society, copyright 2013. (c) 1H ENDOR spectra of
Au25(S-C2H4)180 (black curve), Au25(S-Ph)180 (red curve), Au25(S-Bu)180
(blue curve), and Au25(S-PhCH3)180 (green curve) in toluene solution at
5 K. (d) Baseline-corrected 1H-ENDOR spectrum (blue curve) and simu-
Fig. 42 Reversible switching of the magnetism between Au25(SR)18− lation (red curve) of the neutral Au25(S-Bu)18 in toluene at 5 K.
and Au25(SR)180 nanoclusters. (a) Schematic diagram. (b) Theoretical Reproduced from ref. 376 with permission from The Royal Society of
Kohn–Sham orbital energy-level diagrams (HOMO and LUMO sets) of Chemistry, copyright 2016.
Au25(SR)18− and Au25(SR)180 nanoclusters. (c, d) HOMO distributions at
different orientations. Reproduced from ref. 234 with permission from
American Chemical Society, copyright 2009.
be probed. For example, the atoms can be also affected by the
orbital distribution even if the atoms are as far as 6 Å from the
icosahedral Au13 core.376 Recently, they summed up the mag-
the hyperfine interactions between a surface-delocalized
netism of the neutral Au25(S-C2H4Ph)18 nanocluster
unpaired electron and the gold atom of neutral Au25 nano-
(Fig. 44).377 When in the film state, the [Au25(S-C2H4Ph)18]0 is
cluster were observed.141 In another work, the paramagnetic
paramagnetic; significantly, ferromagnetism is detectable
[Au25(S-C4H9)18]0 was obtained.142 Additionally, the neutral
when this nanocluster is at a low temperature.377 A small
Au25(S-C4H9)18 is capable of forming a nanocluster-based wire,
number of crystals exhibits ferromagnetism, but distinct para-
which exhibited a non-magnetic ground state and could be
magnetic, superparamagnetic, and ferromagnetic behaviors
regarded as a 1D antiferromagnetic system at low
could be observed when the quantity of the crystals is relatively
temperature.142
large. Theoretically, the observed magnetic behaviors respond
The way the electron-transfer affects the magnetism based
to both the spin–orbit (SO) coupling and the crystal
on the Au25(SR)18 nanocluster has been investigated by com-
distortion.377
paring the continuous-wave EPR (cw-EPR) behaviors (Fig. 43a
and b).239 The neutral Au25(SR)18 nanocluster showed the
obvious signal in the cw-EPR spectra, revealing the paramag- 4.5 Thermal stability of Au25 nanoclusters
netism of [Au25(SR)18]0, while small signals in the cw-EPR Previous reports have explained the stability of the Au25 as due
spectra were found in the oxidized products of [Au25(SR)18]0 in to the superatom arrangement of the electronic structures,
the temperature range 6–260 K (i.e., the Au25 with the +1 corresponding to the “jellium” with symmetric electronic
charge state is diamagnetic).239 potential around the metal kernel.328,378,379 The free electron
In 2016, Maran and co-workers assessed the interactions count, Ns, of anion Au25 is 8, being a “magic” number (where
between the metallic kernel and the capped ligands of para- Ns = Nva − L – q; Nva is the number of metallic Au atoms; L is
magnetic Au25(SR)18 nanoclusters by studying the interactions the number of one-electron withdrawing ligands; q is the inte-
of unpaired electrons with pulse electron nuclear double reso- gral charge of the nanocluster).328 With the high thermal stabi-
nance (ENDOR, Fig. 43c and d).376 A pronounced increase in lity in the solid, crystal, or the solution state, various physico-
spin-polarization occurred when the temperature was reduced chemical properties of the nanoclusters could be detected.
to 5 K. In this context, the unpaired electron enabled the inter- Jin and co-workers investigated the stability of two Au–S
actions between the metallic core and the capping ligands to binding modes in the Au25(SG)18 nanocluster according to the

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10793
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 44 (a, b) Experimental (black curves) and calculated (red curves)


cw-EPR spectra of the Au25(S-PhC2H4)180 amorphous film at different
temperatures. (c, d) Hysteresis cw-EPR experiments for
Au25(S-PhC2H4)180 microcrystals with temperature decreasing or
increasing processes. Reproduced from ref. 377 with permission from
American Chemical Society, copyright 2017.

results from the NMR and optical absorption spectra.380 From


the crystal structure of the Au25 nanocluster, there are two
modes of S atoms: (i) S atoms linking the kernel Au and motif
Au; (ii) S atoms linking the two motif Au atoms. NMR and UV-
vis results confirmed that the thermal stability of the Au–S
bond in mode i is much greater than that of mode ii. In this
context, with high energy to inspire the nanocluster, the S
atoms (or ligands) in mode ii tend to fall off from the nano-
cluster surface. Laser desorption ionization (LDI) mass spec-
trometry was performed to verify the above-mentioned con-
clusions, by which the abundant Au25(SG)12 fragment peak Fig. 45 (a) DSC curves for the Au25(S-C6H13)18 with charge states of −1
(with six SG ligands lost from the intact nanocluster) has been (blue curve), 0 (red curve), and +1 (green curve). Reproduced from ref.
381 with permission from American Chemical Society, copyright 2012.
detected.
(b) Time-dependent UV-vis absorptions of Au25 protected by S-C8H17 or
The anionic Au25(SR)18 nanocluster exhibits the “noble-gas Se-C8H17 ligands. Reproduced from ref. 190 with permission from
superatom” electron configuration, that is, the noble-gas-like American Chemical Society, copyright 2012.
(1S)2(1P)6 electron configuration. When the anionic Au25 is oxi-
dized, the open-shell radical (1S)2(1P)5 and diradical (1S)2(1P)4
configurations can be observed in the [Au25(SR)18]0 and
[Au25(SR)18]1+ nanoclusters, respectively. Ackerson and Different capped-ligands affect the electronic configur-
Tofanelli found that the thermal stability of Au25 with close- ations of the nanoclusters, and thus influence their properties.
shell electronic configuration (i.e., anion Au25) was signifi- Han and co-workers theoretically investigated the ligand
cantly higher than the Au25 with the other two charge effects on the thermal stability of nanoclusters.383 Au25 nano-
states.381 Specifically, differential scanning calorimetry (DSC) clusters protected by aliphatic thiols exhibited higher thermal
was performed on the Au25 nanoclusters with different charge stability than those capped by aromatic thiols, mainly because
states. The decomposition temperature of anion Au25 is of the high electrochemical and thermodynamic stability of
227.5 °C, which is higher than 221 °C and 209 °C for the Au250 aliphatic thiols.383
and Au251+ nanoclusters, respectively (Fig. 45a).381 Jiang and Ackerson and co-workers systematically analyzed the
Ouyang found that the environment (such as crystal packing thermal stability of [Au25(SR)18]1−/0/1+.245 Firstly, the stability
mode in the crystal state and the polarity of the solvent) had order of Au25 with different charge states is [Au25(SR)18]1+ <
notable impacts on the ligand–metal combination mode, [Au25(SR)18]0 < [Au25(SR)18]1−. The enhanced thermal stability
which further influenced the thermal stability.382 was observed within the free oxygen environment (vacuum

10794 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

environment). In addition, the Au25 capped by ligands with therapy,407 to name a few. Regarding Au25 nanoclusters, the
long carbon tails showed higher stability. Besides, different fluorescence intensities of Au25 nanoclusters protected by oil-
thermal stabilities of Au25 with the same charge state was also phase ligands are relatively low compared with those protected
observed in different solvents (most stable in the toluene).245 by aqueous-phase ligands (e.g., GSH or BSA ligands).214,274 In
Compared with the S atom in the thiols, the atomic radius, particular, the highly luminescent Au25-BSA nanocluster with
as well as the electronegativity of Se, is closer to gold metal. QY of about 6% at λem = 640 nm is worth noting.214 In this
Consequently, the Au–Se bonds are more covalent and exhibit context, the BSA–Au25 nanocluster has found tremendous
a higher bond energy than Au–S bonds. This is expected to optical applications. In addition, together with the strong
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

prepare more stable Aun(SeR)m than Aun(SR)m nano- luminescence, the high biocompatibility, good photostability,
clusters.96,185,191 the Negishi group and our group have syn- low toxicity, and anticancer activity of such nanoclusters make
thesized the Au25(SePh)18 by ligand-exchange or in situ syn- them highly promising in cell labeling, phototherapy, biosen-
thesis methods.187,192 Time-dependent UV-vis spectra illustrate sing, bioimaging and biotherapy.99–102
the significantly enhanced thermal stability of Au25–Se com- Apart from the optical applications (such as chemical
pared with Au25–S. Specifically, the peaks in the UV-vis spectra sensing) and biological applications, noble metal nanoclusters
of Au25(SePh)18 were maintained even after 2 days. However, have been widely applied in catalysis, due to their distinct fea-
some peaks of Au25(S-C2H4Ph)18 gradually disappeared or tures such as high stability, high surface area, high proportion
broadened after 1 day (Fig. 45b).190,192 With respect to the of metallic atoms, and quantum size effects.25,39,408–419
negative-ion MALDI results, Au25(S-C2H4Ph)18 will generate the Furthermore, the accurate catalytic sites and catalytic pro-
Au21(S-C2H4Ph)14 fragment peak, which emerges by dissociat- cesses could be speculated because of the atomically precise
ing a Au4(S-C2H4Ph)4 unit from the intact nanocluster. structures of these nanoclusters.39,411,413 From these specu-
Relatively small amounts of Au25(SePh)18 nanoclusters are dis- lated conclusions, novel nanoclusters with more enhanced
sociated in the MALDI measurement, demonstrating the catalytic activities can hopefully be obtained through purpose-
higher stability of the Au25(SePh)18 nanocluster relative to fully tailoring the structures and compositions of the template
Au25(S-C2H4Ph)18.190 More studies comparing the Au25 nano- nanoclusters. It should be noted that the capped ligands of
cluster capped by thiolate and tellurolate have been reported, the nanoclusters have significant impacts on the
which also provide evidence for the higher stability of Te–Au25 catalysis.97,219,420–429 In most cases, the catalytic activities of
nanoclusters.193 Theoretically, Pei and co-workers demon- Au nanoclusters capped by ligands are somewhat lower than
strated the higher thermal stability of Se–Au25 compared with those nanoclusters in the partial or completely exposed
S–Au25 by DFT calculations.186 For example, the reaction state.219,420–424 Mild thermal treatments are usually employed
Au25(SR)18 + 18[HSeR] → Au25(SeR)18 + 18[HSR] is indeed an to remove the capped ligands to expose some catalytic sites
exothermic process, regardless of whether –R represents –CH3, (note that the sizes and structures of the nanoclusters are not
–C2H5Ph or –Ph. Furthermore, the enhanced stability was not changed in these processes).422,423–427 On the other hand, in
only observed in the Au25 system, but other nanocluster some cases, the ligands could be utilized in selective catalysis
systems (e.g., Au23, Au24, Au28, Au36, Au38 and Au40) also follow such as reactant selectivity or product selectivity.97,421,428,429
the rules.186 The Au25 nanocluster is the most investigated catalyst due to
In the vacuum environment, the enhanced stability (up to the facile synthesis and high yields. Furthermore, the catalytic
165 °C) of the Au25(SR)18 nanocluster was observed using a sites and reaction processes can be reasonably speculated on
thermogravimetric analyzer (TGA). When the outside tempera- the atomic/molecular level through DFT calculations based on
ture is higher than 165 °C, the weight loss of the Au25 nano- the accurate crystal structure of the Au25 nanocluster.97,411
cluster occurs (the loss comes from the capped ligands and Furthermore, being a versatile catalyst, the Au25 nanocluster is
the counterions).190 In this context, the measurements for capable of inducing thermo-catalysis, electrocatalysis and even
investigating the properties of Au25 nanoclusters with main- photocatalysis.
tained configuration and structure should be below this temp- In this subsection, the tremendous applications of Au25
erature. Besides, it has been confirmed that the bare metals of nanoclusters (mainly BSA–Au25) such as environmental moni-
the nanoclusters significantly affect the catalytic properties, toring, chemical sensing, bioimaging, and photodynamic
such as hydrogenation or oxidized catalysis, which are dis- therapy are summarized. Additionally, some significantly cata-
cussed in Section 5.2. lytic reactions catalyzed by Au25 nanoclusters are highlighted.

5.1 Chemical sensing

5 Application Au nanoparticles have found widespread applications in


chemical sensing, bioimaging37,113 and nanomedicine,430
Due to the intriguing absorption and luminescence properties, which demonstrate their strong luminescence and high bio-
noble metal nanoclusters have been widely used in environ- compatibility.431 However, the first researchers who investi-
mental monitoring and chemical sensing,111,363,384–390 drug gated the Au nanoparticles found that the PL intensity of these
delivery and cancer therapy,391–393 bioimaging and nanomaterials was up to 0.1% or lower,432 and thus not suit-
biolabeling,394–398 nanomedicine,399–406 and photodynamic able for chemical sensing or fluorescence imaging. With the

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10795
View Article Online

Review Nanoscale

in-depth research on the noble metal nanoclusters, it is


suggested that the nanoclusters can exhibit higher PL QY rela-
tive to the nanoparticles (however, the details of the lumine-
scence remain unclear for now).23,433 Among these nano-
clusters, water-soluble Au25 (mainly protected by GSH or BSA
ligands) represents one of the most researched nanoclusters
that have found extensive applications, such as chemical
sensing. Fluorescent nanoclusters for ion and bio-molecule
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

detection are of great significance due to their simplicity and


low detection limit (i.e., high sensitivity).434 In particular, GSH
or BSA has been adopted to produce highly fluorescent Au25
clusters with high chemical stability and simple synthetic
reproducibility in aqueous solutions.214 Furthermore, the
better water solubility of these Au25 nanoclusters compared
with oil-phased Au25 nanoclusters make them promising
chemical sensors for sensing ions and some other molecules.
In 2009, Xie et al. reported the synthesis of the highly fluo-
rescent BSA–Au25 nanocluster.214 Based on this luminescent
nanocluster, they further developed a simple label-free method
for detecting the Hg2+ ions with high selectivity and sensi-
tivity.435 The PL QY of BSA–Au25 was almost maintained when
it was dipped in solutions of other metal ions such as Ag+,
Cu2+, Zn2+, Mg2+, K+, Na+, Ni2+, Mn2+, Cd2+, Fe3+, Pd2+, Pt4+,
Co2+, Pb2+, and Ca2+. In particular, the addition of Hg2+ effec- Fig. 46 (a) Illustration of Hg2+ sensing based on the BSA–Au25 nano-
clusters. (b) PL spectra of BSA–Au25 before and after the sensing
tively quenched the fluorescence of the Au25 nanocluster
process. (c) Photographs of the test strips with the BSA–Au25 nano-
(Fig. 46a–c). BSA–Au25 showed a remarkably high selectivity for cluster under UV light after dipping in the solutions of 50 mM of various
Hg2+ ions compared to other metal ions, and the detection metallic ions. Reproduced from ref. 435 with permission from The Royal
limit of Hg2+ was only 0.5 nM.435 They proposed that the Society of Chemistry, copyright 2010. (d) Illustration of the Hg2+ sensing
sensing mechanism was based on the high-affinity metallophi- by the use of the POSSFF-Au25 system. Reproduced from ref. 436 with
permission from American Chemical Society, copyright 2011.
lic Hg2+–Au+ interactions (note that the Au+ ions are distribu-
ted mainly on the core-surface of the nanocluster).435
Liu and co-workers investigated the fluorescence resonance
energy transfer (FRET for short) between the blue-fluorescent
conjugated-oligomer-substituted polyhedral oligomeric silses- quenching mainly occurred through the oxidation of Au(0) to
quioxane (POSSFF) and red-fluorescent BSA–Au25 nano- Au(I) in the kernel of the BSA–Au25 nanocluster mediated by
clusters.436 The pink fluorescence of the FRET system specifi- nitrite ions. They also designed a BSA–Au25 membrane-based
cally turns into blue in the presence of Hg2+ ions rather than sensor to easily detect the nitrite in urine samples, which
other metal ions (Fig. 46d). The reason is that the strong enabled the detection of nitrite at a level as low as 100 nM in
metallophilic Hg2+–Au+ interaction quenches the red fluo- aqueous solution.437
rescence from the BSA–Au25 nanocluster and thus, only the Ion sensing based on the fluorescent nanoclusters usually
blue fluorescence of POSSFF is exhibited. These FRET systems depends on the weak interactions of the metallophilic bond,
allow for the visual detection and precise quantification of such as Hg2+ and Au+.435–437 In 2013, Yu and co-workers
Hg2+ ions with a low detection limit in aqueous solution or the explored the weak interaction of Hg2+/Cu2+ with BSA–Au25
cell.436 using steady-state and time-resolved luminescence and transi-
Guan et al. developed an effective separation method to ent absorption measurements (Fig. 47).438 The delayed fluo-
separate the pure BSA–Au25 nanocluster from the mixture of rescence (DF) was quenched via an effective triplet state elec-
BSA and BSA–Au25 nanoclusters.389 Specifically, the isolation is tron transfer when the BSA–Au25 nanocluster met the Hg2+
a co-precipitation process, which precipitates the BSA–Au25 ions. On the other hand, the Cu2+ ions do not alter the DF due
nanocluster by reacting with OH− and Zn2+ sequentially. The to the absence of the metallophilic interactions between the
resulting purified BSA–Au25 nanocluster exhibited enhanced Au and Cu. This was the first time the triplet-state quenching
fluorescence, improved sensitivity, and high selectivity in the of the BSA–Au25 nanocluster was observed through the metal-
fluorescence detection of H2O2 or Hg2+ ions.389 lophilic bond.438
Unnikrishnan et al. detected nitrite ions in urine based on Au25(SG)18 can also be used as a chemical sensor. Cai and
the fluorescence quenching of the BSA–Au25 nanocluster and co-workers reported the fluorescent iodide sensing using the
then discussed the probable roots leading to the PL quench- Au25(SG)18 nanocluster.439 They also proposed a particular
ing.437 Based on the detected results, they proposed that the sensing mechanism defined as affinity-induced ratiometric

10796 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 47 (a) Fluorescence spectra of the BSA–Au25 nanocluster with


various concentrations of the Hg2+ ions. (b) Fluorescence time evolution
on the nanosecond timescale (670 nm) with an increasing concentration
of Hg2+ ions. (c) Fluorescence time evolution on the microsecond time-
scale (670 nm) with an increasing concentration of Hg2+ ions. (d)
Transient absorption spectra of the BSA–Au25 nanocluster in the pres-
ence or absence of Hg2+ with a fixed time gate of 10 ns and different
delays. Reproduced from ref. 438 with permission from Wiley-VCH,
copyright 2013.

and enhanced fluorescence, which provided an alternative


strategy for devising nanocluster-based nanosensors.439
Metal ions typically cause PL quenching, as discussed Fig. 48 (a) The fluorescence spectra of the Au25(SG)18 nanocluster with
above.435–439 However, there are also some special cases where different concentrations of Ag+. (b) Linear relation between the fluor-
escence intensity and the Ag+ concentration. Reproduced from ref. 440
the PL QY of the nanocluster sensor is enhanced in the pres- with permission from Wiley-VCH, copyright 2012.
ence of specific ions. The Jin group reported a Ag+ sensor by
detecting the PL enhancement of the Au25(SG)18 nanocluster,
and the detection limit of Ag+ is about 200 nM (Fig. 48).440 The
posed as the CSH etching-induced fluorescence quenching.
good linear fluorescence response between the concentration
They also found that other thiol-containing compounds such
of Ag+ and the PL QY was constructed from 20 nM to 11 μM.
as GSH and cysteine and 19 other natural amino acids did not
Furthermore, the great selectivity among 20 types of metal
interfere with the CSH-induced etching process.443
cations makes Au25(SG)18 an ideal candidate for sensing the
Ag+ ions. Finally, three factors responsible for the unique fluo-
rescence enhancement caused by Ag+ ions were proposed: (i) 5.2 Biological application
the oxidation state alternation of Au25(SG)18; (ii) the interaction Due to the following six reasons, the biological application of
of Ag0 (reduced by Au25(SG)−) with Au25(SG)18; (ii) the inter- the Au25 nanoclusters is highly attractive: (1) The nontoxicity
action of Ag+ with the metallic kernel Au25(SG)18 of the nanoclusters (as small-sized nanoparticles). (2) The
nanocluster.440 ultra-small size that makes them easily enter the cell. (3) The
Besides metal ion detection, the Au25 nanocluster can also high biocompatibility and good photostability. (4) Strong
be adopted to sense the macromolecules or biomolecules. luminescence that can easily label the cell. (5) The potential
Huang and co-workers designed a label-free and simple sensor anti-cancer activity. (6) The functional ligands that cap the
to detect clioquinol using the Cu2+-mediated fluorescent BSA– metallic kernel, enabling the nanocluster to possess the poten-
Au25 nanocluster.441 Shu et al. reported a novel ratiometric tial to be a drug delivery carrier. In this section, the biological
fluorescent sensor based on the BSA–Au25 nanocluster for applications of Au25 nanoclusters, including bioimaging, bio-
label-free, separation-free detection of the tris(2-carboxyethyl) labeling, and biotherapy, are reviewed.
phosphine (TCEP).442 The detection limit of TCEP was 130 5.2.1 Bioimaging and biolabeling. The wide excitation
nM, achieving higher sensitivity compared to previous wavelength of noble metal nanoclusters offers multiplexing
reports.442 Zhang and co-workers demonstrated the label-free capability. This is highly desirable for the applications of these
and separation-free detection of cysteamine (SCH) using the nanomaterials in the biosensing, fluorescence biological
BSA–Au25 nanocluster.443 The detection mechanism was pro- imaging, and biolabeling fields.1,99,111,201,399,444–448

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10797
View Article Online

Review Nanoscale

Liu and co-workers facilitated the FRET system (discussed clusters and found that the small-sized nanoclusters exhibited
in section 5.1.1) for use in the in-vitro cells.436 The preserved notably stronger antimicrobial activity relative to large nano-
ion-selective FRET in cells make this FRET system (composed particles (Fig. 49).452 Specifically, the nanoclusters with ultra-
by the POSSFF and BSA–Au25 nanocluster) effective for multi- small size could kill both Gram-positive and Gram-negative
color intracellular sensing of Hg2+ ions.436 Based on the bio- bacteria. They proposed that the interaction between the nano-
imaging in vivo or in vitro, the accurate sites that the biomater- cluster and bacteria-induced a metabolic imbalance in bac-
ials act on, or the precision processes of the biomaterials terial cells, which led to an increase in intracellular reactive
absorptions and deabsorptions, can be acquired. The Zheng oxygen species production that killed the bacteria.452
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

group carried out extensive work in nano-biomedicine with the


assistance of bioimaging and biolabeling.201,399,445–447 The
Ackerson group investigated the structure–activity relation- 5.3 Catalysis
ships for the biodistribution, pharmacokinetics, and excretion Catalysis is one of the most significant properties/applications
of atomically precise nanoclusters (e.g., the Au25(SG)18 nano- in the nanocluster field.17,25,39,408–419 Despite the chemical
cluster) in a murine model through bioimaging and inertness of gold element or gold on a large-sized scale (this is
biolabeling.100 also the reason why Au was not pursued for catalytic appli-
5.2.2 Biotherapy. Noble metal nanoclusters are attractive cations at an earlier date), gold at small sizes (small nano-
for biological and biotherapeutic applications. A current trend particles or so-called nanoclusters) always possesses high
in the development of nanoclusters in the biological field is to activity in several catalytic fields.453 There have been many
promote photodynamic or photothermal therapy for cancer review articles addressing the catalytic applications of noble
treatment.101,102,449,450 metal nanoparticles and nanoclusters in
The generation of highly reactive singlet oxygen 1O2 is of detail.16,17,25,30,33,40,454–456 In this context, some significant
significance for the application of photodynamic therapy for
cancer treatment. Jin and co-workers directly photosensitized
the Au25(SR)18 nanoclusters and demonstrated that 1O2 can be
efficiently produced in this process.407 The water-soluble Au25
nanocluster was explored for the cytocompatibility and photo-
dynamic activity toward cancer cells. Specifically, the 1O2 can
be produced on the Au25 nanocluster under near-IR photo-
excitation at 808 nm, which is very attractive for photodynamic
applications.407
Katla et al. investigated the photothermal process of the
Au25(SG)18 nanocluster.101 Using the 808 nm laser source with
power of 10 W cm−2, MDA-MB-231 breast cancer cells were
completely broken and died due to the photothermal activities
of Au25(SG)18 nanoclusters.101 This was the first time that the
Au25(SG)18 nanocluster was demonstrated to be effective in
photothermal therapy.101
Xie and co-workers presented a new type of radiosensitizer
(Au25(SG)18 nanocluster) for cancer radiotherapy, which led to
strong enhancement for cancer radiotherapy.102 Specifically,
the Au25(SG)18 nanocluster preferentially accumulated in the
tumor via the enhanced permeability and retention effect of
this nanomaterial, leading to the strong enhancement of the
cancer radiotherapy in the tumor. More importantly, these
small sized Au25(SG)18 nanoclusters can be efficiently cleared
by the kidneys, which minimizes any potential side effects.102
The attractive features of the small-sized nanoclusters, includ-
ing excellent tumor accumulation, strong radiation enhance-
ment, and low toxicity (efficient renal clearance), might make
them promising bio-therapeutic drugs.102
Bulk gold (or Au nanoparticles) is well known to be chemi-
cally inert. However, when the size of the nanoparticle
Fig. 49 (a) Higher killing efficiency of Au25 nanoclusters relative to Au
decreases to the sub-nanometer dimension, the obtained
complexes and nanoparticles toward the aureus bacteria. (b) The high
nanoclusters begin to exhibit multiple physicochemical pro- killing efficiency of the Au25 nanocluster observed in other bacteria
perties and high activities.94,103,159,450,451 The Xie group moni- models. Reproduced from ref. 452 with permission from American
tored the antimicrobial effect of gold nanoparticles and nano- Chemical Society, copyright 2017.

10798 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

examples focusing on the catalytic properties of Au25 nano-


clusters are summarized here.
Au25 nanoclusters are known for their alterable electronic
structure (−1, 0, +1), 456 and they can be shifted into each
other by oxidation or reduction.231–234 Among them, the Au25−
nanoclusters are the most investigated catalysts or precursors25
due to the facile synthesis and high yield.92 Owing to their
surface geometric effect (i.e., surface atom arrangement and
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

affluent low-coordinated metal atoms), the quantum size


effect, as well as the electronic effect,458 satisfactory conversion
and excellent selectivity could be achieved. Besides, the well-
defined crystal structure of the Au25 nanocluster brings us a
glimpse into the catalytic mechanism and the relationship
between structures and properties. As versatile catalysts, the
Au25 nanoclusters could induce thermo-catalysis, electrocataly-
sis and even photocatalysis.
5.3.1 Active site and charge effect of the Au25 nanocluster.
The mechanism for styrene oxidation catalyzed by gold nano-
clusters has been investigated. Zeng and co-workers proposed
two plausible mechanisms to demonstrate that the Au(I) sites
in the semi-ring staple motifs on Au25(SR)18 are the major
active sites for styrene oxidation rather than the inner metallic Fig. 50 (a, c) Correlation between reactant binding energy and reaction
Au sites.252 Au(I) is oxidized into the tetra-coordinated Au(III) turnover frequency (TOF) catalysed by the Au25(SR)18 nanoclusters with
structure upon O2 adsorption, which is considered a key step different charge states. (b, d) Proposed CO2 and H+ or CO and OH− co-
for the activation of O2 molecules. The mechanisms provide a adsorption model. Reproduced from ref. 459 with permission from The
Royal Society of Chemistry, copyright 2014.
deep understanding of other thiolated nanocluster-catalyzed
reactions on the nanocluster surfaces.252
The Au25 nanoclusters are known for their regular charge
Ligands can affect not only the stability of the gold nano-
states. Notably, the charge states (−1, 0, +1) have significant
clusters, 383,460 but the catalytic properties as well. It is well
influence on their electrocatalytic activities.459 Charge state-
established that the removal of the capping ligands will expose
dependent reactivity was observed for the O2 reduction, CO
more naked gold atoms for reactant access, which usually
oxidation and CO2 reduction reactions (Fig. 50). The Au25−
leads to conversion enhancement. However, the residual
nanocluster shows promoted CO2 reduction performance by
ligands also affect the selectivity. Tsukuda et al. confirmed
stabilizing CO2 and H+ co-adsorption, while the Au25+ nano-
that fully covered Au25 nanoclusters are inert in the aerobic
cluster exhibits enhanced CO oxidation activity by stabilizing
oxidation of benzyl alcohol, due to site blocking, but the
CO and OH− co-adsorption.459 In addition, the Au25+ nano-
selectivity toward benzaldehyde increases along with decreas-
cluster displays reduced O2 reduction activity due to their
ing ligand residuals (Fig. 51).421 It is speculated that the role
stronger binding ability to the OH− reaction products. These
results indicate that the catalytic activity could be tailored by
changing the electronic structure.459
5.3.2 Support effect and ligand effect. The support effect
on the catalytic activity of heterogeneous gold nanoclusters
has been evaluated.423 Au25 nanoclusters have been deposited
on various inorganic supports via the conventional impreg-
nation method and calcined under N2 stream. Since particle
size is the key factor for the catalytic activity and overheating
will lead to severe aggregation, all the samples are treated
under 300 °C.423 The strength of the Au-support interaction
plays a primary role in the growth of the Au25 nanocluster.
Hydroxyapatite (HAP) and TiO2 could effectively prevent the
growth of Au nanoclusters during the sintering, while aggrega-
tion would occur by using active carbon (AC), pyrolyzed gra-
phene oxide (PGO) and SiO2. Correspondingly, the HAP sup-
ported Au25 catalyst displays the best performance in nitro- Fig. 51 Ligand effect in the aerobic alcohol oxidation catalysed by the
benzene hydrogenation and styrene oxidation, highlighting Au25(SR)18 nanocluster. Reproduced from ref. 421 with permission from
the importance of the support effect.423 American Chemical Society, copyright 2014.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10799
View Article Online

Review Nanoscale

of thiolates is to weaken the oxidation capability of Au25 by


electron withdrawing behavior and to prohibit the esterifica-
tion reaction on the nanocluster surface by site isolation.421
Besides residual concentration of the ligand, the chain
length and functional moiety are factors tuning the catalytic
activity of gold nanoclusters.461 Shorter chain length gives
reactants better ability to access the metallic core, which will
hopefully exhibit better catalytic activity. The aromatic ligands
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

are particularly effective in increasing the activity, which is


probably facilitated by the π-π stacking effect between the
ligand and 4-nitrophenol.461 On the contrary, ligands with
amine groups lower the conversion mainly due to the steric
blocking and electronic modifications to the metallic kernel.
These results demonstrate that the functional groups of
ligands play crucial roles in catalysis.461
Li et al. also confirmed that the activity and selectivity of
the catalysts are largely influenced by the chemical nature of Fig. 52 (a) Illustration of CO oxidation catalysed by Au25-CeO2. (b)
the protecting thiolate ligands.154 In Ullmann coupling reac- Reaction-temperature-dependent CO conversion by Au25(SR)18/MOx
catalysts. (c) Reaction-temperature-dependent CO conversion over
tions catalyzed by CeO2 supported Au25 nanoclusters, aromatic
Au25(SR)18/CeO2 after different pre-treatments. (d) CO conversion at
ligands not only give rise to a higher conversion but also dra- 80 °C over the Au25(SR)18/CeO2 pre-treated at different temperatures
matically enhance the yield of the heterocoupling products. with or without H2O. Reproduced from ref. 464 with permission from
Specifically, the S-Nap protected Au25 nanocluster exhibits sig- American Chemical Society, copyright 2012.
nificantly improved thermal stability, antioxidation properties
and higher heterocoupling selectivity than those of the
Au25(S-PhC2H4)18 nanocluster.154 kind of Au species: Auδ+ (0 < δ < 1), Au+, and Auδ− (0 < δ < 1).
Liu and co-workers designed a continuous flow catalysis of The former is characteristic of CO oxidation at low tempera-
Au25 nanoclusters based on the high aspect-ratio carbon nano- ture, while the other two are just the opposite.465
tubes.462 They proposed that the catalytic effect originated Au25 nanoclusters immobilized inside mesoporous SBA-15
from the gold metallic core of the nanocluster. In addition, the as precursors have been proved to be good catalysts for benzyl
capped ligands played two key roles in this catalysis: (i) serving alcohol oxidation.466 After calcination, all the ligands on the
as a well-defined surfactant to effectively assemble the nano- nanocluster surface are removed and thus the catalyst demon-
clusters onto the carbon nanotubes (CNTs) in aqueous solu- strates its best activity. Without calcination, over-oxidation of
tion; (ii) serving as efficient protecting ligands for the metallic alcohol is inevitable, resulting in the low selectivity of the cata-
kernel to avoid agglomeration.462 lytic products. Au atoms in the staple motif (–S–Au–S–Au–S–)
5.3.3 Oxidation. Since Haruta found the Au nanoparticles from the outer shell where the O2 dissociation occurs, serve as
to be effective for the oxidation of Co in 1987,463 a variety of Au the primary active sites for the catalytic oxidation. Increasing
nanoparticles prepared by different methods have emerged as calcination temperature leads to more accessible surfaces and
efficient catalysts for this conversion. When referred to the CO more reduced metal nanoparticles, which give rise to better
oxidation in the nanocluster field, the CeO2 supported Au25 conversion.466
nanocluster shows the better performance relative to Fe2O3 For styrene oxidation, CeO2 supported Au25 nanoclusters
and TiO2, indicating that the active sites are located on the per- completely consume the substrate with the aid of TBHP (tert-
imeter of the interface (Fig. 52).464 Two contributors are estab- butyl hydroperoxide).467 As a weak oxidation agent, O2 helps to
lished for the oxidation: oxygen incubation of the catalyst and achieve 99% selectivity toward benzyldehyde, albeit the conver-
moisture. Oxygen pretreatment of the nanoclusters at 423 K sion is low (5%). XAFS revealed the contraction of Au–Au bond
for 1.5 h notably boosts the catalytic activity, while the water and depletion of d-band in Au25 due to strong ligand
vapor benefits the transformation.464 binding.468 The positively charged shell Au atoms could there-
Thiolate ligands could protect the integrity of the nano- fore be preferentially anchored to the surface of CeO2, where
clusters, but they have to be partially/completely removed to the atoms activate the nucleophilic CvC bond of styrene.
activate CO oxidation.465 The capping ligands hinder the CO Compared with the (111) surface of CeO2 nanorods, Au25 on
access to the metal surface and hide the active sites, which the (110) surface of CeO2 displays better selectivity towards the
have been demonstrated by DFT calculations. Intriguingly, the styrene epoxide, which is assigned to the different charge dis-
MvK mechanism undertakes the dominant role to convert CO, tributions and electron transfer of Au25 on different surfaces of
which firstly reacts with the lattice oxygen of CeO2, and then CeO2 (Fig. 53).467
gas-phase O2 replenishes the consumed oxygen, while the L-H However, without oxide support, the homogeneous Au25
mechanism of CO and O2 activated by exposed Au sites takes nanocluster is unstable in oxidizing conditions for catalytic
the rest. The removal of the capping ligands will expose three styrene oxidation.469 XPS results confirmed that the mono-

10800 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

lytic activity toward 4-nitrophenol reduction and styrene


oxidation.475,476
ZnAl-hydrotalcite-supported Au25 nanocluster showed com-
plete conversion of 3-nitrostyrene and selectivity with respect
to 3-vinyl aniline over a wide reaction duration and tempera-
ture range.477 The inertness to vinyl is due to the fact that only
the nitro group is selectively adsorbed and activated on the
catalyst surface. The diameter of Au nanoclusters is one of the
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

most important factors affecting the catalytic activity.


Fortunately, residual S and the epitaxial interaction between
Au (in nanocluster) and the supports contribute to the high
thermostability of the Au25 nanocluster.477
The Jin group reported the first example of the stereo-
selective hydrogenation of bicyclic ketone by the
Fig. 53 Illustration of the proposed mechanism for the selective oxi- Au25(S-PhC2H4)18 nanocluster, though both the metal core and
dation of styrene by Au25 nanocluster immobilized on the (110) surface the protecting thiolates are non-chiral.478 The electron-rich
of CeO2 nanorods. Reproduced from ref. 467 with permission from The Au13 kernel activates the CvO bond, while the electron-
Royal Society of Chemistry, copyright 2013.
deficient Au12 shell provides active sites for H2 adsorption and
dissociation. The first H atom attacked CvO bond will not
change the stereoselectivity; however, the second H selectively
nuclear Au(I) species are the active sites, which are able to con- attacks in the direction of axial position, ascribed to the
tinue to catalyze the conversion after the decomposition of the spatial environment of Au25. The combination of the confined
Au25 nanocluster. Notably, Au25 nanoclusters were completely construction of nanoclusters and the configuration of the
degraded before the product was monitored.469 bicyclic ketone leads to a nearly 100% selectivity.478
The purging of ligands usually brings about better activity, The inertness to vinyl and preference for carbonyl of the
but not in the oxidation of trans-stilbene by SBA-15 supported Au25 nanocluster make it available for selective hydrogenation
Au25 nanoclusters.470 After immobilization, the Au metal was of α,β-unsaturated ketones.479 Intriguingly, the catalyst was
reduced to neutral while the 4-aminothiophenolate ligand was active with 100% selectivity even at 0 °C. Moreover, the unne-
oxidized. Given that the active species Au(I) is absent, lower glectable support effect was observed, and the best oxide
conversion is reasonable. In addition, the oxidized ligands lost support was Fe2O3. With the exterior twelve Au atoms forming
the ability to protect the nanoclusters, which resulted in the an incomplete shell on the icosahedral core, eight sites were
decomposition of the latter.470 left open for substrate access (Fig. 54a). Starting with the unsa-
5.3.4 Reduction. Though the Au25 nanocluster is not the turated ketone, the disassociation of H2 takes place between
best catalyst for the reduction of 4-nitrophenol among the the Au atoms from the motif and the CvO of the substrate
atomically precise Au nanocluster family, 471 the 4-nitrophenol with the assistance of solvent ethanol, forming hydrogen
reduction is usually employed as a model reaction to explore bonds with the partially hydrogenated carbonyl to stabilize it.
the application of these nanomaterials. The monolayer pro- Finally, the second H is transferred to the partially hydrogen-
tected nanoclusters are very stable during the catalysis, and ated group and forms the unsaturated alcohol (Fig. 54b).480
these nanoclusters could be isolated and characterized after Hydrogenation of aldehyde occurs with the assistance of
the reaction was completed.472 EXAFS analysis demonstrated pyridine. Ammonia, a better lone-pair donor, leads to higher
the partial removal of thiolate stabilizers under mild heating conversion. Addition of transition-metal ions, such as Cu+,
conditions, and the resulting species were catalytically active Cu2+, Ni2+ and Co2+, to the system as Lewis acids significantly
for 4-nitrophenol reduction.473 This may explain why the sele- enhance the catalytic conversion of aldehyde hydrogen-
nate-stabilized and hyperstar polymers encapsulated Au25 ation.481 Detachment of the “Au-SR” unit exposes the Au13-core
nanoclusters give similar responses to the reduction of 4-nitro- of Au24(SR)17 to reactants, which provides active sites for the
phenol.194,474 Two factors are suggested that influence the catalytic hydrogenation (Fig. 54c). Moreover, Mz+ ions
catalytic activity: the removal degree of the thiol and the adsorbed on the Au13-core surface produce new possible active
growth of nanoclusters via sintering. Thermal treatment under sites for the aldehyde hydrogenation.481
air at 200 °C leads to the nearly exhaustive removal of thiolate The mechanism of the TiO2-supported spherical Au25 nano-
ligands with little to no growth in nanocluster size, while little clusters-catalyzed semi-hydrogenation of terminal alkynes has
size enhancement was observed when the nanocluster exhibi- been established.97 Three external gold atoms from the three
ted the best maximum activity (at 250 °C pretreatment). When adjacent staples comprise a triangular Au3 site, which gives
coated with silica, the Au25 nanocluster showed high thermal easy access to the reactants (Fig. 54d). The terminal alkynes
stability.475,476 After high temperature calcination, the capping undergo deprotonation in the first place, then both H2 and
ligands were removed and the active sites on the nanocluster deprotonated alkynes are adsorbed on the triangular facet
surface were accessible to the reactants, resulting in good cata- before being activated in the presence of base. Finally, the

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10801
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 54 Proposed mechanisms in catalysis based on the structure of the Au25(SR)18 nanocluster. (a) Proposed mechanism for the chemo-selective
hydrogenation of α,β-unsaturated ketones to unsaturated alcohols catalysed by the Au25(SR)18 nanocluster. Reproduced from ref. 479 with per-
mission from Wiley-VCH, copyright 2010. (b) Catalytic cycle for the selective hydrogenation of benzalacetone to the unsaturated alcohol by the
Au25(SR)18 catalyst. Reproduced from ref. 480 with permission from American Chemical Society, copyright 2015. (c) Mechanistic understanding of
aldehyde hydrogenation with Au25(SR)18 nanoclusters. Reproduced from ref. 481 with permission from American Chemical Society, copyright 2015.
(d) Proposed mechanism for the ligand-on Au25(SR)18 catalysed semi-hydrogenation of terminal alkynes to alkenes using H2. Reproduced from ref.
97 with permission from American Chemical Society, copyright 2014.

alkene product is generated. Since deprotonation is essential Li et al. demonstrated the activity in the homocoupling of
for catalysis, internal alkynes cannot be catalyzed by the aryl iodides and excellent recyclability of the Au25(SR)18/CeO2
“ligand-on” Au25 nanoclusters due to no terminal H being nanocluster.410 The reaction was performed at 130 °C for 2
available in internal alkynes. days, which is critical for homogeneous Au25. Unsupported
5.3.5 C–C coupling. The C–C coupling reactions are widely Au25(SR)18 nanoclusters result in few products (12.1%) due to
used in the manufacturing industry for the organic synthesis of the decomposition behaviour, while the supported ones give
natural products, pharmaceuticals, polymers, etc.482 The use of 99.8% conversion.410 Besides, the heterogeneous catalysts
atomically precise nanoclusters in C–C coupling shows their show only a slight activity drop after 5 rounds of reaction.
great value in practical application. Jin and co-workers have DMF is beneficial to the transformation since it can protect
reviewed the C–C coupling reactions catalyzed by ultra-small the nanoclusters.410 The catalytic circle follows oxidative
nanoclusters.483 Here, we just focus on the Au25 catalyzed C–C addition and reductive elimination processes. The substrate is
coupling reactions, including the Ullmann-type homocoupling, absorbed on the gold surface via a weak I-Au bond to form
Sonogashira cross-coupling, and Suzuki coupling.154,410,484,485 [Au25](PhI)2, and subsequently the C–I bond can be activated
In 2016, the Jin group researched the Ullmann coupling by the exterior Au12 shell to generate [Au25Ph2]I2 intermediates.
reactions catalyzed by the [Au25(SNap)18]− nanocluster.154 In Finally, the biphenyl is yielded by decomposition.410
Ullmann coupling reactions catalyzed by CeO2 supported Au25 By employing the same oxide support, base, and solvent,
nanoclusters, aromatic ligands not only give rise to a higher the composite catalyst gives rise to excellent conversion
conversion, but also dramatically enhance the yield of the het- (96.1%) and satisfactory selectivity (88.1%) in the Sonogashira
erocoupling products.154 cross-coupling.484 The sterically unhindered facets on the

10802 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

nanocluster surface allow easy access for both reactants. The sequently decreases with further addition. It is speculated that
open facet with three external gold atoms, which is named the the pair-up behaviour of electrons results in the decrease of
A3 triangle, serves as the active site. The two phenyl rings from the odd electronic density.487
each reactant as well as two functional groups point toward 2-Nitrobenzonitrile is a good electron acceptor that is
three different gold atoms forming a special configuration to capable of grabbing electrons from Au25−.242 After the elec-
lower the adsorption energy, resulting in good catalytic tron-transfer process, the substrate forms N radicals and Au25−
activity.484 is oxidized to Au250. On the other hand, in the presence of
Only in the presence of ionic liquids can the TiO2 sup- NaBH4, neutral Au25 will be reduced negative again, and the
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

ported Au25(SR)18 catalyse Suzuki coupling.485 The interaction radical is further evolved and reduced.242 Two of the adjacent
between nanoclusters and imidazolium cations partially functional groups of the substrate are involved in the intra-
removes thiolate ligands (-SR) and “Au-SR” units, which pro- molecular catalysis. After the catalytic circle is completed, two
motes cross-coupling. The acidic hydrogen atom of imidazo- electrons are consumed, and the nanoclusters stay untouched,
lium cations exposes low-coordinated naked gold atoms on the functioning like a bridge transferring electrons from the cata-
surface to serve as active sites for catalytic reactions. The lyst to the substrate (Fig. 56).242
mechanism was further confirmed by UV-vis and MALDI-MS 5.3.7 Three-component coupling reaction. The Au25(SR)18
measurements.485 nanocluster is also efficient for the three-component coupling
5.3.6 Electron-transfer catalysis. Au25− can be oxidized into reaction (A3 reaction), which contains aldehydes, amines, and
Au250 and Au25+ by losing one or two electrons.486 The good alkynes.488 Under Ar atmosphere a wide range of substitutes
stability of each species proves that Au25 nanoclusters are with different functional groups could be catalyzed into corres-
efficient electron-transfer mediators when the redox couples ponding propargylamines in good to excellent yields. The
Au250/Au25− and Au25+/Au250 are employed. The transferred open sites without thiolates protecting them serve as active
electrons from Au25− have been captured by TEMPO+BF4− and sites. However, excess alkyne and amine are required to make
monitored by EPR (Fig. 55).241 TEMPO+ (2 equiv.) is needed to the best conversion since (i) alkynes compensate for the slow
make the EPR signal reach a plateau, implying that the elec- activation through the reaction of the alkyne at active sites,
tron consuming rate is 1/2. The linearly raised signal indicates and (ii) amines are necessary for enamine formation and
that the electron transfer process is consecutive. Compounds alkyne activation.488
that seize electrons to form radicals are very active in organic 5.3.8 Photocatalysis. The photocatalytic properties of Au25
reactions. As a result, Au25z nanoclusters could be utilized as nanoclusters have been investigated. Au25− is capable of facili-
radical initiators.241 tating the absorption and activation of molecular oxygen. The
Apart from salts and peroxides, the electrons from Au25− absorption gap of the Au25(SR)18 nanocluster (about 1.3 eV) is
could also be transferred to organic compounds. For instance, much larger than the energy of 1O2 (0.97 eV), which allows for
Au25− ends up neutral by losing only one electron to complex the oxygen oxidation. Kauffman et al. found that the electron-
PTZ-TCBQ (phenothiazine/tetrachloro-p-benzoquinone = 1/1).487 transfer process between Au25− and O2 is photo-mediated and
Upon the addition of complexes, the EPR signal reaches the depends on the relative energies of the Au25− LUMO and the
top when 1 equiv. of complexes is introduced, and sub- O2 electron-accepting level.457 Room light as the excitation
source is consistent with Kawasaki’s investigation of visible/
near-IR irradiation.407 After excitation, the molecular oxygen is
oxidized into the active singlet oxygen radical (Fig. 57a). Both
Au25− and Au250 can generate singlet oxygen and the latter can
be used to oxidize sulfoxide.407

Fig. 55 ESR signals of the reaction between Au25(SR)18− and


TEMPO+BF4− that monitor the electron-transfer process. Reproduced Fig. 56 The proposed mechanism for the Au251−/0 catalysed intra-
from ref. 241 with permission from American Chemical Society, copy- molecular cascade reaction. Reproduced from ref. 242 with permission
right 2011. from Nature Publishing Group, copyright 2013.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10803
View Article Online

Review Nanoscale

Visible-light-driven photocatalysis of Au25-modified TiO2 thione ligand) significantly increases the photocatalytic activity
has been investigated. It was established that the photoexcited while the degradation sharply decreases after calcination at
electrons from gold nanoclusters with different sizes can be 400 °C as the oxidized sulphur recombines with the photo-
injected into TiO2.489 When Au25 is employed, the maximum generated electrons and holes.492
internal quantum yield of 60% for the photoelectric conver- Besides TiO2, BaLa4Ti4O15 has also been used as a support
sion is achieved at 400–900 nm.489 Au25 loaded on TiO2 to load the Au25(SG)18 nanocluster, and the composite catalyst
enhances visible light adsorption and increases visible light exhibits enhanced photocatalytic activity (calculated as 2.6
photocatalytic activity by 1.6 times.490 Under visible light times) on water splitting relative to the catalysts loaded with
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

irradiation, the photogenerated electrons in Au25 can be larger gold nanoparticles (10–30 nm) via conventional
injected into the conduction band (CB) of anatase and rutile photodeposition.493
TiO2, thus facilitating electron–hole separation. Singlet oxygen 5.3.9 Electrocatalysis. Chen and co-workers investigated
(1O2), which is responsible for the decomposition of methyl gold nanoclusters with “ligands-on” for the electrochemical
orange, is produced by Au25(SR)18 after light adsorption reduction of O2 in fuel cells.494 Among the tested samples,
(Fig. 57b), indicating that a small loading amount of smaller nanoclusters exhibited much higher electrocatalytic
Au25(SR)19 can significantly enhance the photocatalytic activity activity for oxygen reduction. The good performance could be
of the composite in decomposing the dye.490 comparable to commercial Pt catalysts loaded on carbon. The
Under visible light irradiation, even 860 nm light, TiO2 sup- appreciable voltammetric currents detected in the ligand pas-
ported Au25(SG)18 is excited to induce the oxidation of phenol sivated gold nanoclusters denoted that the oxygen can readily
derivates and ferrocyanide as well as the reduction of Ag+, Cu2+ access the nanocluster surface with low impedance to inter-
and dissolved oxygen.491 Au25(SG)18 decorated TiO2 shows facial charge transfer.494
good catalytic activity toward Uniblue A degradation.492 For electrooxidation, GSH protected Au25 also showed excel-
Thermal treatment is crucial to the activation of composite cat- lent activity.363 The electrode modified by Au25(SG)18 could
alysts. Calcination at 250 °C ( partially removing the gluta- oxidize and sense ascorbic acid and dopamine in the range
between 0.71 to 44.4 mM. Due to the consequences of electro-
static attraction/repulsion between the charged Au25 nano-
clusters and the charged analytes, pH-dependent electro-
catalytic activity was observed.363
CO2 adsorption redistribution charge within the Au25 nano-
cluster could be used for the electrochemical conversion of
CO2 to CO in aqueous media.495 The Au25 nanocluster displays
enhanced electrocatalytic activity compared to the gold nano-
particles and bulk gold due to the anionic charge of Au25 that
promoted CO2 adsorption and unique reactive sites on the
surface of Au25, favoring the CvO bond activation and H
formation.495
The Lee group prepared an electro-catalyzer composed of
reduced graphene oxide and Au25 nanoclusters, and the thick-
ness of the Au25 layer on the graphene (from 0 to 15 layers)
could be precisely controlled according to the preparation con-
ditions.496 The electrocatalytic activities of these nano-
composites were tested by the reduction of [Ru(NH3)6]3+ (on
the surface region because of the fast electrode reaction,
Fig. 58a) and the oxygen reduction reaction in neutral media
(on or inside the film because of the kinetically controlled
process, Fig. 58b). The rotating disk electrode and rotating
ring-disk electrode voltammetry showed that the electro-
catalytic activity supported an efficient four-electron reduction
of oxygen.496

Fig. 57 (a) Generation of singlet oxygen by photoexcited Au25(SR)18


6 Composite materials
nanoclusters. Reproduced from ref. 407 with permission from American
The marriage of noble metal nanoclusters and other func-
Chemical Society, copyright 2014. (b) Mechanisms of the photocatalytic
activity of Au25(SR)18/TiO2 under visible light irradiation. Reproduced
tional materials provides a meaningful avenue for obtaining
from ref. 490 with permission from American Chemical Society, copy- multi-functional materials.497–500 First of all, the combination
right 2013. of different materials can join the functions of each material

10804 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

together. Furthermore, the integration of different functional


materials might hopefully generate novel materials with newly
intriguing physicochemical properties. Accordingly, multi-
functional composite materials are currently highly desired for
prompting the applications of metal nanoclusters in several
fields. In this section, the composite materials based on Au25
nanoclusters are summarized.
Pradeep and Muhammed demonstrated a rapid method for
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

synthesizing monodisperse water-soluble Au25(SG)18 nano-


clusters (with the assistance of (3-mercaptopropyl)trimethoxy-
silane (MPTS)) and then embedded the nanocluster a layer of
silica (Fig. 59a).501 Specifically, the MPTS-mediated Au25(SG)18
was subjected to hydrolysis with pH = 12 and T = 60 °C for
30 minutes to hydrolyze and condense the capped silanol
outside the Au25(SG)18 nanocluster, and subsequently the
Au25(SG)18@SiO2 composite materials were obtained.501 The
gold nanoclusters protected by a silica layer provided a new
way to prepare multi-functional composite materials due to
the modifiable, capped silica layer and inner nanoclusters.501
Kim et al. prepared the composite material composed of
the BSA–Au25 nanocluster and Ag-SiO2 nanoparticles
(Fig. 59b).502 The Ag-SiO2 core–shell nanoparticles were firstly
prepared and then the BSA–Au25 nanocluster was adsorbed on
the surface of this nanoparticle. A significant enhancement in
PL intensity was observed, compared to the composite
material with the free BSA–Au25 nanoclusters, which was deter-
mined by two competing processes of near-field enhancement
and fluorescence resonance energy transfer via the plasmonic
coupling. In addition, this “Au25-adsorbed Ag@SiO2” compo- Fig. 59 (a) Synthesis of the Au25@SiO2 composite material. Reproduced
site material has been applied as a highly sensitive and selec- from ref. 501 with permission from Wiley-VCH, copyright 2011. (b)
Synthesis of the Au25-adsorbed Ag@SiO2 composite material.
tive “turn-off” sensor for detecting Cu2+ ions.502
Reproduced from ref. 502 with permission from The Royal Society of
The interaction of ultrasmall nanoclusters with surfaces of Chemistry, copyright 2017. (c) Nanocluster size-conversion from Au25 to
graphene is significant for developing new promising appli- Au135 on the RGO. Reproduced from ref. 503 with permission from
cations of graphene materials. Ghosh et al. chemically mixed American Chemical Society, copyright 2014.
the Au25(SR)18 nanocluster with reduced graphene oxide (RGO)
in THF solution, and detected the increase in size of the nano-
cluster on the surface of RGO, due to the inter-reaction the nanoclusters with the surface. Finally, this transformation
between them (Fig. 59c).503 MALDI-MS results determined that could be dynamically tailored by controlling the concentration
the generated nanocluster was Au135(SR)57. Theoretical calcu- of the Au25 nanocluster and RGO in solution.503
lations illustrated that the transformation was influenced by The combination of semiconductors and metal nano-
interfacial fluctuations and the energy scales of interaction of clusters was of great significance to prepare new photovoltaic
or catalytic nanocomposite materials. Han et al. bonded the
Au25(SG)18, Au144(GS)60, and Au807(GS)163 nanoclusters on the
surface of the ZnO semiconductor.504 The prepared nano-
composite materials showed size-dependent photocatalytic
activities in reducing the azobenzene. It was suggested that the
azobenzene photoreduction was a proton-coupled electron
transfer (PCET) process. The photoreduction of azobenzene
exhibited clear thresholds in the size of the nanocluster (com-
pared the activity of Au25 with Au807) or the excitation intensity,
highlighting the multielectron/multiproton process.504
Samanta and co-workers researched the ultrafast transient
absorption of new composite materials comprising semi-
conductor quantum dots CdTe and Au25 nanoclusters.505 The
Fig. 58 Reactions at Au25-rGO-modified electrodes. (a) The reduction
process of [Ru(NH3)6]3+. (b) Oxygen reduction process. Reproduced photoinduced 2-way electron transfer between these two con-
from ref. 496 with permission from Wiley-VCH, copyright 2016. stituents of the nanocomposites was observed; i.e., the Au

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10805
View Article Online

Review Nanoscale

nanoclusters, which always act as electron donors, are able to selectivity of Au25(SG)18 within the ZIF-8 matrix in the catalytic
perform as electron acceptors when used as photosensitizers. reduction of 4-nitrophenol was observed, which is different
Interestingly, despite the big size of BSA ligands capping the from the Au25(SG)18 outside the MOFs, and could allow unrest-
Au25 nanocluster, the electron shuttling speed of BSA–Au25 is ricted access to the reactant. Additionally, the large variation
much larger than the Au25 nanoclusters protected by other of the PL of Au25(SG)18 inside or outside the ZIF-8 demon-
ligands such as GSH or histidine.505 strated the different host–guest interactions of these nano-
Due to the great water-soluble properties of the Au25 nano- cluster-MOF composite materials in different forms.508
cluster protected by GSH ligands, the Au25(SG)18 nanocluster The incorporation of atomically precise metal nanoclusters
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

can combine with the protein to create some water-soluble bio- with other functional materials has emerged as a promising
composites materials. Hartje et al. transferred the Au25(SG)18 subject. However, the current research on these composite
nanocluster into the pores of protein crystals.506 The size of materials is still in the primary stages. For future work, new
the pore in the host-protein crystal was about 13 nm, which methods for nanocluster-based composite materials with
was large enough for adsorption-coupled diffusion of the Au25 multi-function are highly desired, and more practical appli-
nanoclusters.506 Experimental and computational investi- cations are to be mapped out.
gations suggested that each large-pore protein crystal unit cell
can cover up to 29 nanoclusters. The pore diffusion coefficient
of the protein crystal was reduced significantly by 3 orders of 7 Inter-nanocluster reaction
magnitude to 3.4 × 10−10 cm2 s−1 when the concentration of
the nanocluster was too high, mainly because of the effects of Atomically precise analysis of the inter-reactions between
the pore occlusion.506 molecules enables chemists to understand the physico-
The frame and porous structure of metal–organic frame- chemical properties and detect the reflection process in detail.
works (MOFs) empower them to be the carriers to package or Reactions of nanoclusters with small molecular compounds or
surface-attach the small-size nanomaterials to generate multi- metal ions have been explored since the beginning of the
functional composite materials. Our group has selectively syn- nanocluster research. For example, the oxidizing H2O2 could
thesized the nanoclusters with confined size in the hollow/ decompose the Au25(S-PhC2H4)18 nanocluster, which is of
pores of the MOFs (i.e., ZIF-8, Zn(MeIm)2, where MeIm rep- great significance in synthesizing the single-Pt-doped
resents 2-methylimidazole).507 Very recently, Luo et al. Pt1Au24(S-PhC2H4)18 nanocluster.509 In addition, the existence
assembled the ZIF-8 and Au25(SG)18 nanocluster.508 With of NaX (X = Cl or Br) will decompose the [Au25(SR)18]0 nano-
different strategies (summarized in Fig. 60), the Au25(SG)18 cluster, while the transformation of the molecular valence
nanocluster could be selectively encapsulated into the state is detected from [Au25(SR)18]0 to [Au25(SR)18]1− when the
ZIF-8 matrix or just arranged along the outer surface of the NaX changes to TOAX.510 Capturing these reaction processes
ZIF-8 crystals, and the spatial distribution could be easily con- will not only help us fully understand the structures, compo-
trolled in each pattern.508 For comparison, the special size sitions, and properties of metal nanoclusters, but some
specific reactions can be designed based on these reaction
principles.
Muhammed and Pradeep reported the decomposition of
water-soluble Au25(SG)18 nanoclusters by the addition of a
series of metal ions such as Ag+, Fe3+, Cu2+, Ni2+, Cd2+, Zn2+,
and Sr2+.511 They found that there was no direct relationship
between the chemical activity and the electrochemical poten-
tial of these added ions.511 Li et al. introduced a novel size-
controlled synthesis of the Au44(S-PhC2H4)32 nanocluster by
reacting the Au25(S-PhC2H4)18 with redundant Cu2+ ions in an
oxidation-decomposition-recombination manner.512 The
obtained Au44 nanocluster exhibited enhanced catalytic activity
in reducing the 4-nitrophenol, compared with Au25 nano-
cluster, indicating the effect of the nanocluster size, structure,
and configuration on the catalytic properties.512 Wu and co-
workers researched the reactions between different-sized Au
nanoclusters and acetic acid.84 They found that small-sized
nanoclusters (e.g., Au25(SR)18 nanocluster) were easily decom-
posed in the presence of acetic acid, while larger nanoclusters
Fig. 60 Schematic illustration of the synthetic procedures for (a)
can withstand the acetic acid and maintain the composition
Au25(SG)18@ZIF-8 (assembling of Au25(SG)18 into ZIF-8) and (b)
Au25(SG)18/ZIF-8 (impregnating Au25(SG)18 on the surface of ZIF-8).
and construction.84 The reaction between the Au25(SR)18 nano-
Reproduced from ref. 508 with permission from Wiley-VCH, copyright cluster with massive PPh3 ligands will generate rod-like Au25
2017. nanoclusters co-protected by SR and PPh3 ligands (in the pres-

10806 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

ence of thiolate ligands), via the PPh3 protected Au13 and Au11 modify the optoelectronic properties of Au25(SBB)18 nano-
nanoclusters.513 This is essentially a ligand-exchange process, clusters, but also enhance the stability as well as their optical
which is similar to, for example, the generation of the responses. The capped CD molecule protects the Au25(SBB)18
Au28(TBBT)20 nanocluster by reacting the Au25(SR)18 nano- nanocluster like an umbrella, which helps the nanocluster to
cluster with excess TBBT ligands (where TBBT is 4-tert- resist redundant ligands, metal ions, and so on. The optimized
butylbenzenethiolate).53 structure of Au25(SBB)18@CD4 has been demonstrated by DFT
Yao et al. researched the amphiphilicity of the Au25(MHA)18 calculations and molecular modeling, in which the four CD
nanocluster by patching the hydrophilic Au25(MHA)18 with molecules are arranged in a symmetrical structure
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

hydrophobic cations (cetyltrimethylammonium ion, CTA+ in (Fig. 61b).146


this research).514 The acquired amphiphilic The Pradeep group carried out extensive work on the inter-
Au25(MHA)18@xCTA nanoclusters (where x = 6–9, half of a nanocluster reaction between the Au25(SR)18 nanocluster and
monolayer coverage on the templated Au25(MHA)18 nano- other metal nanoclusters, such as Ag44(SR)32,515 Ag25(SR)18,516
cluster) can exhibit diverse polarities (εr ranging from 4.15 to and Ir9(SR)6.116,517 Based on these reaction models, the active
41.25) in different solutions, which enable them to self-wrap sequence of each site on the crystal structure of each nano-
or completely extend (Fig. 61a). Consequently, the amphiphilic cluster has been determined and proved by DFT calcu-
Au25(MHA)18@xCTA nanoclusters can self-assemble into a bi- lations.116,518,519 The first example is the interparticle reaction
layer based film at the air-liquid interface, similar to the be- between the Au25(SR)18 and Ag44(SR)32 model nanoclusters.515
havior of a molecular amphiphile.514 Mass spectrometric measurements including MALDI-MS and
The stability of the nanocluster is of great significance due ESI-MS revealed that the interparticle reaction favors a metal-
to its decisive effect on the properties and practical appli- exchange process, in which the Ag atoms from the Ag44(SR)32
cations of nanoclusters. Mathew et al. partially encircled the nanocluster can exchange into the Au25 nanocluster (Fig. 62a).
Au25(SBB)18 with CD to yield Au25(SBB)18@CDx (x = 1–4, Synchronously, the ligands from the Ag44(SR)32 nanocluster
depicted in Fig. 61b).146 The CD enwrapping would not only can transfer onto the surface of the Au25(SR)18 nanocluster to
produce bi-ligand capped bi-metallic Au25−xAgx nano-
clusters.515 The interparticle metal/ligand-exchange process
can be separated into the metal-atom exchange as well as the
metal-thiolate fragment exchange. Up to 20 Ag atoms and 9
ligands from Ag44(SR)18 can be exchanged into/onto the
Au25(SR)18 nanocluster.515 It should be noted that the opposite
process, exchanging the Au atoms and ligands from the
Au25(SR)18 nanocluster to the Ag44(SR)32 nanocluster, also
occurs in this interparticle reaction. DFT calculations have
illustrated that the Au atoms from the Au25 nanocluster pre-
ferred to occupy the icosahedral core followed by the outer
staples as well as the inner dodecahedron positions, generat-
ing the bi-metallic Ag44−xAx(SR)32 nanocluster.515 In addition,
due to the energy lowering tendency of substituting Au and
Au-SR into/onto the Ag44(SR)32 nanocluster, accompanied by
the substitution of Ag and Ag-SR into the Au25(SR)18 nano-
cluster, the overall reaction is energetically favorable.515 After
this work, the structure- and topology-preserving chemical
reactions between two archetypal nanoclusters (i.e., Ag25(SR)18
and Au25(SR)18) were studied.516 First of all, similar to the
interparticle reaction between Au25(SR)18 and Ag44(SR)32, the
bi-metallic AuxAg25−x(SR)18 (x = 1–24) nanoclusters have been
observed through mass analysis (Fig. 62b), in which the value
of “x” can be controlled by changing the reactant compo-
sitions. It should be noted that the construction of the
M25(SR)18 nanocluster is maintained in this process.516
Significantly, an interesting mass peak centered at about m/z =
6279 Da with “−2” charge state has been captured in the early
Fig. 61 (a) Schematic illustration of the self-assembly process of stages of the process, which perfectly matches the dianionic
Au25(MHA)18@xCTA nanoclusters. Reproduced from ref. 514 with per-
adduct, i.e., (Ag25Au25(SR)36)2− (Fig. 62c). DFT calculations
mission from American Chemical Society, copyright 2015. (b) Schematic
representations of the Au25(SBB)18 nanocluster with different amounts
suggest that the metal-exchange and the ligand-exchange pro-
of CD molecules. Reproduced from ref. 146 with permission from cesses will occur through the formation of this adduct. After
American Chemical Society, copyright 2014. these exchange processes are complete, two bi-metallic

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10807
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Fig. 62 (a) Interparticle reaction between Au25(SR)18 and Ag44(SR)32. Reproduced from ref. 515 with permission from American Chemical Society,
copyright 2016. (b) Interparticle reaction between Au25(SR)18 and Ag25(SR)18. (c) ESI-MS and optimized structure of (Ag25Au25(SR)36)2−. Reproduced
from ref. 516 with permission from Nature Publishing Group, copyright 2016. (d) Interparticle reaction between Au25(SR)18 and Ir9(SR)6; the optimized
structure of Au25−xIrx(SR)18 nanoclusters (x = 1–3). Reproduced from ref. 517 with permission from American Chemical Society, copyright 2017.

M25(SR)18 nanoclusters split and generate the final metal- cluster (Fig. 62d).517 DFT calculations were performed to opti-
exchanged resultants.516 mize the favorable geometry of the Au22Ir3(SR)18 nanocluster,
Besides reacting with Ag44(SR)32 and Ag25(SR)18 nano- where the first Ir atom preferred to occupy the central position
clusters, the Au25(SR)18 nanocluster has also reacted with the of the Au25(SR)18 nanocluster, and the second/third Ir atom
Ir9(SR)6 nanocluster, generating the bi-metallic Au22(Ir)3 nano- goes through metal-exchange with the Au atoms on the surface

10808 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

of the icosahedral Ir1Au12 kernel, finally generating the center- clusters with atomically precise structures have been
doped Ir3Au22(SR)18 nanocluster (Fig. 62d).517 reported.332,342,520–537 Furthermore, no silver nanocluster
On the basis of the interparticle reaction between having identical geometric and electrical configurations to
Au25(SR)18 and Ag44(SR)32 nanoclusters, the active sites that are known Au nanoclusters were published until Bakr and co-
easily metal-exchanged can be determined. The Pradeep group workers reported the Ag25(SR)18 nanocluster, representing the
demonstrated the structure–activity correlations of three structural counterpart of the Au25(SR)18 nanoclusters
model bi-metallic nanoclusters—AuxAg25−x(SR)18, (Fig. 63).520 Specifically, similar to the Au25(SR)18 nanocluster,
AgxAu25−x(SR)18, and AuxAg44−x(SR)18.518 Note that these three the Ag25(SR)18 nanocluster also contains an icosahedral M13
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

nanoclusters were obtained by the metal-exchange process core (M = Au or Ag) and six M2(SR)3 dimeric staple motifs. The
from the parent Ag25(SR)18, Au25(SR)18, and Ag44(SR)32 nano- charge state of the Ag25(SR)18 nanocluster is the same as that
clusters. Experimental and theoretical results reveal that the of the Au25(SR)18 nanocluster, being “−1”.520 The PL intensity
Au atoms on the staple motifs in the AuxAg25−x(SR)18 and of Ag25(SR)18 is significantly larger than Au25(SR)18 (although
AuxAg44−x(SR)18 nanoclusters, i.e., M2(SR)3 or M2(SR)5 staple the absolute QY of Ag25(SR)18 is also too low to be detected
motifs, are more easily substituted relative to the inner sites credibly).520
(e.g., each site on the icosahedral kernel). A similar phenom- Ag44(SR)32 and Ag25(SR)18 nanoclusters have become the
enon was observed for Ag atoms in the AgxAu25−x(SR)18 nano- star molecules in the silver nanocluster field. Bootharaju et al.
cluster.518 Furthermore, Ag atoms on the icosahedral Ag13 found that these two silver nanoclusters could be inter-trans-
kernel of the Ag25(SR)18 nanocluster can be completely substi- formed using the ligand-exchange method.538 Specifically, the
tuted by Au atoms; however, Au atoms in M13 of the Au25(SR)18 transformation from the Ag25(SR)18 to the Ag44(SR)32 nano-
nanocluster cannot, indicating the more rigid M13 icosahedral cluster was achieved by reacting the former nanocluster with
configuration in Au25(SR)18 compared to that of the Ag25(SR)18 4-fluorobenzenethiolate. On the contrary, etching the
nanocluster. Moreover, the Au atoms in the Au1Ag24(SR)18 Ag44(SR)32 nanocluster with the HS-PhMe2 ligand would finally
nanocluster and subter-12 Au atoms in the AuxAg44−x(SR)32 generate Ag25(SR)18 nanocluster.538 It should be noted that the
nanocluster cannot be substituted by Ag atoms, proving that reaction time of the size reducing process (from Ag44(SR)32 to
the locations of these Au atoms are in the core or on the Ag25(SR)18) is pretty long compared to the size increasing
surface of the icosahedral M13 kernel, respectively.518 process. Combining these two reverse processes led to the
In the interparticle reaction between Au25(SR)18 and control of the icosahedral kernel in the hollow or nonhollow
Ag44(SR)32 nanoclusters, not only can the Au atoms in the state.538
Au25(SR)18 nanocluster be substituted by Ag atoms, but the Zheng and co-workers in situ synthesized the Pd/Pt central
metal-exchange process will take place on the Ag44(SR)32 nano- doped Ag25 nanoclusters—Pd1Ag24(SR)18 and Pt1Ag24(SR)18.539
cluster to generate the AuxAg44−x(SR)32 nanocluster (x = Because of the close atomic numbers, X-ray crystallographic
1–12).515 Based on these results, Pradeep and co-workers
demonstrated the manifestation of the shell structure of such
nanoclusters in the inter-nanocluster reactivity.519 The Ag-to-
Au substitution on the Ag44(SR)32 mono-metallic nanocluster
only occurred within the hollow Ag12 kernel and in a systema-
tic manner, generating the stable AuxAg44−x(SR)32 nanocluster
(x = 1–12) with a closed geometric as well as electronic shell
structure. Additionally, the substitution of Ag atoms in the
middle dodecahedral shell and the outermost mount sites to
Au is also possible; however, the obtained AuxAg44−x(SR)18
(where x is greater than 12) nanoclusters exhibited geometri-
cally/electronically open shells and an unexpected superatom-
nonsuperatom transition occurred. All of these results reveal
the structure–activity correlation in the interparticle reaction
chemistry.116,519

8 The “golden” Ag25 nanocluster


During the past two decades, the Au25(SR)18 nanocluster has
been thoroughly studied as the model system to explore the
structure–property correlations as well as the relevant practical
Fig. 63 (a) Atomically precise structure of the Ag25(SR)18 nanocluster
applications. Despite there being a battery of Au nanoclusters with different modes. (b) UV-vis, PL, and ESI mass spectra of the
that have been successfully synthesized and structurally deter- Ag25(SR)18 nanocluster. Reproduced from ref. 520 with permission from
mined, relatively small number of mono-metallic Ag nano- American Chemical Society, copyright 2015.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10809
View Article Online

Review Nanoscale

analysis can hardly distinguish Pd from Ag. Accordingly, DFT


calculations were performed to determine the accurate posi-
tion of the single Pd atom in the crystal structure of
Pd1Ag24(SR)18. The overall reaction was energetically favorable
when the Pd atom was in the central position of the M13 icosa-
hedral kernel, which is similar to the single Pt or Pd atom in
the center of the Pd1Au24(SR)18 and Pt1Au24(SR)18 nano-
clusters.509,540 In addition, the strong π⋯π interactions were
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

observed in Pd1/Pt1Ag24(SR)18 between 6 pairs of benzene rings


on the ligands. Furthermore, not was only the single Pt-doped
Pt1Ag24(SR)18 observed in the ESI-MS result, but the peak of
the Pt2Ag23(SR)18 nanocluster with relatively small proportion
compared to Pt1Ag24(SR)18 was detected (however, only one Pt
atom was determined in the crystal structure data of the Pt-
doped Ag25 nanocluster).539
It has been mentioned above that the PL intensity of the
Ag25(SR)18 nanocluster is still too weak to be observed with the
naked eye. Alloying is a broad-spectrum strategy to tailor the
physicochemical properties of mono-metallic nanoclusters,
which might hopefully solve this defect. Bakr and co-workers
metal-exchanged the templated Ag25(SR)18 nanocluster with Au
(PPh3)Cl complexes, which generated center-doped
Au1Ag24(SR)18 nanoclusters without changing the construction
of the Ag25 nanocluster (Fig. 64).541 Each optical absorption of
Au1Ag24(SR)18 exhibited a slight blue-shift compared with the
corresponding peak of the mono-metallic Ag25(SR)18 nano-
cluster. The PL boost (25-times enhancement), as the result of Fig. 65 (a) The metal-exchange process of the Pt1Ag24(SR)18 nano-
the exchange process, was an important finding. In addition, cluster with Au(I)SR complexes. Reproduced from ref. 542 with per-
the absorption onset of the Ag25(SR)18 nanocluster occurred at mission from Wiley-VCH, copyright 2016. (b) The proposed mechanism
for the Au doping of Pd1Ag24(SR)18 nanoclusters. (c) The proposed
850 nm, and was blue-shifted to 800 nm accompanied by the
mechanism for the Au doping of Pt1Ag24(SR)18 nanoclusters.
Au doping.541 Reproduced from ref. 543 with permission from The Royal Society of
Considering the stable occupation of Pt or Pd atoms in the Chemistry, copyright 2016.
central position of the Ag25(SR)18 nanocluster, we metal-
exchanged the Pt1Ag24(SR)18 nanocluster with Au(I)-SR com-
plexes, which generated the tri-metallic Pt1Au6.4Ag17.6(SR)18
nanocluster as the exchange result (Fig. 65).542 X-ray crystallo-
graphy revealed that the obtained tri-metallic nanocluster was
in a tristratified arrangement—[Pt(center)@Au/Ag-(shell)@Ag
(exterior)]. The Ag atoms in the exterior staple motifs, as well
as the Pt atom in the center of the icosahedral kernel, were
maintained, while only partial Ag atoms on the surface of the
Pt1Ag12 kernel were substituted by Au atoms. DFT calculations
demonstrated that the average distribution of Ag and Au atoms
on the surface of the icosahedral kernel followed the energy-
lowering tendency. However, the PL of Pt1Ag24(SR)18 nano-
cluster was likely to be quenched after the Au exchange
process.542 At the same time, Bakr and co-workers researched
the mechanism of the metal-exchange process by analyzing
the ESI-MS results of doping Pt1Ag24(SR)18 or Pd1Ag24(SR)18
nanoclusters with Au(I)PPh3Cl complexes (Fig. 65b and c).543
The central Pd atom would be substituted by Au atoms when
doping the Pd1Ag24(SR)18 nanocluster with Au(I)PPh3Cl, and
thus the Au1Ag24(SR)18 bi-metallic nanoclusters were generated
Fig. 64 Optical property comparison of Ag25 and its centrally doped
Au1Ag24 and Pd1Ag24 nanoclusters: (a) optical absorption, (b) PL spectra,
(Fig. 65b). However, Au(I)PPh3Cl doping would not exchange
(c) transient absorptions, and (d) fluorescence lifetimes. Reproduced the central Pt atom probably because of the more stable occu-
from ref. 541 with permission from Wiley-VCH, copyright 2016. pation of Pt atoms relative to Pd atoms. In this situation, the

10810 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

final productions were identified to be tri-metallic nanoclusters derived from the DPV measurements are consist-
Au1Pt1Ag23(SR)18 and Au2Pt1Ag22(SR)18 nanoclusters ent with those calculated from the optical absorption
(Fig. 65c).543 It should be noted that the tendency to maintain spectra.356
the central Pt atom in the metal-exchange process is identical It has been mentioned above that the PL intensity of the
in these two works.542,543 Ag25(SR)18 nanocluster is greater than the homologous
The target of single-atom engineering within a maintained Au25(SR)18 nanocluster.520 In addition, the PL QY has been
template has been achieved by combining the Ag25(SR)18 nano- further promoted through central-Au doping.541 In this
cluster and center-doped Au/Pt/Pd to obtain Au1Ag24(SR)18, context, the PL comparison of these nanoclusters can
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Pt1Ag24(SR)18, and Pd1Ag24(SR)18 nanoclusters.520,539,541 enlighten us as to how to tailor the fluorescence in the nano-
Investigating the differences in the structures and properties cluster field. Wu and co-workers investigated the crystal and
of these four homologous nanoclusters will not only help us solution photo-luminescence of M1Ag24(SR)18 nanoclusters
access the structure/composition-property correlations on the (M = Ag/Au/Pd/Pt).544 First of all, the crystal photo-luminescence
atomic level but also enable us to controllably synthesize the intensity sequence demonstrated that the mechanism of the
nanoclusters with desirable physicochemical properties. The fluorescence is a core-atom-directing charge transfer from the
heteroatom effects on the optical and electrochemical pro- capped ligands to the metallic kernels, which was further sup-
perties based on the structures of these four nanoclusters have ported by the theoretical natural population analysis charge
been investigated.356 Central doping with the single Au/Pt and the experimental Aginner–Sterminal bond length.544 In
atom significantly blue-shifts the optical absorption of the addition, all of these nanoclusters exhibited diverse PL intensi-
Ag25(SR)18 nanocluster, which illustrates the unique electronic ties in different solvents; i.e., showing the solvent effect on the
perturbation caused by the incorporated heteroatom. In con- nanocluster photo-luminescence. For instance, the PL QY of
trast, the substitution of the central Ag atom into Pd will not the Pt1Ag24(SR)18 nanocluster in the CH3CN was about 18.6%,
largely impact the optical absorption.356 Voltammetric resulting in a very strong contrast to that in CH2Cl2 with PL QY
measurements suggested that the sequence of O1-O2 potential of 0.2% (about 100-times enhancement); however, the emis-
spacing is Pd1Ag24 < Ag25 < Pt1Ag24 < Au1Ag24 (Fig. 66), match- sion wavelength of the Pt1Ag24(SR)18 nanocluster was almost
ing the stability of these four nanoclusters (e.g., enhanced maintained in different solutions.544
stability of Au1Ag24 compared with the Ag25 nanocluster). More Weerawardene and Aikens theoretically studied the origin
importantly, the HOMO–LUMO transition gaps of these four of the photoluminescence of the Ag25(SR)18 nanocluster.545
The Stokes shift of the Ag25(SH)18 nanocluster was calculated
as 0.37 eV, which slightly decreased when the capped ligands
were changed from -SH into the -SPhMe2. Note that this
decrease in the Stokes shift was based on the Ag25 nanocluster
being different from the Au25 nanocluster.545 DFT calculations
revealed that the luminescence of Ag25 arises from a HOMO–
LUMO transition, similar to the Au25 nanocluster, where the
core-based superatomic P and D orbitals are involved.
Furthermore, doping Ag25 with heteroatoms such as Au, Pt
and Pd does not affect the origin of the photoemission.545
Fully understanding the structure–property correlation
based on the successfully determined structures of templated
M1Ag24(SR)18 (M = Au/Ag/Pt/Pd) is a great opportunity for us to
controllably synthesize the functional nanoclusters with
enhanced physicochemical properties and wider application
prospects. For future work, deeper investigations into the
structure/composition-induced properties based on templated
nanoclusters (e.g., Ag25(SR)18 nanocluster) are also urgent, and
new synthetic principles for designing functional nanoclusters
are to be mapped out.

9 Alloying on Au25(SR)18
The last two decades have witnessed the rapid development of
the Au25(SR)18 nanoclusters in fundamental research as well as
Fig. 66 DPV spectra of Ag25 and its central doped nanoclusters.
practical applications.18,22,24,32,109,110,546,547 In the wake of the
Reproduced from ref. 356 with permission from Wiley-VCH, copyright extensive development of monometallic nanoclusters, in
2016. recent years, bi- or multimetallic nanoclusters have quickly

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10811
View Article Online

Review Nanoscale

Table 4 Alloyed nanoclusters (with crystal structures) based on the Au25(SR)18 template

Nanocluster Synthetic method Heteroatom Substitution sites Ref.

Pd1Au24(SR)18 (SR = S-PhC2H4) In situ synthesis Central position 540 and 568
Cd1Au24(SR)18 (SR = S-PhC2H4) Metal exchange Central position 562
Pt1Au24(SR)18 (SR = S-PhC2H4) In situ synthesis Central position 568
Cd1Au24(SR)18 (SR = S-PhC2H4) Anti-galvanic reduction Icosahedral shell 571
Hg1Au24(SR)18 (SR = S-PhC2H4) Anti-galvanic reduction Staple motif 572
AgxAu25−x(SR)18 (Ag ∼6.7) (SR = S-PhC2H4) In situ synthesis Icosahedral shell 575
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

AgxAu25−x(SR)18 (Ag ∼19.4) (SR = S-C6H11) Metal exchange Icosahedral shell/staple motif 576
AgxAu25−x(SR)18 (Ag ∼20.25) (SR = S-C6H11) In situ synthesis Icosahedral shell/staple motif 577
Hg1AgxAu24−x(SR)18 (Ag ∼7.2) (SR = S-PhC2H4) Metal exchange Ag-Icosahedral shell/Hg-Staple motif 590

emerged as promising nanomaterials.96,104,185 Due to the com- the second type of thiolate (i.e., C12H25SH) to substitute for the
plexity of the compositions and structures, alloying enables host thiolate (i.e., PhC2H4SH) and separated the obtained
the nanoclusters to possess enhanced physical and chemical mixed isomer by using HPLC. The isomer distribution demon-
properties, such as enhanced stability, optical, catalytic, strated that the ligand-exchange process preferred to occur at
electrochemical, magnetic properties and so the thiolates binding the icosahedral Pd1Au12 kernel.557 Bürgi
on.419,509,541,548–554 Au25(SR)18, as the molecular star in the and co-workers ligand-exchanged the Pd1Au24(SR)18 nano-
nanocluster field, has been extensively researched in view of cluster with BIANP.558 HPLC results illustrated that only one
the alloyed nanoclusters. Table 4 reviews the alloyed isomer was generated when only one BINAP was exchanged on
MxAu25−x(SR)18 nanoclusters (with atomically precise crystal the Pd1Au24; however, at least six kinds of isomers could be
structures) by comparing the synthetic methods, compo- detected when the second BINAP ligand was linked on the
sitions, substitution sites, and corresponding measurements. surface of the Pd1Au24 nanocluster.558
Similar to Pd, the Pt doping into the Au25(SR)18 template
9.1 Synthetic method generated the single-Pt doped Pt1Au24(SR)18 nanocluster. Note
So far, the synthetic strategies for alloying MxAu25−x(SR)18 that the single-Pt/Pd-doping into the M25(SR)18 template was
nanoclusters (where M represents the doping heteroatom) can also observed in the Ag system.539 The Jin group synthesized
be classified as follows: (i) in situ synthesis by reducing the the Au25(SR)18 nanocluster in the presence of H2PtCl6, and
mixture of Au-SR and M-SR complexes; (ii) doping the tem- found that the products were the mixture of mono-metallic
plated Au25(SR)18 nanocluster with heteroatom complexes. The Au25(SR)18 and bi-metallic Pt1Au24(SR)18 nanoclusters.509
early in situ synthesis of alloyed MxAu25−x(SR)18 nanoclusters Thanks to the decomposition of the Au25(SR)18 nanocluster via
was mostly based on the synthetic method of monometallic the oxidation by H2O2, as well as the high stability of the
Au25(SR)18 nanoclusters. A battery of measurements (such as Pt1Au24(SR)18 nanocluster, the high-pure Pt1Au24(SR)18 nano-
UV-vis, MALDI-MS, ESI-MS, CV, DPV and so on) was performed cluster was obtained. DFT calculations suggest that the only Pt
to determine whether the alloy process occurred or not and atom was located at the center of the metallic kernel. The
how many heteroatoms were alloyed in the Au25(SR)18 doped PtAu24 exhibited obvious differences in either optical or
nanocluster. catalytic properties. Specifically, the optical absorption of
Murray et al. synthesized the Pd doped Pd1Au24 nanocluster Pt1Au24 showed an 80 nm blue-shift compared with the Au25
by modifying the synthetic procedure of the Au25(SR)18 nano- nanocluster. In addition, enhanced catalytic activity for styrene
cluster.555 Specifically, they transformed the Au-SR precursor oxidation was observed on Pt1Au24 compared to the Au25
into the Au–Pd-SR mixture precursor, which resulted in the nanocluster.509
Pd-doped resultants. MALDI-MS results revealed that only one Alloying Ag into the Au25(SR)18 nanocluster is different
Pd atom could be doped into the Au25(SR)18.555 However, the from the process with Pt/Pd dopants. Only one Pt/Pd hetero-
heteroatom substitution site of this single Pd atom could not atom can be doped into the center of the Au25(SR)18 nano-
be determined because of the difficulty in obtaining crystals of cluster; however, more than one Au atom (always 5–7 in value,
the prepared Pd1Au24(SR)18 nanocluster. Murray and co- i.e., half the number of atoms on the surface of the icosahedral
workers also found that the Pd1Au25(SR)18 alloy nanocluster kernel) can be substituted by Ag. Note that the Ag cannot
had very different electrochemical properties from the mono- occupy the central position in the M13 kernel when prepared
metallic Au25(SR)18.555 Negishi and co-workers isolated the through the in situ synthetic method or metal-exchange
high-purity Pd1Au24(SR)18 alloy nanocluster using high-per- method. Negishi et al. synthesized a series of AgxAu25−x(SR)18
formance liquid chromatography (HPLC).556 A significant nanoclusters with different average x (x = 0–11) by altering the
enhancement in stability against decomposition and laser dis- initial molar ratio of HAuCl4 : AgNO3 in the synthesis.559 The
sociation was observed, compared to the Pd-doped nanocluster electronic structure of M25(SR)18 nanocluster is sensitive to the
with the mono-metallic Au25 nanocluster.556 Based on the degree of Ag doping, which is reflected in the optical
high-purity Pd1Au24(SR)18 nanocluster, they further exchanged spectra.559

10812 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

Transforming the Au-SR into Au–Ag-SR precursor to prepare the stability of CuxAu25−x(SR)18 is far weaker compared to the
the bi-metallic (AuAg)25(SR)18 nanocluster can also be applied Au25(SR)18 nanoclusters.189 Negishi et al. also found the same
to the preparation of water-soluble (AuAg)25 nanoclusters. Yao situation.561 In addition, they experimentally and theoretically
and co-workers prepared the Au17.6Ag7.4(SG)18 nanocluster by demonstrated that the Cu doping into the Au25(SR)18 nano-
reducing the Au–Ag-SG precursor at the feed mole ratio of Au/ cluster would alter the optical and redox potentials, signifi-
Ag = 3/1.347 The ratio of Au/Ag in the final compositions can cantly distort the geometric structure of Au25(SR)18, and reduce
be regulated by altering the feed ratio.347 The Xie group also the nanocluster stability in solution.561 In comparison, the
prepared a series of water-soluble AgxAu25−x nanoclusters pro- SePh protected (AuCu)25(SePh)18 nanoclusters are capable of
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

tected by mono- and bi-thiol ligands via the NaOH-mediated containing a maximum of nine Cu atoms.188 More interest-
NaBH4 reduction strategy.560 The Au/Ag ratio in the metallic ingly, the synthesized (AuCu)25(SePh)18 nanoclusters are stable
kernel, as well as the composition of the capped ligands can enough to remain steady in solution for a long time (Fig. 67b).
be tailored by changing the feeding precursors.560 The difference in stability might originate from the ligand
Cu doping is another special example. Gottlieb et al. syn- effect on the structures and properties of the Au–Cu alloy
thesized the CuxAu25−x(SR)18 nanocluster in situ and deter- nanoclusters.188
mined the Cu-doping atoms by the use of MALDI-MS.189 It was Another important strategy to prepare alloy M25(SR)18
suggested that a maximum of four Au atoms can be substi- nanoclusters is the metal-exchange method based on the pre-
tuted by Cu atoms. Interestingly, the de-alloying process was pared Au25(SR)18 nanocluster. Compared with the in situ syn-
observed with time, and the mono-metallic Au25(SR)18 nano- thetic method, the metal-exchange strategy might allow
cluster existed only to the end of this process (Fig. 67a); i.e., mechanistic insights into the synergistic effects at the
atomic level, which is of major significance for synthesizing
more functionalized nanoclusters. Our group devised a
broad-spectrum method to prepare alloy MxAu25−x(SR)18
nanoclusters through reacting the templated Au25(SR)18
nanocluster with thiolated complexes of CuII, AgI, CdII, and
HgII (Fig. 68).562 Experimental results suggest that the metal-
exchange process does not necessarily follow the metal
activity order, but is associated with the closed electron shell
(8e shell in the Au25 system) as well as the structural stability
of the nanocluster. From these principles, the Ag-, Cd-, Cu-,
and Hg-doped alloy nanoclusters with maintained M25(SR)18
framework were obtained; however, the Ni, Pd, and Pt metal
cannot be doped into the Au25(SR)18 nanocluster using this
method.562 Note that the in situ synthetic strategy can
prepare the single Pd- and Pt-doped M1Au24(SR)18 nano-
clusters (M = Pd/Pt).509,540 The Hg1Au24(SR)18 and
Cd1Au24(SR)18 were synthesized by the metal-exchange
method and were obtained in very high yields (about 100%);
the crystal structure of the center-doped Cd1Au24(SR)18 nano-
cluster was also successfully obtained.562 Xia and Wu alloyed
the Au25(SR)18 nanocluster by reacting it with Ag30(Capt)18 or
Cux(Capt)y nanoclusters, and detected the signals of Ag- or
Cu-doped MxAu25−x(SR)18 (M = Ag/Cu) alloy nanoclusters
with the maintained M25(SR)18 template.563 The Pradeep
group also reported a battery of alloying studies with these
so-called interparticle reactions.116,515–519
Following the periodic table of elements, highly active
metals can displace the low-active metal from its metal-salt,
known as galvanic reduction. Wu reported anti-galvanic
reduction (AGR) in the nanocluster field.564 The Au(0) in
Au25(SR)18 can reduce the less noble metal ions (for example,
reduce the Ag(I) into Ag(0), or Cu(I) into Cu(0)) and the
reduced metal ions in the natural state consist of the metallic
Fig. 67 (a) MALDI-MS of the de-alloying process of copper doped
kernel of the final MxAu25−x(SR)18 (M = Ag/Cu) nanoclusters.564
CuxAu25−x(SR)18 nanoclusters. Reproduced from ref. 189 with permission
from Wiley-VCH, copyright 2013. (b) Time-dependent MALDI mass
There are also some special examples synthesizing alloyed
spectra of CunAu25−n(SeR)18 nanoclusters. Reproduced from ref. 188 Au25(SR)18 nanoclusters through new procedures or with novel
with permission from The Royal Society of Chemistry, copyright 2013. final-compositions, which are discussed in section 9.2.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10813
View Article Online

Review Nanoscale
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Scheme 7 Referentially substituted sites of heteroatoms in the struc-


ture of Au25(SR)18 nanocluster (corresponding to ref. 540, 562, 568, 571,
572, 575, 576, 577 and 590). Color legend: orange sphere, Au; blue
sphere, Pd; green sphere, Pt, pink sphere, Cd; brown sphere, Hg; grey
sphere, Ag; red sphere, S. For clarity, the carbon and hydrogen atoms on
phosphine ligands are not shown.
Fig. 68 Metal exchange processes and the corresponding MALSI-MS
results. Reproduced from ref. 562 with permission from American
Chemical Society, copyright 2015. closed-shell species, and holes in the 1P HOMO manifold are
predicted on the [Pd1Au24(SR)18]z nanocluster with z = −1 ≤ z ≤
+3. However, it should be noted that the later Pt1Au24(SR)18
and Pd1Au24(SR)18 nanoclusters with crystal structures are all
9.2 Alloying site in neutral valence state.540,568 The elemental specificity of the
The variations in optical or electrochemical properties EXAFS spectroscopy makes it an ideal tool for studying the
accompanied by the alloying process will not indicate the accu- likely structure of the nanoclusters. Tsukuda and co-workers
rate sites of the doped heteroatoms in the final nanoclusters. proved the central position of the doped Pd heteroatom in the
The X-ray crystal structure is most trustworthy for grasping the Pd1Au24(SR)18 nanocluster.569 At the same time, Zhang and co-
atomically precise information of the alloyed nanoclusters (see workers reported the EXAFS results of Pt1Au24(SR)18 nano-
Scheme 7 for the preferentially substituted sites of hetero- cluster, which also indicated that the central Au atom in the
atoms in the structure of the Au25(SR)18 nanocluster, deter- Au25(SR)18 nanocluster had been substituted by Pt atoms to
mined by X-ray crystallography). In addition, theoretical calcu- generate the alloyed Pt1Au24(SR)18 nanocluster.570 Recently, the
lations and the XAS measurements enable us to predict the crystal structures of Pt1Au24(SR)18 and Pd1Au24(SR)18 nano-
most likely sites of the heteroatoms as well as the optimized clusters were reported, which completely confirmed the
structures of the alloy nanoclusters. central-doping of Pd and Pt into the Au25(SR)18
Despite not determining the crystal structure of the nanocluster.540,568
Pt1Au24(SR)18 and Pd1Au24(SR)18 nanoclusters (i.e., the doping One of the most attractive aspects of the nanoclusters is
site of the single Pt atom remained unknown) when they was their atomically precise structures. To date, the target of
firstly synthesized, DFT calculations were performed to predict single-atom engineering has been achieved based on the
the central position of the Pt/Pd atom.509 Walter and Moseler M1Ag24(SR)18 template (M = Au/Ag/Pt/Pd).520,539,541 However, as
predicted the central position of the single Pt in the for the Au25(SR)18 nanocluster, the central atom can be substi-
Pt1Au24(SR)18 nanocluster, which maintained the high sym- tuted by Pd or Pt,540,568 but cannot be replaced by Ag atom,
metry of the Au25(SR)18 nanocluster.565 The Jiang and the possibly because of the different electrode potential of the Ag
Häkkinen groups also predicted the optimized structure of compared to Au that makes the thermal stability of central-Ag-
M1Au24(SR)18 nanocluster.566,567 The di-anionic doped Ag1Au24 nanocluster much lower than the M25(SR)18
2−
[Pd1Au24(SR)18] nanocluster is a corresponding eight-electron nanoclusters with central-Au occupancy.

10814 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

Through the metal-exchange process, the single-Cd-doped


Au25(SR)18—Cd1Au24(SR)18 nanocluster was synthesized. The
crystal structure of this Cd1Au24(SR)18 nanocluster revealed
that the single Cd atom occupies the central position of the
alloy nanocluster, similar to the Pd/Pt-doped Au25(SR)18 nano-
cluster.562 In contrast, the single Cd heteroatom will be placed
on the surface of the icosahedral kernel if the templated
Au25(SR)18 nanocluster reacts with Cd(NO3)2 salts instead of
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Cd-SR complexes.571 Another study by the Wu group demon-


strated that the single Hg atom would occupy the staple motif
site to form a -SR-Au-SR-Hg-SR- dimeric motif.572 Bhat et al.
reported the formation of the Ir3Au22(SR)18 nanocluster.517
DFT calculations illustrated that the first Ir atom would substi-
tute the central Au atom, and the second and the third Ir
atoms would be positioned on the surface of the icosahedral
M13 kernel.517
Before the crystal structure of AgxAu25−x(SR)18 was reported,
Kauffman and co-workers theoretically predicted the possible
positions of the Ag heteroatoms in the alloyed (AuAg)25(SR)18
nanocluster.573 Specifically, the Ag atoms will be uniformly dis-
tributed on the surface of the M13 kernel, with the alloy nano-
clusters possessing the lowest energy.573 Tsukuda and co-
workers reached the same conclusion based on the theoretical
predictions as well as the XAFS results.574 In addition, they
Fig. 69 (a) X-ray crystal structure of the slightly doped AgxAu25−x(SR)18
further predicted that the doped Cu atoms would occupy the nanocluster, with the staple-motif metals being entirely Au. Reproduced
positions on the staple motifs.574 from ref. 575 with permission from American Chemical Society, copy-
Dass and co-workers reported the first crystal structure of right 2014. (b) X-ray crystal structure of the heavily doped
the AgxAu25−x(SR)18 nanocluster.575 The central atom and the AgxAu25−x(SR)18 nanocluster with the staple-motif metals being Au-SR
or Ag-SR. Reproduced from ref. 576 with permission from The Royal
metallic atoms in the six -SR-Au-SR-Au-SR- staple motifs are
Society of Chemistry, copyright 2016.
exclusively Au, i.e., with 100% gold occupancy (Fig. 69a). All of
the Ag atoms were selectively incorporated onto the 12 vertices
of the M13 icosahedral kernel. The final composition of the
AgxAu25−x(SR)18 nanocluster as determined by X-ray crystallo- between different metals. More importantly, the physico-
graphy was Au18.3Ag6.7(SC2H4Ph)18, which can be described as chemical properties of nanoclusters can hopefully be tailored
Au1@Au5.3Ag6.7@6 × [-SR-Au-SR-Au-SR-]. It is possible that through controllable doping, thus showing enhanced
because of the small proportion of Ag in the whole compo- efficiency in several application fields (such as optics, electro-
sition, the staple motif of this nanocluster is only composed of chemistry, and catalysis). By controlling the types, numbers
Au atoms.575 Li et al. reported the heavily doped and sites of the doped heteroatoms, the structure–property
AgxAu25−x(SR)18 nanocluster, in which the silver went from the correlations based on the multi-functional MxAu25−x(SR)18
kernel to the surface (Fig. 69b).576 This alloyed (AuAg)25(SR)18 have been fully investigated.540,568,571–577
nanocluster was synthesized by the metal-exchange method The heteroatom doping would have a significant effect on
from the Au23(SR)16 nanocluster. The average Ag atom in this the electronic structure as well as the optical properties.
heavily doped nanocluster was about 19. Jin and co-workers Negishi et al. found that the electronic structure of
reported the one-phase synthesis of alloyed AgxAu25−x(SR)18 AgxAu25−x(SR)18 could be continuously modulated by increas-
with x of about 21. To date, it is the most heavily Ag-doped ing the incorporation of Ag atoms.559 Specifically, continuous
Au25(SR)18 nanocluster with atomically precise structure.577 blue-shifts have been observed not only on the absorption
Interestingly, the heavy Ag-doping did not significantly alter peaks, but also on the photoexcitation spectra accompanied by
the electron–phonon coupling strength or the surface phonon the increasing number of the Ag doping. The decreasing
frequency.577 HOMO–LUMO transition gap has also been detected.559 Jin
and co-workers also theoretically demonstrated the increasing
9.3 Property regulation HOMO–LUMO transition gap with the Ag doping into the
Expanding the versatility of well-defined nanoclusters by Au25(SR)18 nanocluster.577 For example, the HOMO–LUMO gap
seeking distinctively physicochemical behavior is a relevant of 1.36 eV of the monometallic Au25(SR)18 nanocluster would
issue in designing functional nanomaterials. Alloying enables be significantly enhanced to 1.84 eV when 22 Au atoms were
the nanoclusters to exhibit distinctive architectures and com- substituted by Ag atoms (i.e., Au3Ag22(SR)18 nanocluster).577
positions due to the complexity of the combination modes Aikens and co-workers theoretically investigated the electronic

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10815
View Article Online

Review Nanoscale

structures of Au25−xAgx(SH)18 (x = 1, 2, 4, 6, 8, 10, 12) was suggested to be a result of the Jahn–Teller-like distortion
systems.578 First of all, doping the first Ag atom into the icosa- of the 6-electron [MAu24(SR)18]0 (with 1S21P4 superatom con-
hedral shell is more energetically favorable, and the other Ag figuration, M = /Pt).244 The Wu group investigated the mono-
atoms are designed to occupy the other positions on the mercury doping effect on the electrochemical/magnetic pro-
surface of the icosahedral kernel for simplifying the algorithm. perties of Au25(SR)18 nanocluster (Fig. 70c).572 The O1 and R1
Furthermore, DFT calculations suggested that the silver peaks in the DPV spectrum of the Au25(SR)18 nanocluster sig-
dopants tend to be dispersedly distributed on the icosahedral nificantly red-shifted when doping the Hg heteroatom into the
surface. Moreover, a blue-shift in the low-energy HOMO– staple motif. In addition, a slight reduction in the HOMO–
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

LUMO transition peak (1P → 1D) and the enhanced intensity LUMO transition gap derived from the DPV measurement was
of the absorption peak at 2.5 eV were observed as the number also observed. Specifically, the HOMO–LUMO gap of
of doping Ag atoms increased.578 Gonzalez et al. demonstrated Au25(S-PhC2H4)18 nanocluster was calculated as 1.63 eV, which
that the bonding mode of the regular s-type interaction in fell to 1.52 eV for the Hg1Au24(S-PhC2H4)18 nanocluster.572
Au25(SR)18 nanoclusters tended to exhibit more extended and Bi-metallic nanoclusters are attracting great interest due to
covalent modes owing to the generation of s- and p-type inter- their enhanced catalytic properties as a result of the synergetic
actions when the central Au atom was substituted by a group effect between their constituent metals. The Jin group reported
XIV elements, such as Si, Ge, Sn and Pb.579 the enhanced catalytic activity of the Pt1Au24(SR)18 nanocluster
The electrochemical properties are also sensitive to the compared with the monometallic Au25(SR)18 nanocluster in
doping effect. Lee and co-workers measured and compared the styrene oxidation (enhanced the conversation from 58.9% to
SWV spectra of Au25, Pd1Au24 and Pt1Au24 nanoclusters 90.8%).509 Xie et al. reported that the single Pd atom doping
(Fig. 70a).244 The HOMO–LUMO transition gap of the mono- remarkably improved the catalytic performance of the
metallic Au25(SR)18 nanocluster was about 1.32 eV, which was Au25(SR)18 nanocluster in the aerobic oxidation of benzyl
significantly reduced to 0.32 and 0.29 eV for Pd- and Pt-doped alcohol.580 Specifically, the conversion of the oxidation cata-
nanoclusters in the zero valence state. DFT calculations were lyzed by Au25@CNTs of 22% was enhanced to 74% when the
performed to confirm the reducing tendency (Fig. 70b), which Pd1Au24@CNTs was used as catalyst.580 Wu and co-workers
prepared the alloyed Au25Ag2(SR)18 nanocluster by collocating
two Ag atoms on the surface of the Au25(SR)18 nanocluster.581
The bonding Ag atoms triggered the hydrolysis of the 1,3-
diphenylprop-2-ynyl acetate of the Au25(SR)18 nanocluster from
the conversion of 11% to 52%.581 Lee and co-workers found
that the catalytic activity of the bi-metallic Pt1Au24(SR)18 nano-
cluster for the hydrogen production was significantly higher
compared to the previous catalysts.582 DFT calculations were
performed to simulate the hydrogen-production process.582,583
Negishi and co-workers found a remarkable enhancement in
the ligand-exchange reactivity on comparing the reaction time
and rate of Pd1Au24(SR)18 to the mono-metallic Au25(SR)18
nanocluster.584 The same phenomenon was also observed in
the Ag- and Cu-doped Au25(SR)18 nanoclusters.585
Zhou et al. analyzed the effect of doping on the photodyna-
mical properties based on the Au25(SR)18 template.586,587
Compared with the Pd1Au24(SR)18 nanocluster, the coupling
between the surface ligands and the metallic core was acceler-
ated by 30% in the case of the Pt1Au24(SR)18 nanocluster.586 In
addition, much more drastic changes have been observed in
the steady-state and the transient absorption of Pt1Au24(SR)18,
resulting from the more strongly mixed Au (5d and 6s) and Pt
(5d) orbitals, compared with the Pd dopants.586 In another
work, they also revealed the determining role of the valence-
Fig. 70 (a) SWVs spectra of Au25(SR)18 (black curve), Pd1Au24(SR)18 (red
electron number in the photodynamic properties of alloyed
curve), and Pt1Au24(SR)18 (blue curve) nanoclusters. (b) Electronic energy M1Au24(SR)18 nanoclusters (M = Cd/Hg/Pt/Pd).587 First of all,
levels of 6-electron or 8-electron systems and the corresponding Jahn– the doping of the Cd heteroatom enhanced the electron–
Teller-like distortion in the Pd1Au12 icosahedral kernel. Reproduced from phonon coupling while the Hg doping did not. Furthermore,
ref. 244 with permission from American Chemical Society, copyright
similar features were observed in the relaxation dynamics of
2015. (c) DPV spectra of different Au25 and Hg1Au24 nanoclusters. (d)
Simplified energy level and ESR signals of Au25(S-PhC2H4)180 and
the 8-electron [M1Au24(SR)18]0 (M = Hg/Cd) nanoclusters with
Hg1Au24(S-PhC2H4)180 nanoclusters. Reproduced from ref. 572 with per- the negatively charged Au25 nanocluster. In contrast, the non-
mission from American Chemical Society, copyright 2015. 8-electron M1Au24 (M = Pd/Pt) nanoclusters showed much

10816 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

shorter excited-state lifetimes, which resembled the case of the


neutral [Au25(SR)18]0 nanocluster.587
The Wu group investigated the main source of magnetism
in the neutral [Au25(SR)18]0 nanocluster by comparing the mag-
netic properties of the [Au25(SR)18]0 and alloyed
[M1Au24(SR)18]0 nanoclusters (M = Pt/Pd).568 EPR spectroscopy
revealed the diamagnetic properties of both [Pd1Au24(SR)18]0
and [Pt1Au24(SR)18]0 nanoclusters; however, the [Au25(SR)18]0 is
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

paramagnetic. These results indicate that the doping of Pt and


Pd turned off the magnetism of the Au25 nanocluster, which
was proposed to be the result of the electron configuration
conversion from the 7-electron to 6-electron configuration.568
They further experimentally demonstrated that the icosahedral
Au13 kernel was the major magnetism source for the neutral
[Au25(SR)18]0 nanocluster in comparison to the surface staple
motifs.568

9.4 Multi-metallic M25(SR)18 nanoclusters


Different properties of different heteroatoms enable them
occupy different positions in the Au25(SR)18 nanocluster tem-
plate. For example, the Pd/Pt heteroatom prefers to substitute
the central-Au atom in the icosahedral Au13 kernel.540,568 Hg
and Cu atoms tend to occupy the staple motif position.574 Ag Fig. 71 (a) Illustration of the two-step metal-exchange method for
synthesizing the tri-metallic Cd1/Hg1AgxAu24−x(SR)18 nanocluster. (b)
preferentially replaces Au at the surface of the M13 kernel.575
MALDI-MS results of the bi- or tri-metallic nanoclusters. Reproduced
Cd has been reported to substitute the Au atoms both in the from ref. 588 with permission from The Royal Society of Chemistry,
central site and on the shell site of the icosahedral copyright 2015.
kernel.562,571 Consequently, it seems promising to prepare
multi-metallic nanoclusters that contain three or even four
metals with the guidance of preferential doping-sites for each
heteroatom. Depending on the different preferential doping-sites of Hg
We reported the first example of the synthetic method of and Ag atoms, Yan et al. reported the synthesis and crystal
tri-metallic nanoclusters.588 The tri-metallic structure of Hg1AgxAu24−x(SR)18 nanocluster.590 The final com-
Cd1AgxAu24−x(SR)18 nanocluster was prepared through two position and construction of this tri-metallic nanocluster were
steps (Fig. 71a): (i) doping the Ag atoms onto the surface of the confirmed by SC-XRD and TDDFT calculations, revealing that
icosahedral Au13 kernel to obtain the bi-metallic the final alloy nanocluster was in a tri-stratified arrangement—
AgxAu25−x(SR)18 nanocluster; (ii) the substitution of the Au1@Ag7.2Au5.8@5*Au2(SR)3 + Au1Hg1(SR)3.590 It has been
central-Au atom by a Cd atom to prepare the tri-metallic reported that the single-Hg doping would reduce the HOMO–
Cd1AgxAu24−x(SR)18 nanocluster. MALDI-MS results suggested LUMO transition gap of Au25(SR)18 from 1.64 to 1.52 eV.572
that up to 6 Ag atoms can be doped in the final alloy nano- Additionally, further alloying the bi-metallic Hg1Au24(SR)18
cluster (Fig. 71b). The structure of the tri-metallic nanocluster nanocluster with Ag dopants would enlarge the HOMO–LUMO
follows the Cd1@(AuAg)12@6*Au2(SR)3 configuration. The vari- gap back to 1.63 eV (Fig. 72). Furthermore, accompanied by
ation in the optical spectra illustrated that the second and the alloying process from mono-metallic Au25 to bi-metallic
third foreign metals would significantly influence the elec- Hg1Au24, and then to the tri-metallic Hg1AgxAu24−x, the cata-
tronic structure of the monometallic Au25(SR)18 nanocluster.588 lytic activity was continuously enhanced from 23% to 48% and
Negishi and co-workers synthesized the mixture of 91%, respectively.590 It should be noted that the synthetic
Pd1CuxAu24−x(SR)18 nanocluster in situ by reducing the Au–Pd- methods and the crystal structures of the tri-metallic nano-
Cu-SR precursor in the presence of the TOAB.589 Cu doping clusters based on the Ag25(SR)18 template have also been
makes the CuxAu25−x(SR)18 alloy unstable, and it tends to de- reported, which were reviewed in section 8.542,543
alloy as time goes by.189,561 A single Pd doped into the Negishi and co-workers reported the first example of
Cu1Au24(SR)18 nanocluster would significantly enhance the synthesizing alloy nanoclusters containing four different
stability of this unstable nanocluster; that is, the tri-metallic metals.591 The stepwise metal-exchange method has been
Pd1Cu1Au23(SR)18 nanocluster is stable.589 However, different exploited to prepare these multi-metallic nanoclusters: (i) the
effects have been observed when the number of Cu substituted in situ synthesis of the bi-metallic Pd1Au24(SR)18 nanocluster;
is different. Specifically, the formation of the tri-metallic (ii) reacting the Pd1Au24(SR)18 with Ag-SR complexes to syn-
Pd1CuxAu24−x(SR)18 nanocluster is inhibited if x is equal to or thesize the tri-metallic Pd1AgxAu24−x(SR)18 nanocluster; (iii)
greater than 4.589 the metal-exchange process between the Pd1AgxAu24−x(SR)18

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10817
View Article Online

Review Nanoscale

erty investigation, and practical application of Au25 nano-


clusters. In addition, the composite materials combining the
Au25 and other functional materials provide a meaningful
avenue to obtain multi-functional nanomaterials, which have
been summarized in this review. Because of the homologous
structures and compositions of the Ag25(SR)18 and Au25(SR)18
nanoclusters, the development of Ag25(SR)18 was discussed.
Furthermore, we also summed up the synthetic strategies and
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

property variations of the alloyed MxAu25−x(SR)18 nanoclusters


for their importance not only in grasping the structure–prop-
Fig. 72 (a) Illustration of the two-step doping procedure to synthesize erty correlations, but also in guiding the controllable synthesis
the tri-metallic Hg1AgxAu24−x(SR)18 nanocluster. (b) Comparison of the of the nanoclusters with desirable structures/compositions
first oxidation and reduction potentials of mono-metallic Au25, bi-metal- and intriguing physicochemical properties.
lic Au24Hg1, Au25−xAgx, and tri-metallic Hg1AgAu24−x, respectively. Although there has been outstanding progress in the nano-
Reproduced from ref. 590 with permission from American Chemical
cluster field, further developments in fundamental investi-
Society, copyright 2015.
gations and practical applications are still facing many chal-
lenges. These issues also represent the innovations and break-
through points that are capable of making unprecedented pro-
nanocluster and the Cu-SR complexes to obtain the tetra-
gress with the nanoclusters. In the following, these challenges
metallic Pd1AgxCuyAu24−x–y(SR)18 nanocluster.591 In the second
(or breakthrough points) are discussed in brief, followed by
step, stability tests suggested that the most stable tri-metallic
the synthesis-structure–property-applications.
composition was the Pd1Ag4Au20(SR)18 nanocluster. EXAFS
analyses were performed on the tetra-metallic nanocluster to 10.1 Mechanism for the formation of Au25
determine the substituted positions of each metal. (i) The
Ultrasmall nanoclusters fill the gap between discrete atoms
single Pd atom occupied the central position of the icosa-
and plasmonic nanoparticles. The atomically precise struc-
hedral kernel; (ii) Ag atoms were dispersedly arranged on the
tures as well as their accurate packing modes make it possible
surface of the icosahedral M13 kernel; (iii) Cu atoms only sub-
to speculate the evolutionary mechanism from the small
stituted the Au atoms in the outermost M2(SR)3 staple motifs;
metallic complexes to monodisperse nanoclusters. The Xie
(iv) the rest of the Au atoms were distributed on the surface of
group carried out extensive studies on deducing the mecha-
the icosahedral M13 kernel and at the outermost staple motifs.
nism of the formation of Au nanoclusters.105,221–223 They pro-
Finally, optical spectroscopic and theoretical investigations
posed that the formation of the stable nanoclusters (for
demonstrated that the synergistic effects of multiple metals on
example, Au25 nanocluster) included two stages: (i) a size-
the electronic structures were additive in nature.591
increasing process from Au-SR complexes to polydisperse Au
It is significant to research the manner of heteroatom intro-
nanoclusters; (ii) a size-focusing process from polydisperse Au
duction in greater depth for its guidance in synthesizing alloy
nanoclusters to monodisperse nanoclusters.222,223 It should be
nanoclusters with controllable structure/composition and
noted that the free electron number increases with this
desirable properties. In the future, new approaches to prepar-
process.223
ing alloy nanoclusters should be further explored.
The aforementioned mechanism of the nanocluster for-
mation was proposed by analyzing the mass spectra results. To
date, the in situ analysis of the structure is being developed to
10 Conclusions and future catch a glimpse of the transformation of structures
perspective directly.592–595 In this regard, XAS measurement (i.e., operando
spectroscopy analysis) is perhaps one of the most useful
In the past few decades, explosive progress has been achieved approaches for grasping the formation mechanism of the
in researching the noble metal nanoclusters from the funda- nanoclusters.
mental studies to the practical applications. The atomically
precise structures, as well as the promising structure-depen- 10.2 Homologous structure, different metals
dent physicochemical properties, have enabled the nano- Attempts towards correlating the structures and properties in
clusters to draw extensive attention. To date, more than three the nanocluster field based on the homologous structures with
hundred nanoclusters with accurate structures or compo- different homo-components are limited because of the lack of
sitions have been reported. Among these nanoclusters, Au25 appropriate materials. The discovery of the homologous
has been regarded as the molecular star in the development of system will provide a platform for directly and unambiguously
the nanocluster, and is capable of overviewing the past, comparing the physicochemical properties of Ag and Au nano-
present and future of these promising nanomaterials. This clusters, thus enabling theoretical progress for fully investi-
review has highlighted the recent advances (most in the past gating the fundamental differences in their optical, electro-
decade) in material preparation, structure elucidation, prop- chemical, catalytic, and magnetic properties. The Ag25(SR)18

10818 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

nanocluster is the first (and the only until now) silver nano- The homo-thiolated Pt and Pd nanoclusters have not been
cluster with an exact analogue in the Au25(SR)18 nanocluster in reported mainly due to their different bonding modes with Au
terms of the size, composition, superatom electronic configur- or Ag metals. Compared to the Au and Ag atoms that contain
ation, and the crystal structure.520 The Ag25(SR)18 nanocluster one free valence s-electron (d10s1 electronic configuration), Pd
shares the same icosahedral M13 kernel and six dimeric -SR-M- metal has a closed d shell but no s-electrons (d10s0 electronic
SR-M-SR- staple motifs with its homologous Au25(SR)18 nano- configuration), while Pt metal has both d/s shells open (d 9s1).
cluster. Although the optical absorption of Ag25 is similar to Previous works demonstrated that the Pd or Pt atom prefers to
the Au25 nanocluster, just displaying some blue-shifts, the PL occupy the central position in the metallic
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

intensity of Ag25 is much stronger than the Au25 nano- kernel.509,539,540,544,551–553,597–600 More recently, Shen and
cluster.520 Different charge states endow the Au25(SR)18 nano- Mizuta reported the atomically precise structure of Pt5Ag22 bi-
cluster with diverse properties in optical absorption, magnet- metallic nanocluster, in which four alkynyl-platinum based
ism and so on. However, the Ag25(SR)18 nanocluster has only “crucifix motif” was observed.601 All of these research efforts
been observed for the “−1” charge state. DPV measurements of illustrate that the bonding modes between Pt (or Pd) with
Ag25(SR)18 suggested the existence of Ag25(SR)18 nanoclusters other metals or ligands are totally different with those of Au
with different charge-states, such as the neutral state.356 and Ag metals. Liu et al. reported the synthetic method of Pt
Future works should identify and research the Ag25(SR)18 nano- nanocluster with the mass determining as about 8000 Da.602
cluster with more charge states. Charge transfer from Pt to S of the thiolate ligand in the nano-
Hayton and co-workers synthesized the Cu nanocluster cluster has been clarified. The mass of 8000 Da suggested that
which contained 25 Cu atoms, 12 PPh3 and 22 H ligands, and the metal number in this platinum nanocluster is around 25.
the molecular formula is Cu25H22(PPh3)12.596 The crystal struc- More efforts should be made on the structural determination
ture of the Cu25 contains a Cu-based icosahedral M13 kernel, of Pt/Pd-based thiolated nanoclusters, or at least on the
same as the M13 kernel observed in the Ag25(SR)18 and further understanding the bonding mode between Pt/Pd with
Au25(SR)18 nanoclusters (Fig. 73). The feature of partial Cu(0) thiolates. Such information is of significance to guide us to
character in the Cu13 kernel has been confirmed.596 Although prepare more nanoclusters with novel compositions and
the Cu25 nanocluster embodies different constructions in the structures.
staple motif with Au25(SR)18 and Ag25(SR)18 nanoclusters, the
discovery of the Cu25 nanocluster is also significant for 10.3 New strategy to prepare alloy MxAu25 nanoclusters
directing the preparation of new metal nanoclusters with such Compared with the single-component Au25(SR)18 nanoclusters,
active metals. Furthermore, the Cu25 nanoclusters exhibit a alloyed MxAu25−x(SR)18 nanoclusters often exhibit unique
unique opportunity to study the reactivity of Cu nanoclusters, physicochemical properties due to the synergistic effect
such as more structure–property correlations.596 between different metals. These properties create more oppor-
tunities in applications of catalysis, biosensing, surface pat-
terning, molecular electronics, and so on.96,104 The previously
reported preparations of alloyed MxAu25−x(SR)18 nanoclusters
can mainly be classified into two types: (i) in situ synthesis by
reducing the M–Au-SR mixture complexes; (ii) templated
metal-exchange process on as-prepared Au25(SR)18 nano-
clusters. Recently, the Pradeep group reported the interparticle
reaction between Au25(SR)18 and other nanoclusters such as Ag
and Ir nanoclusters.116,515–517 The prepared AgxAu25−x(SR)18 or
Ir3Au22(SR)18 nanoclusters were characterized by mass spec-
troscopy and the optimized structures of these alloy nano-
clusters were predicted by DFT calculations.516,517 Xia and Wu
also reported the interparticle reaction between the Au25 nano-
cluster and the Ag30 nanocluster, Ag nanoparticles, or Cu
nanoparticles, which generated AgxAu25−x(SR)18 or
CuxAu25−x(SR)18 nanoclusters, respectively.563 The increasingly
popular interparticle reaction strategy can be seen as the third
way to prepare MxAu25−x(SR)18 alloy nanoclusters.
To date, alloying based on the template of Au25(SR)18 has
been achieved on Ag, Pt, Pd, Cu, Hg, Cd, Ir metals. DFT calcu-
lations have also been performed on the heteroatoms of Ge,
Fig. 73 Crystal structure of the Cu25H22(PPh3)12 nanocluster. (a) Whole
Sn, Pb, Zn, Ni, Fe, Co, Rh and so on. The reaction conditions
structure with carbon atoms depicted in wireframe. (b) Side view
showing only the Cu and P atoms. (c) H positions on the structure pre-
including solvent, temperature, oxygen, reaction time, stirring
dicted by DFT calculations. Reproduced from ref. 596 with permission speed are all significant in preparing the alloy nanoclusters
from American Chemical Society, copyright 2015. with different composition and stability. Future studies are

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10819
View Article Online

Review Nanoscale

expected to provide more insight into the alloying effect on the 10.6 Precise structure of water-soluble Au25
electronic structures as well as the physicochemical properties. SC-XRD is the most credible means of capturing the atomically
In addition, preferential doping sites of each heteroatom precise structure, binding mode, and packing order of a mole-
should be further analyzed for guiding the synthetic method. cule. Several crystal structures of Au25(SR)18 with different
Furthermore, alloying the homo-gold Au25(SR)18 nanocluster capped ligands or charge states have been reported since the
with new types of metals (better to be Ni, Zn and Fe active first report of the accurate structure of
metals) is necessary for exploiting novel functional [Au25(S-PhC2H4)18][TOA].51,52 It should be noted that all of
nanoclusters.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

these nanoclusters are oil-soluble. Meanwhile, water-soluble


Au25 nanoclusters (such as Au25(SG)18 and BSA–Au25) have
10.4 Mechanism of the metal-exchange process been prepared and applied in the fields of sensors or
biology.135,136,214 However, the atomically precise structures of
The atomic precision of nanoclusters provides a platform to these water-soluble nanoclusters cannot be obtained due to
structurally analyze the evolution of the the difficulty in crystallization. It is by default that the struc-
nanocluster.47,541,598,603–611 The single-heteroatom in the alloy- tures and bonding modes of these water-soluble Au25 nano-
ing process hopefully acts as a “tracing agent” for the metal- clusters are the same as the oil-soluble ones. However, the
exchange process, just like the “isotopic tracer method” in complexity of the aqueous ligands such as BSA makes it poss-
organic reactions.541,598,603,605 For example, Bakr and co- ible that the metal–ligands binding modes and the total struc-
workers have proposed the mechanism of substituting the tures of water-soluble Au25 nanocluster are different from what
central-Ag atom with the doping Au atom in the synthesis of we considered previously. New breakthroughs are urgently
the Au1Ag24(SR)18 bi-metallic nanocluster: (i) the Au+ ion is needed for determining the atomically accurate structures of
firstly reduced into Au0 and replace a Ag atom on the icosahe- these water-soluble nanoclusters, which are of significance for
dral Ag13 kernel; (ii) Au diffuses into the core and pushes the the fundamental research in the nanocluster field.
central Ag atom to the icosahedron surface because of the
thermodynamically favorable process.541 The mechanism of
Ag- or Cu-doping into the Au25(SR)18 nanocluster has also 10.7 Origin of the properties
been proposed as the direct substitution of the surface Au Strong quantum size effects in the nanocluster field have been
atoms with the introduced heteroatoms.562,573–575 However, no manifested in their physicochemical properties such as mul-
experimental evidence has been presented so far. tiple absorption bands due to one-electron transitions, discrete
The preferential exchange sites of different heteroatoms energy levels of electrons, enhanced photo-luminescence, mag-
have been theoretically and experimentally netism, catalytic reactivity, and nonlinear optical properties, to
investigated.96,562,566–573 Despite reports on several crystal name a few. However, the origins of some intriguing properties
structures of alloyed MxAu25−x(SR)18 nanoclusters, the exact of Au nanoclusters are still not clear for now, despite reports
mechanism of doping is not clear. Doping on the icosahedral from a battery of related works.274,279–281,283–285 As for the
M13 surface or the staple motifs is easy to understand to some luminescence, it was suggested that the kernel-surface coup-
extent because the dissociative heteroatom complexes have the ling in the nanocluster is responsible for the photo-lumine-
change to contact these sites. In contrast, the central doping is scence; however, the correlations between them remains to be
more complex. Stepwise processes certainly exist, and how to determined.
monitor these processes remains challenging. Further explora- The available crystal structure of the Au25(SR)18 nanocluster
tion of the metal-exchange process will open up new areas in permits the correlation of its catalytic reactivity with the
nanocluster science. atomic structure. Li and Jin suspected the “A3 site” as the
active site in catalysis.97 Our group proposed the mechanism
for the Au25(SR)18 nanocluster in catalyzing styrene oxidation;
10.5 More composite materials based on Au25 612
however, the actual catalytic sites remain unknown.
The marriage of Au25 nanocluster with graphene,503 semi- Theoretical methods (such as DFT, TDDFT calculations) are
conductors,504 and MOFs508 poses a promising approach to capable of simulating the likely mechanism of the catalysis,
preparing composite materials, which may hopefully display which will hopefully find their power in mechanistic studies.
novel optical, magnetic, catalytic, electrochemical properties, The origin of the paramagnetism of Au25(SR)18 with zero
and so on. The combination of two functional materials will charge state is also a mystery. Previous theoretical calculations
not only maintain the physicochemical properties of the build- revealed that the icosahedral Au13 kernel was the major mag-
ing blocks but can also enable some new properties due to the netism source.234 Wu and co-workers also suggested that the
interaction between them. The development of the composite Au13 kernel was the major magnetism location for
materials based on Au25 is still in the initial stages. More [Au25(SR)18]0 nanocluster in comparison with the surface
efforts should be made on the combination of the Au25 nano- motif.568 It was suggested the paramagnetism/diamagnetism
cluster with diversely functional nanomaterials, which will is related to the free electron number, while the specific influ-
lead to the discovery of more joint materials with amazing encing modes and other influencing factors should be further
behaviors. investigated. The mysterious structure–property correlations of

10820 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

giant nanoclusters will eventually be unveiled through such


efforts.

10.8 Ligand engineering to enhance the Au25 properties


The significance of the ligand effect has been evaluated in
several aspects such as controllable synthesis and property
regulation.54,332,536,537,613–616 Through engineering the electro-
negativity, steric hindrance, or other properties of the pro-
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

tected ligands, researchers can tailor the optical,139,279–281 cata-


lytic,154 and chiral properties137,138,172 of the Au25 nano-
clusters. Important questions that naturally arise are as
follows: What is the driving force for these intriguing ligand
effects? How does the effect influence the integrated properties
of the nanocluster? Can we prepare the multi-functional nano-
clusters by just combining the different functional ligands on
the surface of the nanocluster? The future application of the
nanoclusters urge us to map out more details of the ligand’s
role/effect, and thus more nanoclusters with enhanced func-
tions can be prepared using the rules.

10.9 Theoretical investigation on the Au25 nanocluster


Theoretically investigating the nanocluster will provide
fundamental insight into the structural
evolution47,107,108,247–252,606–610,617 and the mechanisms of the
optical52,248 and magnetic234 properties. At the same time that
the crystal structure of Au25(SR)18 was reported, Akola et al.
reported the optimized structure of the Au25(S-CH3)18 nano-
cluster (matching the construction of experimental structure)
on the basis of the first principles calculations.247 DFT and
TDDFT are current primary methods to simulate the electronic
structures as well as diverse properties of the nano-
clusters.107,108,617 Nonetheless, the computed results still
cannot match the experimental data perfectly. For instance,
red-shifts are always observed on comparing the UV-vis absorp-
tion peaks predicted by DFT calculations with the experi-
Fig. 74 (a) Comparison of the size exclusive/gel filtration chromato-
mental ones.194 Besides, the increasing demands of theoreti-
graphy and the glomerular filtration of the sub-nanometer gold nano-
cally analyzing the complex constructions and properties of clusters. (b) X-ray images of mice after being intravenously injected with
nanoclusters require greater progress in the computational nanoclusters of different sizes. (c) Comparison of the renal clearance
science. As a result, the further development of DFT calcu- efficiencies of different-sized nanoclusters or nanoparticles.
lations or the formulation of a more powerfully computational Reproduced from ref. 98 with permission from Nature Publishing Group,
copyright 2017.
tool is the order of the day.

10.10 Details in biological processes


Au nanoclusters have received enormous attention for biologi- (Fig. 74a).98 The renal clearance of Au nanoclusters increases
cal applications.1,37,98,99,102,111,618,619 BSA–Au25 nanoclusters or exponentially with increasing the number of the atoms in a
Au25(SG)18 nanoclusters have always been for bioimaging, bio- nanocluster (from Au10 to Au15, Au18 and Au25 nanoclusters,
labeling, or biomedical applications.98,101,102 It is known that Fig. 74b and c). The working mode in glomerular filtration is
the nanoparticles will be retained by glomerular filtration no longer the pore size-dependent filtration when in the sub-
barrier if their sizes are larger than 6–8 nm.620 Nanoclusters nanometer region, and the novel inverse size dependency
always possess a small-sized metallic kernel of 1–3 nm, which comes from the more easily physically retained feature of the
means that they can be metabolized out of the body smoothly. nanoclusters by the endothelial glycocalyx of the glomerulus.98
Zheng and co-workers found that the glomerulus behaved Future studies are expected to explore more biological appli-
as an atomically precise “bandpass” barrier to slow down the cations of these water-soluble nanoclusters based on their
renal clearance of the water-soluble nanoclusters (like size promising characteristics of low toxicity, biocompatibility,
exclusive chromatography by gel filtration chromatography), high photo-luminescence, photostability, atomic precision,
which is different from its usual function as a “size cut-off” slit bioselective absorption, and high permeability. In addition,

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10821
View Article Online

Review Nanoscale

new technologies are to be performed to analyze the biological 14 M. Grzelczak, J. Vermant, E. M. Furst and L. M. Liz-
processes in detail. Marzán, ACS Nano, 2010, 4, 3591–3605.
In summary, the nanocluster ship has been continuously 15 J. Kao, K. Thorkelsson, P. Bai, B. J. Rancatore and T. Xu,
expanding, due to the special sizes (filling the gap between Chem. Soc. Rev., 2013, 42, 2654–2678.
discrete atoms and plasmonic nanoparticles), atomically 16 K. Zhou and Y. Li, Angew. Chem., Int. Ed., 2012, 51, 602–
precise structures, intriguing physicochemical properties, and 613.
promising applications (see ref. 18–27). The Au25 nanocluster, 17 S. Yamazoe, K. Koyasu and T. Tsukuda, Acc. Chem. Res.,
as the captain of the great nanocluster ship, is capable of 2014, 47, 816–824.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

reviewing the history and guiding the future of the nano- 18 R. Jin, Nanoscale, 2015, 7, 1549–1565.
clusters. Accompanied by the further developments and wide 19 W. Kurashige, Y. Niihori, S. Sharma and Y. Negishi, Coord.
applications of Au25, a bright future lies ahead for Chem. Rev., 2016, 238, 320–250.
nanoclusters. 20 R. Jin, C. Zeng, M. Zhou and Y. Chen, Chem. Rev., 2016,
116, 10346–10413.
21 I. Chakraborty and T. Pradeep, Chem. Rev., 2017, 117,
Conflicts of interest 8208–8271.
22 A. Fernando, K. L. D. M. Weerawardene, N. V. Karimova
There are no conflicts to declare. and C. M. Aikens, Chem. Rev., 2015, 115, 6112–6216.
23 A. Cantelli, G. Guidetti, J. Manzi, V. Caponetti and
M. Montalti, Eur. J. Inorg. Chem., 2017, 5068–5084.
Acknowledgements 24 C. Zeng and R. Jin, Chem. – Asian J., 2017, 12, 1839–1850.
We acknowledge the financial support by NSFC (21372006, 25 G. Li and R. Jin, Acc. Chem. Res., 2013, 46, 1749–1758.
U1532141& 21631001), the Ministry of Education, the 26 H. Qian, M. Zhu, Z. Wu and R. Jin, Acc. Chem. Res., 2012,
Education Department of Anhui Province, 211 Project of Anhui 45, 1470–1479.
University. 27 S. H. Yau, O. Varnavski and T. Goodson III, Acc. Chem.
Res., 2013, 46, 1506–1516.
28 R. Jin, S. Zhao, Y. Xing and R. Jin, CrystEngComm, 2016,
Notes and references 18, 3996–4005.
29 T. Higaki, C. Zeng, Y. Chen, E. Hussain and R. Jin,
1 N. L. Rosi and C. A. Mirkin, Chem. Rev., 2005, 105, 1547– CrystEngComm, 2016, 18, 6979–6986.
1562. 30 Y. Zhu, R. Jin and Y. Sun, Catalysts, 2011, 1, 3–17.
2 D. V. Talapin, J. S. Lee, M. V. Kovalenko and 31 J. Fang, B. Zhang, Q. Yao, Y. Yang, J. Xie and N. Yan,
E. V. Shevchenko, Chem. Rev., 2010, 110, 389–458. Coord. Chem. Rev., 2016, 322, 1–29.
3 A. K. Gupta and M. Gupta, Biomaterials, 2005, 26, 3995– 32 A. Mathew and T. Pradeep, Part. Part. Syst. Charact., 2014,
4021. 31, 1017–1053.
4 R. J. Moon, A. Martini, J. Nairn, J. Simonsen and 33 S. Zhao, R. Jin and R. Jin, ACS Energy Lett., 2018, 3, 452–
J. Youngblood, Chem. Soc. Rev., 2011, 40, 3941–3994. 462.
5 P. K. Jain, X. Huang, I. H. El-Sayed and M. A. El-Sayed, Acc. 34 L. Nie, X. Xiao and H. Yang, J. Nanosci. Nanotechnol.,
Chem. Res., 2008, 12, 1578–1586. 2016, 16, 8164–8175.
6 E. Katz and I. Willner, Angew. Chem., Int. Ed., 2004, 43, 35 P. Yu, X. Wen, Y. Toh, X. Ma and J. Tang, Part. Part. Syst.
6042–6108. Charact., 2015, 32, 142–163.
7 M. E. Stewart, C. R. Anderton, L. B. Thompson, J. Maria, 36 T. Udayabhaskararao and T. Pradeep, J. Phys. Chem. Lett.,
S. K. Gray, J. A. Rogers and R. G. Nuzzo, Chem. Rev., 2008, 2013, 4, 1553–1564.
108, 494–521. 37 Y. Tao, M. Li, J. Ren and X. Qu, Chem. Soc. Rev., 2015, 44,
8 R. A. Petros and J. M. DeSimone, Nat. Rev. Drug Discovery, 8636–8663.
2010, 9, 615–627. 38 N. Goswami, Q. Yao, Z. Luo, J. Li, T. Chen and J. Xie,
9 C. J. Murphy, A. M. Gole, J. W. Stone, P. N. Sisco, J. Phys. Chem. Lett., 2016, 7, 962–975.
A. M. Alkilany, E. C. Goldsmith and S. C. Baxter, Acc. 39 P. Liu, R. Qin, G. Fu and N. Zheng, J. Am. Chem. Soc.,
Chem. Res., 2008, 41, 1721–1730. 2017, 139, 2122–2131.
10 D. A. Giljohann, D. S. Seferos, W. L. Daniel, 40 X. Liu and D. Astruc, Coord. Chem. Rev., 2018, 359, 112–
M. D. Massich, P. C. Patel and C. A. Mirkin, Angew. Chem., 126.
Int. Ed., 2010, 49, 3280–3294. 41 L. Zhang and E. Wang, Nano Today, 2014, 9,
11 A. R. Tao, S. Habas and P. Yang, Small, 2008, 4, 310–325. 132–157.
12 J. M. Slocik and R. R. Naik, Chem. Soc. Rev., 2010, 39, 42 P. D. Jadzinsky, G. Calero, C. J. Ackerson, D. A. Bushnell
3454–3463. and R. D. Kornberg, Science, 2007, 318, 430–433.
13 M. Grzelczak, J. Pérez-Juste, P. Mulvaney and L. M. Liz- 43 A. Das, C. Liu, H. Y. Byun, K. Nobusada, S. Zhao, N. Rosi
Marzán, Chem. Soc. Rev., 2008, 37, 1783–1791. and R. Jin, Angew. Chem., Int. Ed., 2015, 54, 3140–3144.

10822 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

44 S. Chen, S. Wang, J. Zhong, Y. Song, J. Zhang, H. Sheng, 67 C. Zeng, Y. Chen, C. Liu, K. Nobusada, N. L. Rosi and
Y. Pei and M. Zhu, Angew. Chem., Int. Ed., 2015, 54, 3145– R. Jin, Sci. Adv., 2015, 1, e1500425.
3149. 68 H. Dong, L. Liao and Z. Wu, J. Phys. Chem. Lett., 2017, 8,
45 C. Zeng, C. Liu, Y. Chen, N. L. Rosi and R. Jin, J. Am. 5338–5343.
Chem. Soc., 2014, 136, 11922–11925. 69 C. Zeng, Y. Chen, K. Iida, K. Nobusada, K. Kirschbaum,
46 T. C. Jones, L. Sementa, M. Stener, K. J. Gagnon, K. J. Lambright and R. Jin, J. Am. Chem. Soc., 2016, 138,
V. D. Thanthirige, G. Ramakrishna, A. Fortunelli and 3950–3953.
A. Dass, J. Phys. Chem. C, 2017, 121, 10865–10869. 70 L. Liao, S. Zhuang, C. Yao, N. Yan, J. Chen, C. Wang,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

47 S. Chen, L. Xiong, S. Wang, Z. Ma, S. Jin, H. Sheng, Y. Pei N. Xia, X. Liu, M.-B. Li, L. Li, X. Bao and Z. Wu, J. Am.
and M. Zhu, J. Am. Chem. Soc., 2016, 138, 10754–10757. Chem. Soc., 2016, 138, 10425–10428.
48 A. Das, T. Li, K. Nobusada, C. Zeng, N. L. Rosi and R. Jin, 71 L. Liao, S. Zhuang, P. Wang, Y. Xu, N. Yan, H. Dong,
J. Am. Chem. Soc., 2013, 135, 18264–18267. C. Wang, Y. Zhao, N. Xia, J. Li, H. Deng, Y. Pei, S.-K. Tian
49 A. Das, T. Li, G. Li, K. Nobusada, C. Zeng, N. L. Rosi and and Z. Wu, Angew. Chem., Int. Ed., 2017, 56, 12644–
R. Jin, Nanoscale, 2014, 6, 6458–6462. 12648.
50 Z. Gan, Y. Lin, L. Luo, G. Han, W. Liu, Z. Liu, C. Yao, 72 S. Zhuang, L. Liao, M.-B. Li, C. Yao, Y. Zhao, H. Dong,
L. Weng, L. Liao, J. Chen, X. Liu, Y. Luo, C. Wang, S. Wei J. Li, H. Deng, L. Li and Z. Wu, Nanoscale, 2017, 9, 14809–
and Z. Wu, Angew. Chem., Int. Ed., 2016, 55, 11567– 14813.
11571. 73 Z. Gan, J. Chen, J. Wang, C. Wang, M.-B. Li, C. Yao,
51 M. W. Heaven, A. Dass, P. S. White, K. M. Holt and S. Zhuang, A. Xu, L. Li and Z. Wu, Nat. Commun., 2017, 8,
R. W. Murray, J. Am. Chem. Soc., 2008, 130, 3754–3755. 14739.
52 M. Zhu, C. M. Aikens, F. J. Hollander, G. C. Schatz and 74 Z. Gan, J. Chen, L. Liao, H. Zhang and Z. Wu, J. Phys.
R. Jin, J. Am. Chem. Soc., 2008, 130, 5883–5885. Chem. Lett., 2018, 9, 204–208.
53 C. Zeng, T. Li, A. Das, N. L. Rosi and R. Jin, J. Am. Chem. 75 C. Zeng, C. Liu, Y. Chen, N. L. Rosi and R. Jin, J. Am.
Soc., 2013, 135, 10011–10013. Chem. Soc., 2016, 138, 8710–8713.
54 Y. Chen, C. Liu, Q. Tang, C. Zeng, T. Higaki, A. Das, 76 L. Liao, J. Chen, C. Wang, S. Zhuang, N. Yan, C. Yao,
D.-e. Jiang, N. L. Rosi and R. Jin, J. Am. Chem. Soc., 2016, N. Xia, L. Li, X. Bao and Z. Wu, Chem. Commun., 2016, 52,
138, 1482–1485. 12036–12036.
55 D. Crasto, S. Malola, G. Brosofsky, A. Dass and 77 T. Higaki, C. Liu, M. Zhou, T.-Y. Luo, N. L. Rosi and R. Jin,
H. Häkkinen, J. Am. Chem. Soc., 2014, 136, 5000–5005. J. Am. Chem. Soc., 2017, 139, 9994–10001.
56 D. Crasto, G. Barcaro, M. Stener, L. Sementa, A. Fortunelli 78 Y. Chen, C. Zeng, C. Liu, K. Kirschbaum, C. Gayathri,
and A. Dass, J. Am. Chem. Soc., 2014, 136, 14933–14940. R. R. Gil, N. L. Rosi and R. Jin, J. Am. Chem. Soc., 2015,
57 T. Higaki, C. Liu, C. Zeng, R. Jin, Y. Chen, N. L. Rosi and 137, 10076–10079.
R. Jin, Angew. Chem., Int. Ed., 2016, 55, 6694–6697. 79 A. Dass, S. Theivendran, P. R. Nimmala, C. Kumara,
58 H. Dong, L. Liao, S. Zhuang, C. Yao, J. Chen, S. Tian, V. R. Jupally, A. Fortunelli, L. Sementa, G. Barcaro, X. Zuo
M. Zhu, X. Liu, L. Li and Z. Wu, Nanoscale, 2017, 9, 3742– and B. C. Noll, J. Am. Chem. Soc., 2015, 137, 4610–4613.
3746. 80 C. Zeng, Y. Chen, K. Kirschbaum, K. Appavoo, M. Y. Sfeir
59 S. Zhuang, L. Liao, Y. Zhao, J. Yuan, C. Yao, X. Liu, J. Li, and R. Jin, Sci. Adv., 2015, 1, e1500045.
H. Deng, J. Yang and Z. Wu, Chem. Sci., 2018, 9, 2437– 81 S. Vergara, D. A. Lukes, M. W. Martynowycz, U. Santiago,
2442. G. Plascencia-Villa, S. C. Weiss, M. J. de la Cruz,
60 A. Das, C. Liu, C. Zeng, G. Li, T. Li, N. L. Rosi and R. Jin, D. M. Black, M. M. Alvarez, X. López-Lozano, C. O. Barnes,
J. Phys. Chem. A, 2014, 118, 8264–8269. G. Lin, H.-C. Weissker, R. L. Whetten, T. Gonen,
61 C. Zeng, H. Qian, T. Li, G. Li, N. L. Rosi, B. Yoon, M. J. Yacaman and G. Calero, J. Phys. Chem. Lett., 2017, 8,
R. N. Barnett, R. L. Whetten, U. Landman and R. Jin, 5523–5530.
Angew. Chem., Int. Ed., 2012, 51, 13114–13118. 82 C. Zeng, Y. Chen, K. Kirschbaum, K. J. Lambright and
62 P. R. Nimmala, S. Knoppe, V. R. Jupally, J. H. Delcamp, R. Jin, Science, 2016, 354, 1580–1584.
C. M. Aikens and A. Dass, J. Phys. Chem. B, 2014, 118, 83 N. A. Sakthivel, S. Theivendran, V. Ganeshraj, A. G. Oliver
14157–14167. and A. Dass, J. Am. Chem. Soc., 2017, 139, 15450–15459.
63 S. Yang, J. Chai, Y. Song, X. Kang, H. Sheng, H. Chong 84 N. Xia, Z. Gan, L. Liao, S. Zhuang and Z. Wu, Chem.
and M. Zhu, J. Am. Chem. Soc., 2015, 137, 10033–10035. Commun., 2017, 53, 11646–11649.
64 H. Qian, W. T. Eckenhoff, Y. Zhu, T. Pintauer and R. Jin, 85 G. Li, C. Zeng and R. Jin, J. Am. Chem. Soc., 2014, 136,
J. Am. Chem. Soc., 2010, 132, 8280–8281. 3673–3669.
65 C. Liu, T. Li, G. Li, K. Nobusada, C. Zeng, G. Pang, 86 H. Qian and R. Jin, Nano Lett., 2009, 9, 4083–4087.
N. L. Rosi and R. Jin, Angew. Chem., Int. Ed., 2015, 54, 87 R. L. Whetten, J. T. Khoury, M. M. Alvarez, S. Murthy,
9826–9829. I. Vezmar, Z. L. Wang, P. W. Stephens, C. L. Cleveland,
66 S. Tian, Y. Li, M. Li, J. Yuan, J. Yang, Z. Wu and R. Jin, W. D. Luedtke and U. Landman, Adv. Mater., 1996, 8, 428–
Nat. Commun., 2015, 6, 8667. 433.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10823
View Article Online

Review Nanoscale

88 H. Qian, Y. Zhu and R. Jin, J. Am. Chem. Soc., 2010, 132, 116 K. R. Krishnadas, A. Baksi, A. Ghosh, G. Natarajan,
4583–4585. A. Som and T. Pradeep, Acc. Chem. Res., 2017, 50, 1988–
89 M. Zhu, H. Qian and R. Jin, J. Phys. Chem. Lett., 2010, 1, 1996.
1003–1007. 117 T. C. W. Mak, X.-L. Zhao, Q.-M. Wang and G.-C. Guo,
90 S. Takano, S. Yamazoe, K. Koyasu and T. Tsukuda, J. Am. Coord. Chem. Rev., 2017, 251, 2311–2333.
Chem. Soc., 2015, 137, 7027–7030. 118 Z. Lei and Q.-M. Wang, Coord. Chem. Rev., 2017, DOI:
91 R. L. Donkers, D. Lee and R. W. Murray, Langmuir, 2004, 10.1016/j.ccr.2017.11.001.
20, 1945–1952. 119 C. D. Bain, E. B. Troughton, Y.-T. Tao, J. Evall,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

92 M. Zhu, E. Lanni, N. Garg, M. E. Bier and R. Jin, J. Am. G. M. Whitesides and R. G. Nuzzo, J. Am. Chem. Soc.,
Chem. Soc., 2008, 130, 1138–1139. 1989, 111, 321–335.
93 J. F. Parker, J. E. F. Weaver, F. McCallum, C. A. Fields- 120 R. Jin, G. Wu, Z. Li, C. A. Mirkin and G. C. Schatz, J. Am.
Zinna and R. W. Murray, Langmuir, 2010, 26, 13650– Chem. Soc., 2003, 125, 1643–1654.
13654. 121 C. A. Mirkin, R. L. Letsinger, R. C. Mucic and
94 J. F. Parker, C. A. Fields-Zinna and R. W. Murray, Acc. J. J. Storhoff, Nature, 1996, 382, 607–609.
Chem. Res., 2010, 43, 1289–1296. 122 M. D. Porter, T. B. Bright, D. L. Allara and
95 R. W. Murray, Chem. Rev., 2008, 108, 2688–2720. C. E. D. Chidsey, J. Am. Chem. Soc., 1987, 109, 3559–3568.
96 Y. Niihori, S. Hossain, S. Sharma, B. Kumar, W. Kurashige 123 A. L. Plant, Langmuir, 1993, 9, 2764–2767.
and Y. Negishi, Chem. Rec., 2017, 17, 473–484. 124 M. Brust, M. Walker, D. Bethell, D. J. Schiffrin and
97 G. Li and R. Jin, J. Am. Chem. Soc., 2014, 136, 11347– R. Whyman, J. Chem. Soc., Chem. Commun., 1994, 801–802.
11354. 125 M. J. Hostetler, S. J. Green, J. J. Stokes and R. W. Murray,
98 B. Du, X. Jiang, A. Das, Q. Zhou, M. Yu, R. Jin and J. Am. Chem. Soc., 1996, 118, 4212–4213.
J. Zheng, Nat. Nanotechnol., 2017, 12, 1096. 126 P. J. G. Goulet and R. B. Lennox, J. Am. Chem. Soc., 2010,
99 N. Goswami, K. Zheng and J. Xie, Nanoscale, 2014, 6, 132, 9582–9584.
13328–13347. 127 Y. Li, O. Zaluzhna, B. Xu, Y. Gao, J. M. Modest and
100 O. A. Wong, R. J. Hansen, T. W. Ni, C. L. Heinecke, Y. Y. J. Tong, J. Am. Chem. Soc., 2011, 133, 2092–
W. S. Compel, D. L. Gustafson and C. J. Ackerson, 2095.
Nanoscale, 2013, 5, 10525–10533. 128 Y. Li, O. Zaluzhna and Y. Y. J. Tong, Langmuir, 2011, 27,
101 S. K. Katla, J. Zhang, E. Castro, R. A. Bernal and X. Li, ACS 7366–7370.
Appl. Mater. Interfaces, 2018, 10, 75–82. 129 S. G. Booth, A. Uehara, S.-Y. Chang, C. L. Fontaine,
102 X.-D. Zhang, J. Chen, Z. Luo, D. Wu, X. Shen, S.-S. Song, T. Fujii, Y. Okamoto, T. Imai, S. L. M. Schroeder and
Y.-M. Sun, P.-X. Liu, J. Zhao, S. Huo, S. Fan, F. Fan, R. A. W. Dryfe, Chem. Sci., 2017, 8, 7954–7962.
X.-J. Liang and J. Xie, Adv. Healthcare Mater., 2014, 3, 133– 130 V. L. Jimenez, D. G. Georganopoulou, R. J. White,
141. A. S. Harper, A. J. Mills, D. Lee and R. W. Murray,
103 Y. Lu and W. Chen, Chem. Soc. Rev., 2012, 41, 3594– Langmuir, 2004, 20, 6864–6870.
3623. 131 J. Kim, K. Lema, M. Ukaigwe and D. Lee, Langmuir, 2007,
104 X. Yuan, X. Dou, K. Zheng and J. Xie, Part. Part. Syst. 23, 7853–7585.
Charact., 2015, 32, 613–629. 132 J. B. Tracy, G. Kalyuzhny, M. C. Crowe,
105 N. Goswami, Q. Yao, T. Chen and J. Xie, Coord. Chem. R. Balasubramanian, J.-P. Choi and R. W. Murray, J. Am.
Rev., 2016, 329, 1–15. Chem. Soc., 2007, 129, 6706–6707.
106 Y.-P. Xie, J.-L. Jin, G.-X. Duan, X. Lu and T. C. W. Mak, 133 S. M. Reilly, T. Krick and A. Dass, J. Phys. Chem. C, 2010,
Coord. Chem. Rev., 2017, 331, 54–72. 114, 741–745.
107 H. Häkkien, Chem. Soc. Rev., 2008, 37, 1847–1859. 134 Y. Shichibu, Y. Negishi, H. Tsunoyama, M. Kanehara,
108 Y. Pei and X. C. Zeng, Nanoscale, 2012, 4, 4054–4072. T. Teranishi and T. Tsukuda, Small, 2007, 3, 835–839.
109 G. Schmid, Chem. Soc. Rev., 2008, 37, 1909–1930. 135 Y. Negishi, Y. Takasugi, S. Sato, H. Yao, K. Kimura and
110 H. Yu, B. Rao, W. Jiang, S. Yang and M. Zhu, Coord. Chem. T. Tsukuda, J. Am. Chem. Soc., 2004, 126, 6518–6519.
Rev., 2017, DOI: 10.1016/j.ccr.2017.12.005. 136 Y. Negishi, K. Nobusada and T. Tsukuda, J. Am. Chem.
111 X. Yuan, Z. Luo, Y. Yu, Q. Yao and J. Xie, Chem. – Asian J., Soc., 2005, 127, 5261–5270.
2013, 8, 858–871. 137 M. Zhu, H. Qian, X. Meng, S. Jin, Z. Wu and R. Jin, Nano
112 N. Goswami, Z. Luo, X. Yuan, D. T. Leong and J. Xie, Lett., 2011, 11, 3963–3969.
Mater. Horiz., 2017, 4, 817–831. 138 T. Cao, S. Jin, S. Wang, D. Zhang, X. Meng and M. Zhu,
113 L.-Y. Chen, C.-W. Wang, Z. Yuan and H.-T. Chang, Anal. Nanoscale, 2013, 5, 7589–7595.
Chem., 2015, 87, 216–229. 139 S. Wang, X. Zhu, T. Cao and M. Zhu, Nanoscale, 2014, 6,
114 X.-R. Song, N. Goswami, H.-H. Yang and J. Xie, Analyst, 5777–5781.
2016, 141, 3126–3140. 140 S. Antonello, G. Arrigoni, T. Dainese, M. D. Nardi,
115 Z. Luo, K. Zheng and J. Xie, Chem. Commun., 2014, 50, G. Parisio, L. Perotti, A. René, A. Venzo and F. Maran, ACS
5143–5155. Nano, 2014, 8, 2788–2795.

10824 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

141 T. Dainese, S. Antonello, J. A. Gascón, F. Pan, N. V. Perera, 163 T. G. Schaaff, M. N. Shafigullin, J. T. Khoury, I. Vezmar
M. Ruzzi, A. Venzo, A. Zoleo, K. Rissanen and F. Maran, and R. L. Whetten, J. Phys. Chem. B, 2001, 105, 8785–8796.
ACS Nano, 2014, 8, 3904–3912. 164 Y. Negishi and T. Tsukuda, J. Am. Chem. Soc., 2003, 125,
142 M. D. Nardi, S. Antonello, D.-e. Jiang, F. Pan, K. Rissanen, 4046–4047.
M. Ruzzi, A. Venzo, A. Zoleo and F. Maran, ACS Nano, 165 Y. Negishi and T. Tsukuda, Chem. Phys. Lett., 2004, 383,
2014, 8, 8505–8512. 161–165.
143 J. Akola, K. A. Kacprzak, O. Lopez-Acevedo, M. Walter, 166 R. Balasubramanian, R. Guo, A. J. Mills and R. W. Murray,
H. Grönbeck and H. Häkkinen, J. Phys. Chem. C, 2010, J. Am. Chem. Soc., 2005, 127, 8126–8132.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

114, 15986–15994. 167 H. Tsunoyama, Y. Negishi and T. Tsukuda, J. Am. Chem.


144 A. Baksi, P. Chakraborty, S. Bhat, G. Natarajan and Soc., 2006, 128, 6036–6037.
T. Pradeep, Chem. Commun., 2016, 52, 8397–8400. 168 H. Tsunoyama, P. Nickut, Y. Negishi, K. Al-Shamery,
145 S. Antonello, T. Dainese, F. Pan, K. Rissanen and Y. Matsumoto and T. Tsukuda, J. Phys. Chem. C, 2007,
F. Maran, J. Am. Chem. Soc., 2017, 139, 4168–4174. 111, 4153–4158.
146 A. Mathew, G. Natarajan, L. Lehtovaara, H. Häkkinen, 169 J. F. Parker, K. A. Kacprzak, O. Lopez-Acevedo,
R. M. Kumar, V. Subramanian, A. Jaleel and T. Pradeep, H. Häkkinen and R. W. Murray, J. Phys. Chem. C, 2010,
ACS Nano, 2014, 8, 139–152. 114, 8276–8281.
147 A. Mathew, E. Varghese, S. Choudhury, S. K. Pal and 170 C. A. Fields-Zinna, J. F. Parker and R. W. Murray, J. Am.
T. Pradeep, Nanoscale, 2015, 7, 14305–14315. Chem. Soc., 2010, 132, 17193–17198.
148 Y. Ishida, K. Narita, T. Yonezawa and R. L. Whetten, 171 A. Dass, K. Holt, J. F. Parker, S. W. Feldberg and
J. Phys. Chem. Lett., 2016, 7, 3718–3722. R. W. Murray, J. Phys. Chem. C, 2008, 112, 20276–
149 C. Lavenn, F. Albrieux, G. Bergeret, R. Chiriac, 20283.
P. Delichère, A. Tuel and A. Demessence, Nanoscale, 2012, 172 S. Si, C. Gautier, J. Boudon, R. Taras, S. Gladiali and
4, 7334–7337. T. Bürgi, J. Phys. Chem. C, 2009, 113, 12966–12969.
150 J. Hassinen, P. Pulkkinen, E. Kalenius, T. Pradeep, 173 S. Knoppew and T. Bürgi, Phys. Chem. Chem. Phys., 2013,
H. Tenhu, H. Häkkinen and R. H. A. Ras, J. Phys. Chem. 15, 15816–15820.
Lett., 2014, 5, 585–589. 174 M. S. Devadas, K. Kwak, J.-W. Park, J.-H. Choi, C.-H. Jun,
151 K. Saha, S. S. Agasti, C. Kim, X. Li and V. M. Rotello, E. Sinn, G. Ramakrishna and D. Lee, J. Phys. Chem. Lett.,
Chem. Rev., 2012, 112, 2739–2779. 2010, 1, 1497–1503.
152 Z. Wu, J. Suhan and R. Jin, J. Mater. Chem., 2009, 19, 622– 175 T. Udayabhaskararao, P. K. Kundu, J. Ahrens and R. Klajn,
626. ChemPhysChem, 2016, 17, 1805–1809.
153 H. Dong, M. Zhu, J. A. Yoon, H. Gao, R. Jin and 176 Z. Tian and A. D. Q. Li, Acc. Chem. Res., 2013, 46, 269–279.
K. Matyjaszewsk, J. Am. Chem. Soc., 2008, 130, 12852– 177 T. W. Ni, M. A. Tofanelli, B. D. Phillips and C. J. Ackerson,
12853. Inorg. Chem., 2014, 53, 6500–6502.
154 G. Li, H. Abroshan, C. Liu, S. Zhuo, Z. Li, Y. Xie, H. J. Kim, 178 A. Fernando and C. M. Aikens, J. Phys. Chem. C, 2015, 119,
N. L. Rosi and R. Jin, ACS Nano, 2016, 10, 7998–8005. 20179–20187.
155 J. Lin, W. Li, C. Liu, P. Huang, M. Zhu, Q. Ge and G. Li, 179 T. M. Carducci, R. E. Blackwell and R. W. Murray, J. Phys.
Nanoscale, 2015, 7, 1363–13670. Chem. Lett., 2015, 6, 1299–1302.
156 Y. Negishi, U. Kamimura, M. Ide and M. Hirayama, 180 G. Salassa, A. Sels, F. Mancin and T. Bürgi, ACS Nano,
Nanoscale, 2012, 4, 4263–4268. 2017, 11, 12609–12614.
157 J. B. Tracy, M. C. Crowe, J. F. Parker, O. Hampe, 181 P. Pengo, C. Bazzo, M. Boccalon and L. Pasquato, Chem.
C. A. Fields-Zinna, A. Dass and R. W. Murray, J. Am. Chem. Commun., 2015, 51, 3204–3207.
Soc., 2007, 129, 16209–16215. 182 C. K. Yee, A. Ulman, J. D. Ruiz, A. Parikh, H. White and
158 A. Dass, A. Stevenson, G. R. Dubay, J. B. Tracy and M. Rafailovich, Langmuir, 2003, 19, 9450–9458.
R. W. Murray, J. Am. Chem. Soc., 2008, 130, 5940–5946. 183 B. S. Zelakiewicz, T. Yonezawa and Y. Y. Tong, J. Am.
159 M. M. Alvarez, J. T. Khoury, T. G. Schaaff, M. Shafigullin, Chem. Soc., 2004, 126, 8112–8113.
I. Vezmar and R. L. Whetten, Chem. Phys. Lett., 1997, 266, 184 M. Brust, N. Stuhr-Hansen, K. Nørgaard,
91–98. J. B. Christensen, L. K. Nielsen and T. Bjørnholm, Nano
160 T. G. Schaaff, M. N. Shafigullin, J. T. Khoury, I. Vezmar, Lett., 2001, 1, 189–191.
R. L. Whetten, W. G. Cullen, P. N. First, C. Gutiérrez- 185 Y. Negishi, W. Kurashige, Y. Niihori and K. Nobusada,
Wing, J. Ascensio and M. J. Jose-Yacamán, J. Phys. Chem. Phys. Chem. Chem. Phys., 2013, 15, 18736–18751.
B, 1997, 101, 7885–7891. 186 J. Zhong, X. Tang, J. Tang, J. Su and Y. Pei, J. Phys. Chem.
161 T. G. Schaaff, G. Knight, M. N. Shafigullin, R. F. Borkman C, 2015, 119, 9205–9214.
and R. L. Whetten, J. Phys. Chem. B, 1998, 102, 10643– 187 Y. Negishi, W. Kurashige and U. Kamimura, Langmuir,
10646. 2011, 27, 12289–12292.
162 T. G. Schaaff and R. L. Whetten, J. Phys. Chem. B, 2000, 188 W. Kurashige, K. Munakata, K. Nobusada and Y. Negishi,
104, 2630–2641. Chem. Commun., 2013, 49, 5447–5449.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10825
View Article Online

Review Nanoscale

189 E. Gottlieb, H. Qian and R. Jin, Chem. – Eur. J., 2013, 19, 213 T. Chen, Q. Yao, X. Yuan, R. R. Nasaruddin and J. Xie,
4238–4243. J. Phys. Chem. C, 2017, 121, 10743–10751.
190 W. Kurashige, M. Yamaguchi, K. Nobusada and 214 J. Xie, Y. Zheng and J. Y. Ying, J. Am. Chem. Soc., 2009,
Y. Negishi, J. Phys. Chem. Lett., 2012, 3, 2649–2652. 131, 888–889.
191 E. Pohjolainen, H. Häkkinen and A. Clayborne, J. Phys. 215 Y. Yu, Z. Luo, C. S. Teo, Y. N. Tan and J. Xie, Chem.
Chem. C, 2015, 119, 9587–9594. Commun., 2013, 49, 9740–9742.
192 X. Meng, Q. Xu, S. Wang and M. Zhu, Nanoscale, 2012, 4, 216 S. Raut, R. Chib, R. Rich, D. Shumilov, Z. Gryczynski and
4161–4165. I. Gryczynski, Nanoscale, 2013, 5, 3441–3446.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

193 W. Kurashige, S. Yamazoe, M. Yamaguchi, K. Nishido, 217 X. Yuan, Y. Yu, Q. Yao, Q. Zhang and J. Xie, J. Phys. Chem.
K. Nobusada, T. Tsukuda and Y. Negishi, J. Phys. Chem. Lett., 2012, 3, 2310–2314.
Lett., 2014, 5, 2072–2076. 218 S. Kumar and R. Jin, Nanoscale, 2012, 4, 4222–4227.
194 Y. Song, J. Zhong, S. Yang, S. Wang, T. Cao, J. Zhang, P. Li, 219 A. Shivhare, L. Wang and R. W. J. Scott, Langmuir, 2015,
D. Hu, Y. Pei and M. Zhu, Nanoscale, 2014, 6, 13977– 31, 1835–1841.
13985. 220 E. S. Shibu, M. A. H. Muhammed, T. Tsukuda and
195 S. Hossain, W. Kurashige, S. Wakayama, B. Kumar, T. Pradeep, J. Phys. Chem. C, 2008, 112, 12168–12176.
L. V. Nair, Y. Niihori and Y. Negishi, J. Phys. Chem. C, 221 Y. Yu, Z. Luo, Y. Yu, J. Y. Lee and J. Xie, ACS Nano, 2012, 6,
2016, 120, 25861–25869. 7920–7927.
196 V. Rojas-Cervellera, C. Rovira and J. Akola, J. Phys. Chem. 222 Z. Luo, V. Nachammai, B. Zhang, N. Yan, D. T. Leong,
Lett., 2015, 6, 3859–3865. D.-e. Jiang and J. Xie, J. Am. Chem. Soc., 2014, 136, 10577–
197 J. Yang, N. Xia, X. Wang, X. Liu, A. Xu, Z. Wu and Z. Luo, 10580.
Nanoscale, 2015, 7, 18464–18470. 223 Q. Yao, X. Yuan, V. Fung, Y. Yu, D. T. Leong, D.-e. Jiang
198 C.-A. J. Lin, T.-Y. Yang, C.-H. Lee, S. H. Huang, and J. Xie, Nat. Commun., 2017, 8, 927.
R. A. Sperling, M. Zanella, J. K. Li, J.-L. Shen, H.-H. Wang, 224 M. J. Hostetler, J. E. Wingate, C.-J. Zhong, J. E. Harris,
H.-I. Yeh, W. J. Parak and W. H. Chang, ACS Nano, 2009, R. W. Vachet, M. R. Clark, J. D. Londono, S. J. Green,
3, 395–401. J. J. Stokes, G. D. Wignall, G. L. Glish, M. D. Porter,
199 F. Wen, Y. Dong, L. Feng, S. Wang, S. Zhang and N. D. Evans and R. W. Murray, Langmuir, 1998, 14, 17–30.
X. Zhang, Anal. Chem., 2011, 83, 1193–1196. 225 Y. Yu, Q. Yao, K. Cheng, X. Yuan, Z. Luo and J. Xie, Part.
200 X. Wu, X. He, K. Wang, C. Xie, B. Zhou and Z. Qing, Part. Syst. Charact., 2014, 31, 652–656.
Nanoscale, 2010, 2, 2244–2249. 226 A. C. Dharmaratne, T. Krick and A. Dass, J. Am. Chem.
201 J. Liu, M. Yu, C. Zhou, S. Yang, X. Ning and J. Zheng, Soc., 2009, 131, 13604–13605.
J. Am. Chem. Soc., 2013, 135, 4978–4981. 227 R. Jin, H. Qian, Z. Wu, Y. Zhu, M. Zhu, A. Mohanty and
202 X.-D. Zhang, Z. Luo, J. Chen, X. Shen, S. Song, Y. Sun, N. Garg, J. Phys. Chem. Lett., 2010, 1, 2903–2910.
S. Fan, F. Fan, D. T. Leong and J. Xie, Adv. Mater., 2014, 228 Z. Wu, M. A. MacDonald, J. Chen, P. Zhang and R. Jin,
26, 4565–4568. J. Am. Chem. Soc., 2011, 133, 9670–9673.
203 Y. Yu, S. Y. New, J. Xie, X. Su and Y. N. Tan, Chem. 229 D. M. Black, C. M. Crittenden, J. S. Brodbelt and
Commun., 2014, 50, 13805–13808. R. L. Whetten, J. Phys. Chem. Lett., 2017, 8, 1283–1289.
204 W. Zhou, Y. Cao, D. Sui, W. Guan, C. Lu and J. Xie, 230 T. A. Dreier, W. S. Compel, O. A. Wong and C. J. Ackerson,
Nanoscale, 2016, 8, 9614–9620. J. Phys. Chem. C, 2016, 120, 28288–28294.
205 X.-D. Zhang, Z. Luo, J. Chen, S. Song, X. Yuan, X. Shen, 231 Y. Negishi, N. K. Chaki, Y. Shichibu, R. L. Whetten and
H. Wang, Y. Sun, K. Gao, L. Zhang, S. Fan, D. T. Leong, T. Tsukuda, J. Am. Chem. Soc., 2007, 129, 11322–11323.
M. Guo and J. Xie, Sci. Rep., 2015, 5, 8669. 232 M. Zhu, W. T. Eckenhoff, T. Pintauer and R. Jin, J. Phys.
206 S. Link, A. Beeby, S. FitzGerald, M. A. El-Sayed, Chem. C, 2008, 112, 14221–14224.
T. G. Schaaff and R. L. Whetten, J. Phys. Chem. B, 2002, 233 M. A. Tofanelli, K. Salorinne, T. W. Ni, S. Malola,
106, 3410–3415. B. Newell, B. Phillips, H. Häkkinen and C. J. Ackerson,
207 S. Link, M. A. El-Sayed, T. G. Schaaff and R. L. Whetten, Chem. Sci., 2016, 7, 1882–1890.
Chem. Phys. Lett., 2002, 356, 240–246. 234 M. Zhu, C. M. Aikens, M. P. Hendrich, R. Gupta, H. Qian,
208 Y. Shichibu, Y. Negishi, T. Tsukuda and T. Teranishi, G. C. Schatz and R. Jin, J. Am. Chem. Soc., 2009, 131,
J. Am. Chem. Soc., 2005, 127, 13464–13465. 2490–2492.
209 L. C. McKenzie, T. O. Zaikova and J. E. Hutchison, J. Am. 235 Z. Wu, C. Gayathri, R. R. Gil and R. Jin, J. Am. Chem. Soc.,
Chem. Soc., 2014, 136, 13426–13435. 2009, 131, 6535–6542.
210 Y. Yu, X. Chen, Q. Yao, Y. Yu, N. Yan and J. Xie, Chem. 236 H. Qian, M. Y. Sfei and R. Jin, J. Phys. Chem. C, 2010, 114,
Mater., 2013, 25, 946–952. 19935–19940.
211 S. Chen, H. Yao and K. Kimura, Langmuir, 2001, 17, 733– 237 J. F. Parker, J.-P. Choi, W. Wang and R. W. Murray, J. Phys.
739. Chem. C, 2008, 112, 13976–13981.
212 X. Yuan, B. Zhang, Z. Luo, Q. Yao, D. T. Leong, N. Yan 238 A. Venzo, S. Antonello, J. A. Gascón, I. Guryanov,
and J. Xie, Angew. Chem., Int. Ed., 2014, 53, 4623–4627. R. D. Leapman, N. V. Perera, A. Sousa, M. Zamuner,

10826 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

A. Zanella and F. Maran, Anal. Chem., 2011, 83, 6355– 266 C. M. Aikens, J. Phys. Chem. Lett., 2011, 2, 99–104.
6362. 267 M. S. Devadas, S. Bairu, H. Qian, E. Sinn, R. Jin and
239 S. Antonello, N. V. Perera, M. Ruzzi, J. A. Gascón and G. Ramakrishna, J. Phys. Chem. Lett., 2011, 2, 2752–2758.
F. Maran, J. Am. Chem. Soc., 2013, 135, 15585–15594. 268 L. Sementa, G. Barcaro, A. Dass, M. Stener and
240 S. Antonello, T. Dainese and F. Maran, Electroanalysis, A. Fortunelli, Chem. Commun., 2015, 51, 7935–7938.
2016, 28, 2771–2776. 269 X. Yuan, N. Goswami, W. Chen, Q. Yao and J. Xie, Chem.
241 Z. Liu, M. Zhu, X. Meng, G. Xu and R. Jin, J. Phys. Chem. Commun., 2016, 52, 5234–5237.
Lett., 2011, 2, 2104–2109. 270 E. M. S. Stennett, M. A. Ciuba and M. Levitus, Chem. Soc.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

242 H. Chong, P. Li, S. Wang, F. Fu, J. Xiang, M. Zhu and Y. Li, Rev., 2014, 43, 1057–1075.
Sci. Rep., 2013, 3, 3214. 271 Q. Zheng, M. F. Juette, S. Jockusch, M. R. Wasserman,
243 H. A. Jahn and E. Teller, Proc. R. Soc. London, Ser. A, 1937, Z. Zhou, R. B. Altman and S. C. Blanchard, Chem. Soc.
161, 220–235. Rev., 2014, 43, 1044–1056.
244 K. Kwak, Q. Tang, M. Kim, D.-e. Jiang and D. Lee, J. Am. 272 M. Montalti, L. Prodi, E. Rampazzo and N. Zaccheroni,
Chem. Soc., 2015, 137, 10833–10840. Chem. Soc. Rev., 2014, 43, 4243–4268.
245 C. B. Collins, M. A. Tofanelli, M. F. Crook, B. D. Phillips 273 A. S. Klymchenko, Acc. Chem. Res., 2017, 50, 366–375.
and C. J. Ackerson, RSC Adv., 2017, 7, 45061–45065. 274 Z. Wu and R. Jin, Nano Lett., 2010, 10, 2568–2573.
246 K. Ikeda, Y. Kobayashi, Y. Negishi, M. Seto, T. Iwasa, 275 F. Wang and X. Liu, Chem. Soc. Rev., 2009, 38,
K. Nobusada, T. Tsukuda and N. Kojima, J. Am. Chem. 976–989.
Soc., 2007, 129, 7230–7231. 276 D. Wu, A. C. Sedgwick, T. Gunnlaugsson, E. U. Akkaya,
247 J. Akola, M. Walter, R. L. Whetten, H. Häkkinen and J. Yoon and T. D. James, Chem. Soc. Rev., 2017, 46, 7105–
H. Grönbeck, J. Am. Chem. Soc., 2008, 130, 3756–3757. 7123.
248 C. M. Aikens, J. Phys. Chem. Lett., 2010, 1, 2594–2599. 277 G. H. Carey, A. L. Abdelhady, Z. Ning, S. M. Thon,
249 V. R. Jupally, R. Kota, E. V. Dornshuld, D. L. Mattern, O. M. Bakr and E. H. Sargent, Chem. Rev., 2015, 115,
G. S. Tschumper, D.-e. Jiang and A. Dass, J. Am. Chem. 12732–12763.
Soc., 2011, 133, 20258–20266. 278 P. Ning, W. Wang, M. Chen, Y. Feng and X. Meng, Chin.
250 D.-e. Jiang and M. Walter, Nanoscale, 2012, 4, 4234–4239. Chem. Lett., 2017, 28, 1943–1951.
251 A. Tlahuice-Flores, R. L. Whetten and M. Jose-Yacaman, 279 Z. Tang, B. Xu, B. Wu, D. A. Robinson, N. Bokossa and
J. Phys. Chem. C, 2013, 117, 20867–20875. G. Wang, Langmuir, 2011, 27, 2989–2996.
252 C. Liu, S. Lin, Y. Pei and X. C. Zeng, J. Am. Chem. Soc., 280 Z. Tang, T. Ahuja, S. Wang and G. Wang, Nanoscale, 2012,
2013, 135, 18067–18079. 4, 4119–4124.
253 A. I. Frenkel, Chem. Soc. Rev., 2012, 41, 8163–8178. 281 C. V. Conroy, J. Jiang, C. Zhang, T. Ahuja, Z. Tang,
254 G. Meitzner, G. H. Via, F. W. Lytle and J. H. Sinfelt, J. Phys. C. A. Prickett, J. J. Yang and G. Wang, Nanoscale, 2014, 6,
Chem., 1992, 96, 4960–4964. 7416–7423.
255 A. I. Frenkel, C. W. Hills and R. G. Nuzzo, J. Phys. Chem. B, 282 X. Wen, P. Yu, Y.-R. Toh and J. Tang, J. Phys. Chem. C,
2001, 105, 12689–12703. 2013, 117, 3621–3626.
256 S. Bordiga, E. Groppo, G. Agostini, J. A. v. Bokhoven and 283 T. D. Green, C. Yi, C. Zeng, R. Jin, S. McGill and
C. Lamberti, Chem. Rev., 2013, 113, 1736–1850. K. L. Knappenberger, J. Phys. Chem. A, 2014, 118, 10611–
257 P. Zhang, J. Phys. Chem. C, 2014, 118, 25291–25299. 10621.
258 M. A. MacDonald, D. M. Chevrier and P. Zhang, J. Phys. 284 K. L. D. M. Weerawardene and C. M. Aikens, J. Am. Chem.
Chem. C, 2011, 115, 15282–15287. Soc., 2016, 138, 11202–11210.
259 D. M. Chevrier, X. Meng, Q. Tang, D.-e. Jiang, M. Zhu, 285 T. D. Green, P. J. Herbert, C. Yi, C. Zeng, S. McGill, R. Jin
A. Chatt and P. Zhang, J. Phys. Chem. C, 2014, 118, 21730– and K. L. Knappenberger, J. Phys. Chem. C, 2016, 120,
21737. 17784–17790.
260 S. Yamazoe, S. Takano, W. Kurashige, T. Yokoyama, 286 X. Wen, P. Yu, Y.-R. Toh, A.-C. Hsu, Y.-C. Lee and J. Tang,
K. Nitta, Y. Negishi and T. Tsukuda, Nat. Commun., 2016, J. Phys. Chem. C, 2012, 116, 19032–19038.
7, 10414. 287 W. Miao, Chem. Rev., 2008, 108, 2506–2553.
261 M. Faraday, Philos. Trans. R. Soc. London, 1857, 147, 145– 288 M. Sentic, M. Milutinovic, F. Kanoufi, D. Manojlovic,
181. S. Arbault and N. Sojic, Chem. Sci., 2014, 5, 2568–2572.
262 M. M. Alvarez, J. T. Khoury, T. G. Schaaff, 289 K. N. Swanick, M. Hesari, M. S. Workentin and Z. Ding,
M. N. Shafigullin, I. Vezmar and R. L. Whetten, J. Phys. J. Am. Chem. Soc., 2012, 134, 15205–15208.
Chem. B, 1997, 101, 3706–3712. 290 M. Hesari, M. S. Workentin and Z. Ding, Chem. Sci., 2014,
263 C. M. Aikens, J. Phys. Chem. A, 2009, 113, 10811–10817. 5, 3814–3822.
264 D.-e. Jiang, M. Kühn, Q. Tang and F. Weigend, J. Phys. 291 M. Hesari, M. S. Workentin and Z. Ding, RSC Adv., 2014,
Chem. Lett., 2014, 5, 3286–3289. 4, 29559–29562.
265 R. Guo and R. W. Murray, J. Am. Chem. Soc., 2005, 127, 292 M. Hesari, M. S. Workentin and Z. Ding, Chem. – Eur. J.,
12140–12413. 2014, 20, 15116–15121.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10827
View Article Online

Review Nanoscale

293 Y.-M. Fang, J. Song, J. Li, Y.-W. Wang, H.-H. Yang, J.-J. Sun 317 S. A. Miller, C. A. Fields-Zinna, R. W. Murray and
and G.-N. Chen, Chem. Commun., 2011, 47, 2369–2371. A. M. Moran, J. Phys. Chem. Lett., 2010, 1, 1383–1387.
294 L. Li, H. Liu, Y. Shen, J. Zhang and J.-J. Zhu, Anal. Chem., 318 M. S. Devadas, J. Kim, E. Sinn, D. Lee, T. Goodson and
2011, 83, 661–665. G. Ramakrishna, J. Phys. Chem. C, 2010, 114, 22417–
295 R. Philip, P. Chantharasupawong, H. Qian, R. Jin and 22423.
J. Thomas, Nano Lett., 2012, 12, 4661–4667. 319 S. H. Yau, O. Varnavski, J. D. Gilbertson, B. Chandler,
296 I. Russier-Antoine, F. Bertorelle, M. Vojkovic, D. Rayane, G. Ramakrishna and T. Goodson, J. Phys. Chem. C, 2010,
E. Salmon, C. Jonin, P. Dugourd, R. Antoine and 114, 15979–15985.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

P.-F. Brevet, Nanoscale, 2014, 6, 13572–13578. 320 T. D. Green and K. L. Knappenberger, Nanoscale, 2012, 4,
297 S. Knoppe, H. Häkkinen and T. Verbiest, J. Phys. Chem. C, 4111–4118.
2015, 119, 27676–27682. 321 K. G. Stamplecoskie and P. V. Kamat, J. Am. Chem. Soc.,
298 P. N. Day, K. A. Nguyen and R. Pachter, J. Chem. Theory 2014, 136, 11093–11099.
Comput., 2010, 6, 2809–2821. 322 R. D. Senanayake, A. V. Akimov and C. M. Aikens, J. Phys.
299 L. Polavarapu, M. Manna and Q.-H. Xu, Nanoscale, 2011, Chem. C, 2017, 121, 10653–10662.
3, 429–434. 323 T. Stoll, E. Sgrò, J. W. Jarrett, J. Réhault, A. Oriana, L. Sala,
300 G. Ramakrishna, O. Varnavski, J. Kim, D. Lee and F. Branchi, G. Cerullo and K. L. Knappenberger, J. Am.
T. Goodson, J. Am. Chem. Soc., 2008, 130, 5032–5033. Chem. Soc., 2016, 138, 1788–1791.
301 R. Ho-Wu, S. H. Yau and T. Goodson, ACS Nano, 2016, 10, 324 C. Yi, H. Zheng, P. J. Herbert, Y. Chen, R. Jin and
562–572. K. L. Knappenberger, J. Phys. Chem. C, 2017, 121, 24894–
302 N. Abeyasinghe, S. Kumar, K. Sun, J. F. Mansfield, R. Jin 24902.
and T. Goodson, J. Am. Chem. Soc., 2016, 138, 16299– 325 M. Farrag, M. Tschurl, A. Dass and U. Heiz, Phys. Chem.
16307. Chem. Phys., 2013, 15, 12539–12542.
303 S. Knoppe, M. Vanbel, S. v. Cleuvenbergen, L. Vanpraet, 326 B. Varnholt, P. Oulevey, S. Luber, C. Kumara, A. Dass and
T. Bürgi and T. Verbiest, J. Phys. Chem. C, 2015, 119, 6221– T. Bürgi, J. Phys. Chem. C, 2014, 118, 9604–9611.
6226. 327 B. Varnholt, M. J. Guberman-Pfeffer, P. Oulevey,
304 S. L. Raut, D. Shumilov, R. Chib, R. Rich, Z. Gryczynski S. Antonello, T. Dainese, J. A. Gascón, T. Bürgi and
and I. Gryczynski, Chem. Phys. Lett., 2013, 561, 74–76. F. Maran, J. Phys. Chem. C, 2016, 120, 25378–25386.
305 Z. Hu and L. Jensen, Chem. Sci., 2017, 8, 4595–4601. 328 M. Walter, J. Akola, O. Lopez-Acevedo, P. D. Jadzinsky,
306 G. V. Hartland, Chem. Rev., 2011, 111, 3858–3887. G. Calero, C. J. Ackerson, R. L. Whetten, H. Grönbeck and
307 M. Y. Sfeir, H. Qian, K. Nobusada and R. Jin, J. Phys. H. Häkkinen, Proc. Natl. Acad. Sci. U. S. A., 2008, 105,
Chem. C, 2011, 115, 6200–6207. 9157–9162.
308 M. S. Devadas, V. D. Thanthirige, S. Bairu, E. Sinn and 329 K. Hirata, K. Yamashita, S. Muramatsu, S. Takano,
G. Ramakrishna, J. Phys. Chem. C, 2013, 117, 23155– K. Ohshimo, T. Azuma, R. Nakanishi, S. Yamazoe,
23161. K. Koyasu and T. Tsukuda, Nanoscale, 2017, 9, 13409–
309 M. Zhou, S. Vdović, S. Long, M. Zhu, L. Yan, Y. Wang, 13412.
Y. Niu, X. Wang, Q. Guo, R. Jin and A. Xia, J. Phys. Chem. 330 T. Ohta, M. Shibuta, H. Tsunoyama, Y. Negishi, T. Eguchi
A, 2013, 117, 10294–10303. and A. Nakajima, J. Phys. Chem. C, 2013, 117, 3674–
310 S. Mustalahti, P. Myllyperkiö, T. Lahtinen, K. Salorinne, 3679.
S. Malola, J. Koivisto, H. Häkkinen and M. Pettersson, 331 L. Shi, L. Zhu, J. Guo, L. Zhang, Y. Shi, Y. Zhang, K. Hou,
J. Phys. Chem. C, 2014, 118, 18233–18239. Y. Zheng, Y. Zhu, J. Lv, S. Liu and Z. Tang, Angew. Chem.,
311 M. Zhou, S. Long, X. Wan, Y. Li, Y. Niu, Q. Guo, Int. Ed., 2017, 56, 15397–15401.
Q.-M. Wang and A. Xia, Phys. Chem. Chem. Phys., 2014, 16, 332 H. Yang, J. Yan, Y. Wang, G. Deng, H. Su, X. Zhao, C. Xu,
18288–18293. B. K. Teo and N. Zheng, J. Am. Chem. Soc., 2017, 139,
312 C. Yi, H. Zheng, L. M. Tvedte, C. J. Ackerson and 16113–16116.
K. L. Knappenberger, J. Phys. Chem. C, 2015, 119, 6307– 333 J. Chen, L. Liu, X. Liu, L. Liao, S. Zhuang, S. Zhou, J. Yang
6313. and Z. Wu, Chem. – Eur. J., 2017, 23, 18187–18192.
313 M. Zhou, J. Zhong, S. Wang, Q. Guo, M. Zhu, Y. Pei and 334 S. Takano and T. Tsukuda, J. Phys. Chem. Lett., 2016, 7,
A. Xia, J. Phys. Chem. C, 2015, 119, 18790–18797. 4509–4513.
314 C. Yi, M. A. Tofanelli, C. J. Ackerson and 335 X. He, Y. Wang, H. Jiang and L. Zhao, J. Am. Chem. Soc.,
K. L. Knappenberger, J. Am. Chem. Soc., 2013, 135, 18222– 2016, 138, 5634–5643.
18228. 336 I. Dolamic, B. Varnholt and T. Bürgi, Nat. Commun., 2015,
315 O. Varnavski, G. Ramakrishna, J. Kim, D. Lee and 6, 7117.
T. Goodson, J. Am. Chem. Soc., 2010, 132, 16–17. 337 X.-K. Wan, S.-F. Yuan, Z.-W. Lin and Q.-M. Wang, Angew.
316 S. Mustalahti, P. Myllyperkiö, S. Malola, T. Lahtinen, Chem., Int. Ed., 2014, 53, 2923–2926.
K. Salorinne, J. Koivisto, H. Häkkinen and M. Pettersson, 338 I. Dolamic, S. Knoppe, A. Dass and T. Bürgi, Nat.
ACS Nano, 2015, 9, 2328–2335. Commun., 2012, 3, 798.

10828 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

339 J. Yan, H. Su, H. Yang, C. Hu, S. Malola, S. Lin, B. K. Teo, 367 I. Carmeli, G. Leitus, R. Naaman, S. Reich and Z. Vager,
H. Häkkinen and N. Zheng, J. Am. Chem. Soc., 2016, 138, J. Chem. Phys., 2003, 118, 10372–10375.
12751–12754. 368 V. Tuboltsev, A. Savin, A. Pirojenko and J. Räisänen, ACS
340 O. Lopez-Acevedo, H. Tsunoyama, T. Tsukuda, Nano, 2013, 7, 6691–6699.
H. Häkkinen and C. M. Aikens, J. Am. Chem. Soc., 2010, 369 G. L. Nealon, B. Donnio, R. Greget, J.-P. Kappler,
132, 8210–8218. E. Terazzi and J.-L. Gallani, Nanoscale, 2012, 4, 5244–5285.
341 R. Kazan, B. Zhang and T. Bürgi, Dalton Trans., 2017, 46, 370 W. Qin, J. Lohrman and S. Ren, Angew. Chem., Int. Ed.,
7708–7713. 2014, 53, 7316–7319.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

342 X. Zou, S. Jin, W. Du, Y. Li, P. Li, S. Wang and M. Zhu, 371 P. Crespo, R. Litárn, T. C. Rojas, M. Multigner, J. M. de la
Nanoscale, 2017, 9, 16800–16805. Fuente, J. C. Sánchez-López, M. A. García, A. Hernando,
343 Q. Xu, S. Kumar, S. Jin, H. Qian, M. Zhu and R. Jin, Small, S. Penadés and A. Fernández, Phys. Rev. Lett., 2004, 93,
2014, 10, 1008–1014. 087204.
344 H. Qian, M. Zhu, C. Gayathri, R. R. Gil and R. Jin, ACS 372 R. S. McCoy, S. Choi, G. Collins, B. J. Ackerson and
Nano, 2011, 5, 8935–8942. C. J. Ackerson, ACS Nano, 2013, 7, 2610–2616.
345 A. Sánchez-Castillo, C. Noguez and I. L. Garzón, J. Am. 373 Y. Negishi, H. Tsunoyama, M. Suzuki, N. Kawamura,
Chem. Soc., 2010, 132, 1504–1505. M. M. Matsushita, K. Maruyama, T. Sugawara,
346 H. Yao, J. Phys. Chem. Lett., 2012, 3, 1701–1706. T. Yokoyama and T. Tsukuda, J. Am. Chem. Soc., 2006, 128,
347 R. Kobayashi, Y. Nonoguchi, A. Sasaki and H. Yao, J. Phys. 12034–12035.
Chem. C, 2014, 118, 15506–15515. 374 Z. Wu, J. Chen and R. Jin, Adv. Funct. Mater., 2011, 21,
348 H. Yao, T. Fukui and K. Kimura, J. Phys. Chem. C, 2008, 177–183.
112, 16281–16285. 375 K. S. Krishna, P. Tarakeshwar, V. Mujica and C. S. S.
349 H. Yao and S. Yaomura, Langmuir, 2013, 29, 6444–6451. R. Kumar, Small, 2014, 10, 907–911.
350 T. Laaksonen, V. Ruiz, P. Liljeroth and B. M. Quinn, 376 M. Agrachev, S. Antonello, T. Dainese, J. A. Gascón,
Chem. Soc. Rev., 2008, 37, 1836–1846. F. Pan, K. Rissanen, M. Ruzzi, A. Venzo, A. Zoleo and
351 D. Wang, J. W. Padelford, T. Ahuja and G. Wang, ACS F. Maran, Chem. Sci., 2016, 7, 6910–6918.
Nano, 2015, 9, 8344–8351. 377 M. Agrachev, S. Antonello, T. Dainese, M. Ruzzi, A. Zoleo,
352 B. M. Quinn, P. Liljeroth, V. Ruiz, T. Laaksonen and E. Aprà, N. Govind, A. Fortunelli, L. Sementa and
K. Kontturi, J. Am. Chem. Soc., 2003, 125, 6644–6645. F. Maran, ACS Omega, 2017, 2, 2607–2617.
353 Z. Tang, D. A. Robinson, N. Bokossa, B. Xu, S. Wang and 378 J. U. Reveles, S. N. Khanna, P. J. Roach and
G. Wang, J. Am. Chem. Soc., 2011, 133, 16037–16044. A. W. Castleman, Proc. Natl. Acad. Sci. U. S. A., 2006, 103,
354 T. M. Carducci and R. W. Murray, J. Am. Chem. Soc., 2013, 18405–18410.
135, 11351–11356. 379 A. W. Castleman and S. N. Khanna, J. Phys. Chem. C, 2009,
355 T. M. Carducci, R. E. Blackwell and R. W. Murray, J. Am. 113, 2664–2675.
Chem. Soc., 2014, 136, 11182–11187. 380 Z. Wu and R. Jin, ACS Nano, 2009, 3, 2036–2042.
356 X. Kang, S. Chen, S. Jin, Y. Song, Y. Xu, H. Yu, H. Sheng 381 M. A. Tofanelli and C. J. Ackerson, J. Am. Chem. Soc., 2012,
and M. Zhu, ChemElectroChem, 2016, 3, 1261–1265. 134, 16937–16940.
357 D. Lee, R. L. Donkers, J. M. DeSimone and R. W. Murray, 382 R. Ouyang and D.-e. Jiang, J. Phys. Chem. C, 2015, 119,
J. Am. Chem. Soc., 2003, 125, 1182–1183. 21555–21560.
358 D. Lee, R. L. Donkers, G. Wang, A. S. Harper and 383 J. Jung, S. Kang and Y.-K. Han, Nanoscale, 2012, 4, 4206–
R. W. Murray, J. Am. Chem. Soc., 2004, 126, 6193–6199. 4210.
359 W. Wang, D. Lee and R. W. Murray, J. Phys. Chem. B, 2006, 384 Y. Zhang, M. Yan, S. Wang, J. Jiang, P. Gao, G. Zhang,
110, 10258–10265. C. Dong and S. Shuang, RSC Adv., 2016, 6, 8612–8619.
360 J. Kim and D. Lee, J. Am. Chem. Soc., 2006, 128, 4518– 385 S. Roy, G. Palui and A. Banerjee, Nanoscale, 2012, 4, 2734–
4819. 2740.
361 D. García-Raya, R. Madueño, M. Blázquez and T. Pineda, 386 A. Mathew, P. R. Sajanlal and T. Pradeep, Angew. Chem.,
J. Phys. Chem. C, 2009, 113, 8756–8761. Int. Ed., 2012, 51, 9596–9600.
362 S. S. Kumar, K. Kwak and D. Lee, Anal. Chem., 2011, 83, 387 B. Zheng, J. Zheng, T. Yu, A. Sang, J. Du, Y. Guo, D. Xiao
3244–3247. and M. M. F. Choi, Sens. Actuators, B, 2015, 221, 386–392.
363 S. S. Kumar, K. Kwak and D. Lee, Electroanalysis, 2011, 23, 388 P.-C. Chen, P. Roy, L.-Y. Chen, R. Ravindranath and
2116–2124. H.-T. Chang, Part. Part. Syst. Charact., 2014, 31, 917–942.
364 K. Kwak and D. Lee, J. Phys. Chem. Lett., 2012, 3, 2476– 389 G. Guan, S.-Y. Zhang, Y. Cai, S. Liu, M. S. Bharathi,
2481. M. Low, Y. Yu, J. Xie, Y. Zheng, Y.-W. Zhang and
365 S. Antonello, A. H. Holm, E. Instuli and F. Maran, J. Am. M.-Y. Han, Chem. Commun., 2014, 50, 5703–5705.
Chem. Soc., 2007, 129, 9836–9837. 390 M. Shellaiah and K. W. Sun, Chemosensors, 2017, 5, 36.
366 H. Hori, T. Teranishi, Y. Nakae, Y. Seino, M. Miyake and 391 X. Zhang, F.-G. Wu, P. Liu, H.-Y. Wang, N. Gu and
S. Yamada, Phys. Lett. A, 1999, 263, 406–410. Z. Chen, J. Colloid Interface Sci., 2015, 455, 6–15.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10829
View Article Online

Review Nanoscale

392 K. Zheng, X. Yuan, N. Goswami, Q. Zhang and J. Xie, RSC 417 H. Chen, C. Liu, M. Wang, C. Zhang, N. Luo, Y. Wang,
Adv., 2014, 4, 60581–60596. H. Abroshan, G. Li and F. Wang, ACS Catal., 2017, 7,
393 T. Zhao, T. Zhou, Q. Yao, C. Hao and X. Chen, J. Environ. 3632–3638.
Sci. Health, Part C: Environ. Carcinog. Ecotoxicol. Rev., 418 Y. Wang, H. Su, C. Xu, G. Li, L. Gell, S. Lin, Z. Tang,
2015, 33, 168–187. H. Häkkinen and N. Zheng, J. Am. Chem. Soc., 2015, 137,
394 M. C. Alonso, L. Trapiella-Alfonso, J. M. C. Fernández, 4324–4327.
R. Pereiro and A. Sanz-Medel, Biosens. Bioelectron., 2016, 419 Y. Wang, X.-K. Wan, L. Ren, H. Su, G. Li, S. Malola, S. Lin,
77, 1055–1061. Z. Tang, H. Häkkinen, B. K. Teo, Q.-M. Wang and
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

395 A. Samanta, B. B. Dhar and R. N. Devi, J. Phys. Chem. C, N. Zheng, J. Am. Chem. Soc., 2016, 138, 3278–3281.
2012, 116, 1748–1754. 420 B. Panthi, A. Mukhopadhyay, L. Tibbitts, J. Saavedra,
396 S. Xu, H. Yang, K. Zhao, J. Li, L. Mei, Y. Xie and A. Deng, C. J. Pursell, R. M. Rioux and B. D. Chandler, ACS Catal.,
RSC Adv., 2015, 5, 11343–11348. 2015, 5, 2232–2241.
397 S. Palmal, S. Basiruddin, A. R. Maity, S. C. Ray and 421 T. Yoskamtorn, S. Yamazoe, R. Takahata, J.-i. Nishigaki,
N. R. Jana, Chem. – Eur. J., 2013, 19, 943–949. A. Thivasasith, J. Limtrakul and T. Tsukuda, ACS Catal.,
398 Z. Li, H. Peng, J. Liu, Y. Tian, W. Yang, J. Yao, Z. Shao and 2014, 4, 3696–3700.
X. Chen, ACS Appl. Mater. Interfaces, 2018, 10, 83–90. 422 A. Shivhare, K. E. Lee, Y. Hu and R. W. J. Scott, J. Phys.
399 J. Zheng, C. Zhou, M. Yu and J. Liu, Nanoscale, 2012, 4, Chem. C, 2015, 119, 23279–23284.
4073–4083. 423 J. Fang, J. Li, B. Zhang, X. Yuan, H. Asakura, T. Tanaka,
400 Y. Wang, Y. Wang, F. Zhou, P. Kim and Y. Xia, Small, 2012, K. Teramura, J. Xie and N. Yan, Nanoscale, 2015, 7, 6325–
8, 3769–3773. 6333.
401 T. D. Fernández, J. R. Pearson, M. P. Leal, M. J. Torres, 424 S. Das, A. Goswami, M. Hesari, J. F. Al-Sharab,
M. Blanca, C. Mayorga and X. L. Guével, Biomaterials, E. Mikmeková, F. Maran and T. Asefa, Small, 2014, 10,
2015, 43, 1–12. 1473–1478.
402 X.-D. Zhang, J. Yang, S.-S. Song, W. Long, J. Chen, 425 G. Ma, A. Binder, M. Chi, C. Liu, R. Jin, D.-e. Jiang, J. Fan
X. Shen, H. Wang, Y.-M. Sun, P.-X. Liu and S. Fan, and S. Dai, Chem. Commun., 2012, 48, 11413–11415.
Int. J. Nanomed., 2014, 9, 2069–2072. 426 B. Zhang, J. Fang, J. Li, J. J. Lau, D. Mattia, Z. Zhong,
403 X. Jia, J. Li, L. Han, J. Ren, X. Yang and E. Wang, ACS J. Xie and N. Yan, Chem. – Asian J., 2016, 11, 532–539.
Nano, 2012, 6, 3311–3317. 427 W. Li, Q. Ge, X. Ma, Y. Chen, M. Zhu, H. Xu and R. Jin,
404 W. Song, Y. Wang, R.-P. Liang, L. Zhang and J.-D. Qiu, Nanoscale, 2016, 8, 2378–2385.
Biosens. Bioelectron., 2015, 64, 234–240. 428 I. Cano, M. A. Huertos, A. M. Chapman, G. Buntkowsky,
405 K. M. Harkness, B. N. Turner, A. C. Agrawal, Y. Zhang, T. Gutmann, P. B. Groszewicz and P. W. N. M. van
J. A. McLean and D. E. Cliffe, Nanoscale, 2012, 4, 3843–3851. Leeuwen, J. Am. Chem. Soc., 2015, 137, 7718–7727.
406 K. Luo, F. Nie, Y. Yan, S. Wang, X. Zheng and Z. Song, RSC 429 T. Mitsudome, M. Yamamoto, Z. Maeno, T. Mizugaki,
Adv., 2014, 4, 61465–61475. K. Jitsukawa and K. Kaneda, J. Am. Chem. Soc., 2015, 137,
407 H. Kawasaki, S. Kumar, G. Li, C. Zeng, D. R. Kauffman, 13452–13455.
J. Yoshimoto, Y. Iwasaki and R. Jin, Chem. Mater., 2014, 430 X. Yang, M. Yang, B. Pang, M. Vara and Y. Xia, Chem. Rev.,
26, 2777–2788. 2015, 115, 10410–10488.
408 J. Zhang, Z. Li, J. Huang, C. Liu, F. Hong, K. Zheng and 431 K. M. Tsoi, S. A. MacParland, X.-Z. Ma, V. N. Spetzler,
G. Li, Nanoscale, 2017, 9, 16879–16886. J. Echeverri, B. Ouyang, S. M. Fadel, E. A. Sykes,
409 C. Liu, C. Yan, J. Lin, C. Yu, J. Huang and G. Li, J. Mater. N. Goldaracena, J. M. Kaths, J. B. Conneely, B. A. Alman,
Chem. A, 2015, 3, 20167–20173. M. Selzner, M. A. Ostrowski, O. A. Adeyi, A. Zilman,
410 G. Li, C. Liu, Y. Lei and R. Jin, Chem. Commun., 2012, 48, I. D. McGilvray and W. C. W. Chan, Nat. Mater., 2016, 15,
12005–12007. 1212–1221.
411 M.-B. Li, S.-K. Tian and Z. Wu, Nanoscale, 2014, 6, 5714– 432 T. Huang and R. W. Murray, J. Phys. Chem. B, 2001, 105,
5717. 12498–12502.
412 K.-M. Choi, T. Akita, T. Mizugaki, K. Ebitani and 433 X. L. Guével, B. Hötzer, G. Jung, K. Hollemeyer,
K. Kaneda, New J. Chem., 2003, 27, 324–328. V. Trouillet and M. Schneider, J. Phys. Chem. C, 2011, 115,
413 Y. Gao, N. Shao, Y. Pei and X. C. Zeng, Nano Lett., 2010, 10955–10963.
10, 1055–1062. 434 G. Aragay, J. Pon and A. Merkoc, Chem. Rev., 2011, 111,
414 Y. Wu, D. Wang and Y. Li, Chem. Soc. Rev., 2014, 43, 2112– 3433–3458.
2124. 435 J. Xie, Y. Zheng and J. Y. Ying, Chem. Commun., 2010, 46,
415 A. W. Cook, Z. R. Jones, G. Wu, S. L. Scott and 961–963.
T. W. Hayton, J. Am. Chem. Soc., 2018, 140, 394–400. 436 K.-Y. Pu, Z. Luo, K. Li, J. Xie and B. Liu, J. Phys. Chem. C,
416 S. Zhao, R. Jin, H. Abroshan, C. Zeng, H. Zhang, 2011, 115, 13069–13075.
S. D. House, E. Gottlieb, H. J. Kim, J. C. Yang and R. Jin, 437 B. Unnikrishnan, S.-C. Wei, W.-J. Chiu, J. Cang, P.-H. Hsu
J. Am. Chem. Soc., 2017, 139, 1077–1080. and C.-C. Huang, Analyst, 2014, 139, 2221–2228.

10830 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

438 P. Yu, X. Wen, Y.-R. Toh, J. Huang and J. Tang, Part. Part. 465 Z. Wu, D.-e. Jiang, A. K. P. Mann, D. R. Mullins,
Syst. Charact., 2013, 30, 467–472. Z.-A. Qiao, L. F. Allard, C. Zeng, R. Jin and S. H. Overbury,
439 M. Wang, Z. Wu, J. Yang, G. Wang, H. Wang and W. Cai, J. Am. Chem. Soc., 2014, 136, 6111–6122.
Nanoscale, 2012, 4, 4087–4090. 466 C. Lavenn, A. Demessence and A. Tuel, J. Catal., 2015,
440 Z. Wu, M. Wang, J. Yang, X. Zheng, W. Cai, G. Meng, 322, 130–138.
H. Qian, H. Wang and R. Jin, Small, 2012, 8, 2028–2035. 467 P. Huang, G. Chen, Z. Jiang, R. Jin, Y. Zhu and Y. Sun,
441 J. Wang, Y. Chang, P. Zhang, S. Q. Lie, P. F. Gao and Nanoscale, 2013, 5, 3668–3672.
C. Z. Huang, Analyst, 2015, 140, 8194–8200. 468 J. Liu, K. S. Krishna, Y. B. Losovyj, S. Chattopadhyay,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

442 T. Shu, J. Wang, L. Su and X. Zhang, Anal. Chem., 2016, N. Lozova, J. T. Miller, J. J. Spivey and C. S. S. R. Kumar,
88, 11193–11198. Chem. – Eur. J., 2013, 19, 10201–10208.
443 T. Shu, L. Su, J. Wang, C. Li and X. Zhang, Biosens. 469 T. A. Dreier, O. A. Wong and C. J. Ackerson, Chem.
Bioelectron., 2015, 66, 155–161. Commun., 2015, 51, 1240–1243.
444 S. Wang, X. Meng, A. Das, T. Li, Y. Song, T. Cao, X. Zhu, 470 C. Lavenn, A. Demessence and A. Tuel, Catal. Today, 2014,
M. Zhu and R. Jin, Angew. Chem., Int. Ed., 2014, 53, 2376– 235, 72–78.
2380. 471 S. Zhao, A. Das, H. Zhang, R. Jin, Y. Song and R. Jin, Prog.
445 M. Yu, J. Zhou, B. Du, X. Ning, C. Authement, L. Gandee, Nat. Sci.: Mater. Int., 2016, 26, 483–486.
P. Kapur, J.-T. Hsieh and J. Zheng, Angew. Chem., Int. Ed., 472 A. Shivhare, S. J. Ambrose, H. Zhang, R. W. Purves and
2016, 55, 2787–2791. R. W. J. Scott, Chem. Commun., 2013, 49, 276–278.
446 M. Yu and J. Zheng, ACS Nano, 2015, 9, 6655–6674. 473 A. Shivhare, D. M. Chevrier, R. W. Purves and
447 S. Sun, X. Ning, G. Zhang, Y.-C. Wang, C. Peng and R. W. J. Scott, J. Phys. Chem. C, 2013, 117, 20007–20016.
J. Zheng, Angew. Chem., Int. Ed., 2016, 55, 2421– 474 D. Hu, S. Jin, Y. Shi, X. Wang, R. W. Graff, W. Liu, M. Zhu
2424. and H. Gao, Nanoscale, 2017, 9, 3629–3636.
448 M. A. H. Muhammed, P. K. Verma, S. K. Pal, 475 H. Chen, C. Liu, M. Wang, C. Zhang, G. Li and F. Wang,
R. C. A. Kumar, S. Paul, R. V. Omkumar and T. Pradeep, Chin. J. Catal., 2016, 37, 1787–1793.
Chem. – Eur. J., 2009, 15, 10110–10120. 476 V. Sudheeshkumar, A. Shivhare and R. W. J. Scott, Catal.
449 M. Sakamoto, T. Tachikawa, M. Fujitsuka and T. Majima, Sci. Technol., 2017, 7, 272–280.
Langmuir, 2009, 25, 13888–13893. 477 Y. Tan, X. Yan Liu, L. Zhang, A. Wang, L. Li, X. Pan,
450 E. Delamarche, B. Michel, H. A. Biebuyck and C. Gerber, S. Miao, M. Haruta, H. Wei, H. Wang, F. Wang, X. Wang
Adv. Mater., 1996, 8, 719–729. and T. Zhang, Angew. Chem., Int. Ed., 2017, 56, 2709–
451 M. J. Hostetler and R. W. Murray, Curr. Opin. Colloid 2713.
Interface Sci., 1997, 2, 42–50. 478 Y. Zhu, Z. Wu, C. Gayathri, H. Qian, R. R. Gil and R. Jin,
452 K. Zheng, M. I. Setyawati, D. T. Leong and J. Xie, ACS J. Catal., 2010, 271, 155–160.
Nano, 2017, 11, 6904–6910. 479 Y. Zhu, H. Qian, B. A. Drake and R. Jin, Angew. Chem., Int.
453 M. Haruta, Chem. Rec., 2003, 3, 75–87. Ed., 2010, 49, 1295–1298.
454 H. Li, L. Li and Y. Li, Nanotechnol. Rev., 2013, 2, 480 R. Ouyang and D.-e. Jiang, ACS Catal., 2015, 5, 6624–6629.
515–528. 481 G. Li, H. Abroshan, Y. Chen, R. Jin and H. J. Kim, J. Am.
455 K. D. Gilroy, A. Ruditskiy, H.-C. Peng, D. Qin and Y. Xia, Chem. Soc., 2015, 137, 14295–14304.
Chem. Rev., 2016, 116, 10414–10472. 482 C. C. C. J. Seechurn, M. O. Kitching, T. J. Colacot and
456 Y. Zhu, H. Qian, A. Das and R. Jin, Chin. J. Catal., 2011, V. Snieckus, Angew. Chem., Int. Ed., 2012, 51, 5062–5085.
32, 1149–1155. 483 G. Li and R. Jin, Nanotechnol. Rev., 2013, 2, 529–545.
457 D. R. Kauffman, D. Alfonso, C. Matranga, G. Li and R. Jin, 484 G. Li, D.-e. Jiang, C. Liu, C. Yu and R. Jin, J. Catal., 2013,
J. Phys. Chem. Lett., 2013, 4, 195–202. 306, 177–183.
458 H. Chong and M. Zhu, ChemCatChem, 2015, 7, 2296–2304. 485 H. Abroshan, G. Li, J. Lin, H. J. Kim and R. Jin, J. Catal.,
459 D. R. Kauffman, D. Alfonso, C. Matranga, P. Ohodnicki, 2016, 337, 72–79.
X. Deng, R. C. Siva, C. Zeng and R. Jin, Chem. Sci., 2014, 5, 486 S. Antonello, M. Hesari, F. Polo and F. Maran, Nanoscale,
3151–3157. 2012, 4, 5333–5342.
460 X. Yuan, N. Goswami, I. Mathews, Y. Yu and J. Xie, Nano 487 Z. Liu, Q. Xu, S. Jin, S. Wang, G. Xu and M. Zhu,
Res., 2015, 8, 3488–3495. Int. J. Hydrogen Energy, 2013, 38, 16722–16726.
461 J. Li, R. R. Nasaruddin, Y. Feng, J. Yang, N. Yan and J. Xie, 488 Y. Adachi, H. Kawasaki, T. Nagata and Y. Obora, Chem.
Chem. – Eur. J., 2016, 22, 14816–14820. Lett., 2016, 45, 1457–1459.
462 Y. Liu, Y. Zheng, B. Du, R. R. Nasaruddin, T. Chen and 489 N. Sakai and T. Tatsuma, Adv. Mater., 2010, 22, 3185–
J. Xie, Ind. Eng. Chem. Res., 2017, 56, 2999–3007. 3188.
463 M. Haruta, T. Kobayashi, H. Sano and N. Yamada, Chem. 490 C. Yu, G. Li, S. Kumar, H. Kawasaki and R. Jin, J. Phys.
Lett., 1987, 405–408. Chem. Lett., 2013, 4, 2847–2852.
464 X. Nie, H. Qian, Q. Ge, H. Xu and R. Jin, ACS Nano, 2012, 491 A. Kogo, N. Sakai and T. Tatsuma, Electrochem. Commun.,
6, 6014–6022. 2010, 12, 996–999.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10831
View Article Online

Review Nanoscale

492 M. Lee, P. Amaratunga, J. Kim and D. Lee, J. Phys. Chem. 517 S. Bhat, A. Baksi, S. K. Mudedla, G. Natarajan,
C, 2010, 114, 18366–18371. V. Subramanian and T. Pradeep, J. Phys. Chem. Lett., 2017,
493 Y. Negishi, M. Mizuno, M. Hirayama, M. Omatoi, 8, 2787–2793.
T. Takayama, A. Iwase and A. Kudo, Nanoscale, 2013, 5, 518 K. R. Krishnadas, D. Ghosh, A. Ghosh, G. Natarajan and
7188–7192. T. Pradeep, J. Phys. Chem. C, 2017, 121, 23224–23232.
494 W. Chen and S. Chen, Angew. Chem., Int. Ed., 2009, 48, 519 K. R. Krishnadas, A. Baksi, A. Ghosh, G. Natarajan and
4386–4389. T. Pradeep, ACS Nano, 2017, 11, 6015–6023.
495 D. R. Kauffman, D. Alfonso, C. Matranga, H. Qian and 520 C. P. Joshi, M. S. Bootharaju, M. J. Alhilaly and
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

R. Jin, J. Am. Chem. Soc., 2012, 134, 10237–10243. O. M. Bakr, J. Am. Chem. Soc., 2015, 137, 11578–
496 K. Kwak, U. P. Azad, W. Choi, K. Pyo, M. Jang and D. Lee, 11581.
ChemElectroChem, 2016, 3, 1253–1260. 521 H. Yang, J. Lei, B. Wu, Y. Wang, M. Zhou, A. Xia, L. Zheng
497 K. Rajeshwar, N. R. D. Tacconi and and N. Zheng, Chem. Commun., 2013, 49, 300–302.
C. R. Chenthamarakshan, Chem. Mater., 2001, 13, 2765– 522 X.-T. Shen, X.-L. Ma, Q.-L. Ni, M.-X. Ma, L.-C. Gui, C. Hou,
2782. R.-B. Hou and X.-J. Wang, Nanoscale, 2018, 10, 515–519.
498 J. Shen, W. Zhang, R. Qi, Z.-W. Mao and H. Shen, Chem. 523 H. Yang, Y. Wang and N. Zheng, Nanoscale, 2013, 5, 2674–
Soc. Rev., 2017, 47, 1969–1995. 2677.
499 A. M. Dennis, J. B. Delehanty and I. L. Medintz, J. Phys. 524 L. G. AbdulHalim, M. S. Bootharaju, Q. Tang,
Chem. Lett., 2016, 7, 2139–2150. S. D. Gobbo, R. G. AbdulHalim, M. Eddaoudi, D.-e. Jiang
500 S. Chowdhury, Z. Wu, A. Jaquins-Gerstl, S. Liu, and O. M. Bakr, J. Am. Chem. Soc., 2015, 137, 11970–
A. Dembska, B. A. Armitage, R. Jin and L. A. Peteanu, 11975.
J. Phys. Chem. C, 2011, 115, 20105–20112. 525 R. S. Dhayal, J.-H. Liao, Y.-C. Liu, M.-H. Chiang, S. Kahlal,
501 M. A. H. Muhammed and T. Pradeep, Small, 2011, 7, 204– J.-Y. Saillard and C. W. Liu, Angew. Chem., Int. Ed., 2015,
208. 54, 3702–3706.
502 J. K. Kim and D.-J. Jang, J. Mater. Chem. C, 2017, 5, 6037– 526 C. Liu, T. Li, H. Abroshan, Z. Li, C. Zhang, H. J. Kim, G. Li
6046. and R. Jin, Nat. Commun., 2018, 9, 744.
503 A. Ghosh, T. Pradeep and J. Chakrabarti, J. Phys. Chem. C, 527 W. Du, S. Jin, L. Xiong, M. Chen, J. Zhang, X. Zou, Y. Pei,
2014, 118, 13959–13964. S. Wang and M. Zhu, J. Am. Chem. Soc., 2017, 139, 1618–
504 N. S. Han, D. Kim, J. W. Lee, J. Kim, H. S. Shim, Y. Lee, 1624.
D. Lee and J. K. Song, ACS Appl. Mater. Interfaces, 2016, 8, 528 A. Desireddy, B. E. Conn, J. Guo, B. Yoon, R. N. Barnett,
1067–1072. B. M. Monahan, K. Kirschbaum, W. P. Griffith,
505 N. Mondal, S. Paul and A. Samanta, Nanoscale, 2016, 8, R. L. Whetten, U. Landman and T. P. Bigioni, Nature,
14250–14256. 2013, 501, 399–402.
506 L. F. Hartje, B. Munsky, T. W. Ni, C. J. Ackerson and 529 M. J. Alhilaly, M. S. Bootharaju, C. P. Joshi, T. M. Besong,
C. D. Snow, J. Phys. Chem. B, 2017, 121, 7652–7659. A.-H. Emwas, R. Juarez-Mosqueda, S. Kaappa, S. Malola,
507 L. Liu, Y. Song, H. Chong, S. Yang, J. Xiang, S. Jin, K. Adil, A. Shkurenko, H. Häkkinen, M. Eddaoudi and
X. Kang, J. Zhang, H. Yu and M. Zhu, Nanoscale, 2016, 8, O. M. Bakr, J. Am. Chem. Soc., 2016, 138, 14727–14732.
1407–1412. 530 H. Yang, Y. Wang, X. Chen, X. Zhao, L. Gu, H. Huang,
508 Y. Luo, S. Fan, W. Yu, Z. Wu, D. A. Cullen, C. Liang, J. Shi J. Yan, C. Xu, G. Li, J. Wu, A. J. Edwards, B. Dittrich,
and C. Su, Adv. Mater., 2018, 30, 1704576. Z. Tang, D. Wang, L. Lehtovaara, H. Häkkinen and
509 H. Qian, D.-e. Jiang, G. Li, C. Gayathri, A. Das, R. R. Gil N. Zheng, Nat. Commun., 2016, 7, 12809.
and R. Jin, J. Am. Chem. Soc., 2012, 134, 16159–16162. 531 S.-F. Yuan, P. Li, Q. Tang, X.-K. Wan, Z.-A. Nan, D.-e. Jiang
510 M. Zhu, G. Chan, H. Qian and R. Jin, Nanoscale, 2011, 3, and Q.-M. Wang, Nanoscale, 2017, 9, 11405–11409.
1703–1707. 532 M. Qu, H. Li, L.-H. Xie, S.-T. Yan, J.-R. Li, J.-H. Wang,
511 M. A. H. Muhammed and T. Pradeep, Chem. Phys. Lett., C.-Y. Wei, Y.-W. Wu and X.-M. Zhang, J. Am. Chem. Soc.,
2007, 449, 186–190. 2017, 139, 12346–12349.
512 M.-B. Li, S.-K. Tian, Z. Wu and R. Jin, Chem. Commun., 533 L. Ren, P. Yuan, H. Su, S. Malola, S. Lin, Z. Tang,
2015, 51, 4433–4436. B. K. Teo, H. Häkkinen, L. Zheng and N. Zheng, J. Am.
513 M.-B. Li, S.-K. Tian, Z. Wu and R. Jin, Chem. Mater., 2016, Chem. Soc., 2017, 139, 13288–13291.
28, 1022–1025. 534 H. Yang, J. Yan, Y. Wang, H. Su, L. Gell, X. Zhao, C. Xu,
514 Q. Yao, X. Yuan, Y. Yu, Y. Yu, J. Xie and J. Y. Lee, J. Am. B. K. Teo, H. Häkkinen and N. Zheng, J. Am. Chem. Soc.,
Chem. Soc., 2015, 137, 2128–2136. 2017, 139, 31–34.
515 K. R. Krishnadas, A. Ghosh, A. Baksi, I. Chakraborty, 535 R.-W. Huang, Y.-S. Wei, X.-Y. Dong, X.-H. Wu, C.-X. Du,
G. Natarajan and T. Pradeep, J. Am. Chem. Soc., 2016, 138, S.-Q. Zang and T. C. W. Mak, Nat. Chem., 2017, 9, 689–
140–148. 697.
516 K. R. Krishnadas, A. Baksi, A. Ghosh, G. Natarajan and 536 S. Li, X.-S. Du, B. Li, J.-Y. Wang, G.-P. Li, G.-G. Gao and
T. Pradeep, Nat. Commun., 2016, 7, 13447. S.-Q. Zang, J. Am. Chem. Soc., 2018, 140, 594–597.

10832 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018
View Article Online

Nanoscale Review

537 Z.-Y. Wang, M.-Q. Wang, Y.-L. Li, P. Luo, T.-T. Jia, 561 Y. Negishi, K. Munakata, W. Ohgake and K. Nobusada,
R.-W. Huang, S.-Q. Zang and T. C. W. Mak, J. Am. Chem. J. Phys. Chem. Lett., 2012, 3, 2209–2214.
Soc., 2018, 140, 1069–1076. 562 S. Wang, Y. Song, S. Jin, X. Liu, J. Zhang, Y. Pei, X. Meng,
538 M. S. Bootharaju, C. P. Joshi, M. J. Alhilaly and M. Chen, P. Li and M. Zhu, J. Am. Chem. Soc., 2015, 137,
O. M. Bakr, Chem. Mater., 2016, 28, 3292–3297. 4018–4021.
539 J. Yan, H. Su, H. Yang, S. Malola, S. Lin, H. Häkkinen and 563 N. Xia and Z. Wu, J. Mater. Chem. C, 2016, 4, 4125–4128.
N. Zheng, J. Am. Chem. Soc., 2015, 137, 11880–11883. 564 Z. Wu, Angew. Chem., Int. Ed., 2012, 51, 2934–2938.
540 M. A. Tofanelli, T. W. Ni, B. D. Phillips and C. J. Ackerson, 565 M. Walter and M. Moseler, J. Phys. Chem. C, 2009, 113,
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

Inorg. Chem., 2016, 55, 999–1001. 15834–15837.


541 M. S. Bootharaju, C. P. Joshi, M. R. Parida, 566 D.-e. Jiang and S. Dai, Inorg. Chem., 2009, 48, 2720–2722.
O. F. Mohammed and O. M. Bakr, Angew. Chem., Int. Ed., 567 K. A. Kacprzak, L. Lehtovaara, J. Akola, O. Lopez-Acevedo
2016, 55, 922–926. and H. Häkkinen, Phys. Chem. Chem. Phys., 2009, 11,
542 X. Kang, L. Xiong, S. Wang, H. Yu, S. Jin, Y. Song, T. Chen, 7123–7129.
L. Zheng, C. Pan, Y. Pei and M. Zhu, Chem. – Eur. J., 2016, 568 S. Tian, L. Liao, J. Yuan, C. Yao, J. Chen, J. Yang and
22, 17145–17150. Z. Wu, Chem. Commun., 2016, 52, 9873–9876.
543 M. S. Bootharaju, L. Sinatra and O. M. Bakr, Nanoscale, 569 Y. Negishi, W. Kurashige, Y. Kobayashi, S. Yamazoe,
2016, 8, 17333–17339. N. Kojima, M. Seto and T. Tsukuda, J. Phys. Chem. Lett.,
544 X. Liu, J. Yuan, C. Yao, J. Chen, L. Li, X. Bao, J. Yang and 2013, 4, 3579–3583.
Z. Wu, J. Phys. Chem. C, 2017, 121, 13848–13853. 570 S. L. Christensen, M. A. MacDonald, A. Chatt, P. Zhang,
545 K. L. D. M. Weerawardene and C. M. Aikens, J. Phys. H. Qian and R. Jin, J. Phys. Chem. C, 2012, 116, 26932–
Chem. C, 2018, 122, 2440–2447. 26937.
546 R. Jin, Nanoscale, 2010, 2, 343–362. 571 C. Yao, Y.-j. Lin, J. Yuan, L. Liao, M. Zhu, L.-h. Weng,
547 Z. Luo, A. W. Castleman and S. N. Khanna, Chem. Rev., J. Yang and Z. Wu, J. Am. Chem. Soc., 2015, 137, 15350–
2016, 116, 14456–14492. 15353.
548 T. Udayabhaskararao, Y. Sun, N. Goswami, S. K. Pal, 572 L. Liao, S. Zhou, Y. Dai, L. Liu, C. Yao, C. Fu, J. Yang and
K. Balasubramanian and T. Pradeep, Angew. Chem., Int. Z. Wu, J. Am. Chem. Soc., 2015, 137, 9511–9514.
Ed., 2012, 51, 2155–2159. 573 D. R. Kauffman, D. Alfonso, C. Matranga, H. Qian and
549 X. Kang, S. Wang, Y. Song, S. Jin, G. Sun, H. Yu and R. Jin, J. Phys. Chem. C, 2013, 117, 7914–7923.
M. Zhu, Angew. Chem., Int. Ed., 2016, 55, 3611– 574 S. Yamazoe, W. Kurashige, K. Nobusada, Y. Negishi and
3614. T. Tsukuda, J. Phys. Chem. C, 2014, 118, 25284–25290.
550 G. Soldan, M. A. Aljuhani, M. S. Bootharaju, 575 C. Kumara, C. M. Aikens and A. Dass, J. Phys. Chem. Lett.,
L. G. AbdulHalim, M. R. Parida, A.-H. Emwas, 2014, 5, 461–466.
O. F. Mohammed and O. M. Bakr, Angew. Chem., Int. Ed., 576 Q. Li, S. Wang, K. Kirschbaum, K. J. Lambright, A. Das
2016, 55, 5749–5753. and R. Jin, Chem. Commun., 2016, 52, 5194–5197.
551 L. He, J. Yuan, N. Xia, L. Liao, X. Liu, Z. Gan, C. Wang, 577 R. Jin, S. Zhao, C. Liu, M. Zhou, G. Panapitiya, Y. Xing,
J. Yang and Z. Wu, J. Am. Chem. Soc., 2018, 140, 3487– N. L. Rosi, J. P. Lewis and R. Jin, Nanoscale, 2017, 9,
3490. 19183–19190.
552 X. Kang, L. Xiong, S. Wang, Y. Pei and M. Zhu, Chem. 578 E. B. Guidez, V. Mäkinen, H. Häkkinen and C. M. Aikens,
Commun., 2017, 53, 12564–12567. J. Phys. Chem. C, 2012, 116, 20617–20624.
553 X. Kang, M. Zhou, S. Wang, S. Jin, G. Sun, M. Zhu and 579 J. C. Gonzalez and A. Muñoz-Castro, J. Phys. Chem. C,
R. Jin, Chem. Sci., 2017, 8, 2581–2587. 2016, 120, 27019–27026.
554 X. Dou, X. Yuan, Y. Yu, Z. Luo, Q. Yao, D. T. Leong and 580 S. Xie, H. Tsunoyama, W. Kurashige, Y. Negishi and
J. Xie, Nanoscale, 2014, 6, 157–161. T. Tsukuda, ACS Catal., 2012, 2, 1519–1523.
555 C. A. Fields-Zinna, M. C. Crowe, A. Dass, J. E. F. Weaver 581 C. Yao, J. Chen, M.-B. Li, L. Liu, J. Yang and Z. Wu, Nano
and R. W. Murray, Langmuir, 2009, 25, 7704–7710. Lett., 2015, 15, 1281–1287.
556 Y. Negishi, W. Kurashige, Y. Niihori, T. Iwasa and 582 K. Kwak, W. Choi, Q. Tang, M. Kim, Y. Lee, D.-e. Jiang and
K. Nobusada, Phys. Chem. Chem. Phys., 2010, 12, 6219– D. Lee, Nat. Commun., 2017, 8, 14723.
6225. 583 G. Hu, Q. Tang, D. Lee, Z. Wu and D.-e. Jiang, Chem.
557 Y. Niihori, Y. Kikuchi, A. Kato, M. Matsuzaki and Mater., 2017, 29, 4840–4847.
Y. Negishi, ACS Nano, 2015, 9, 9347–9356. 584 Y. Niihori, W. Kurashige, M. Matsuzaki and Y. Negishi,
558 A. Sels, N. Barrabés, S. Knoppe and T. Bürgi, Nanoscale, Nanoscale, 2013, 5, 508–512.
2016, 8, 11130–11135. 585 Y. Niihori, M. Eguro, A. Kato, S. Sharma, B. Kumar,
559 Y. Negishi, T. Iwai and M. Ide, Chem. Commun., 2010, 46, W. Kurashige, K. Nobusada and Y. Negishi, J. Phys. Chem.
4713–4715. C, 2016, 120, 14301–14309.
560 X. Dou, X. Yuan, Q. Yao, Z. Luo, K. Zheng and J. Xie, 586 M. Zhou, H. Qian, M. Y. Sfeir, K. Nobusada and R. Jin,
Chem. Commun., 2014, 50, 7459–7462. Nanoscale, 2016, 8, 7163–7171.

This journal is © The Royal Society of Chemistry 2018 Nanoscale, 2018, 10, 10758–10834 | 10833
View Article Online

Review Nanoscale

587 M. Zhou, C. Yao, M. Y. Sfeir, T. Higaki, Z. Wu and R. Jin, 603 S. Wang, H. Abroshan, C. Liu, T.-Y. Luo, M. Zhu,
J. Phys. Chem. C, 2018, DOI: 10.1021/acs.jpcc.7b11057. H. J. Kim, N. L. Rosi and R. Jin, Nat. Commun., 2017, 8,
588 S. Yang, S. Wang, S. Jin, S. Chen, H. Sheng and M. Zhu, 848.
Nanoscale, 2015, 7, 10005–10007. 604 Q. Yao, Y. Feng, V. Fung, Y. Yu, D.-e. Jiang, J. Yang and
589 S. Sharma, W. Kurashige, K. Nobusada and Y. Negishi, J. Xie, Nat. Commun., 2017, 8, 1555.
Nanoscale, 2015, 7, 10606–10612. 605 X. Kang, L. Xiong, S. Wang, Y. Pei and M. Zhu, Inorg.
590 N. Yan, L. Liao, J. Yuan, Y.-j. Lin, L.-h. Weng, J. Yang and Chem., 2018, 57, 335–342.
Z. Wu, Chem. Mater., 2016, 28, 8240–8247. 606 N. S. Khetrapal, S. S. Bulusu and X. C. Zeng, J. Phys. Chem.
Published on 22 May 2018. Downloaded by Universite Grenoble Alpes INP on 11/16/2023 11:03:36 AM.

591 S. Sharma, S. Yamazoe, T. Ono, W. Kurashige, Y. Niihori, A, 2017, 121, 2466–2474.


K. Nobusada, T. Tsukuda and Y. Negishi, Dalton Trans., 607 W. W. Xu, Y. Li, Y. Gao and X. C. Zeng, Nanoscale, 2016, 8,
2016, 45, 18064–18068. 7396–7401.
592 L. Yang, H. Cheng, Y. Jiang, T. Huang, J. Bao, Z. Sun, 608 S. Pande, W. Huang, N. Shao, L.-M. Wang, N. Khetrapal,
Z. Jiang, J. Ma, F. Sun, Q. Liu, T. Yao, H. Deng, S. Wang, W.-N. Mei, T. Jian, L.-S. Wang and X. C. Zeng, ACS Nano,
M. Zhu and S. Wei, Nanoscale, 2015, 7, 14452–14459. 2016, 10, 10013–10022.
593 T. Huang, L. Huang, Y. Jiang, F. Hu, Z. Sun, G. Pan and 609 Z. Ma, P. Wang and Y. Pei, Nanoscale, 2016, 8, 17044–
S. Wei, Dalton Trans., 2017, 46, 12239–12244. 17054.
594 T. Yao, Z. Sun, Y. Li, Z. Pan, H. Wei, Y. Xie, M. Nomura, 610 P. Wang, X. Sun, X. Liu, L. Xiong, Z. Ma and Y. Pei, J. Phys.
Y. Niwa, W. Yan, Z. Wu, Y. Jiang, Q. Liu and S. Wei, J. Am. Chem. Lett., 2017, 8, 1248–1252.
Chem. Soc., 2010, 132, 7696–7701. 611 C. Zeng, C. Liu, Y. Pei and R. Jin, ACS Nano, 2013, 7,
595 Y. Li, H. Cheng, T. Yao, Z. Sun, W. Yan, Y. Jiang, Y. Xie, 6138–6145.
Y. Sun, Y. Huang, S. Liu, J. Zhang, Y. Xie, T. Hu, L. Yang, 612 J. Chai, H. Chong, S. Wang, S. Yang, M. Wu and M. Zhu,
Z. Wu and S. Wei, J. Am. Chem. Soc., 2012, 134, 17997–18003. RSC Adv., 2016, 6, 111399–111405.
596 T.-A. D. Nguyen, Z. R. Jones, B. R. Goldsmith, 613 C. Zeng, Y. Chen, A. Das and R. Jin, J. Phys. Chem. Lett.,
W. R. Buratto, G. Wu, S. L. Scott and T. W. Hayton, J. Am. 2015, 6, 2976–2986.
Chem. Soc., 2015, 137, 13319–13324. 614 X. Kang, C. Silalai, Y. Lv, G. Sun, S. Chen,
597 M. S. Bootharaju, S. M. Kozlov, Z. Cao, M. Harb, N. Maity, H. Yu, F. Xu and M. Zhu, Eur. J. Inorg. Chem., 2017, 1414–
A. Shkurenko, M. R. Parida, M. N. Hedhili, M. Eddaoudi, 1419.
O. F. Mohammed, O. M. Bakr, L. Cavallo and J.-M. Basset, 615 Y. Chen, C. Zeng, D. R. Kauffman and R. Jin, Nano Lett.,
J. Am. Chem. Soc., 2017, 139, 1053–1056. 2015, 15, 3603–3609.
598 M. S. Bootharaju, S. M. Kozlov, Z. Cao, M. Harb, 616 X.-K. Wan, J.-Q. Wang, Z.-A. Nan and Q.-M. Wang, Sci.
M. R. Parida, M. N. Hedhili, O. F. Mohammed, Adv., 2017, 3, e1701823.
O. M. Bakr, L. Cavallo and J.-M. Basset, Nanoscale, 2017, 617 V. Marjomäki, T. Lahtinen, M. Martikainen, J. Koivisto,
9, 9529–9536. S. Malola, K. Salorinne, M. Pettersson and H. Häkkinen,
599 X. Kang, J. Xiang, Y. Lv, W. Du, H. Yu, S. Wang and Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 1277–1281.
M. Zhu, Chem. Mater., 2017, 29, 6856–6862. 618 S. Choi, R. M. Dickson and J. Yu, Chem. Soc. Rev., 2012,
600 H. Shen and T. Mizuta, Chem. – Asian J., 2017, 12, 2904– 41, 1867–1891.
2907. 619 X.-D. Zhang, Z. Luo, J. Chen, H. Wang, S.-S. Song, X. Shen,
601 H. Shen and T. Mizuta, Chem. – Eur. J., 2017, 23, 17885– W. Long, Y.-M. Sun, S. Fan, K. Zheng, D. T. Leong and
17888. J. Xie, Small, 2015, 11, 1683–1690.
602 C. Liu, G. Li, D. R. Kauffman, G. Pang and R. Jin, 620 B. Haraldsson, J. Nyström and W. M. Deen, Physiol. Rev.,
J. Colloid Interface Sci., 2014, 423, 123–128. 2008, 88, 451–487.

10834 | Nanoscale, 2018, 10, 10758–10834 This journal is © The Royal Society of Chemistry 2018

You might also like