Next Article in Journal
Preparation of a Stable Nanoscale Manganese Residue-Derived FeS@Starch-Derived Carbon Composite for the Adsorption of Safranine T
Next Article in Special Issue
Gold Nanoclusters: Bridging Gold Complexes and Plasmonic Nanoparticles in Photophysical Properties
Previous Article in Journal
Guided Mode Resonance Sensors with Optimized Figure of Merit
Previous Article in Special Issue
Catenane Structures of Homoleptic Thioglycolic Acid-Protected Gold Nanoclusters Evidenced by Ion Mobility-Mass Spectrometry and DFT Calculations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heterogeneous Cross-Coupling over Gold Nanoclusters

1
College of Science, Inner Mongolia Agricultural University, Hohhot 010018, China
2
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(6), 838; https://doi.org/10.3390/nano9060838
Submission received: 23 April 2019 / Revised: 8 May 2019 / Accepted: 12 May 2019 / Published: 1 June 2019

Abstract

:
Au clusters with the precise numbers of gold atoms, a novel nanogold material, have recently attracted increasing interest in the nanoscience because of very unique and unexpected properties. The unique interaction and electron transfer between gold clusters and reactants make the clusters promising catalysts during organic transformations. The AunLm nanoclusters (where L represents organic ligands and n and m mean the number of gold atoms and ligands, respectively) have been well investigated and developed for selective oxidation, hydrogenation, photo-catalysis, and so on. These gold clusters possess unique frameworks, providing insights into the catalytic processes and an excellent arena to correlate the atomic frameworks with their intrinsic catalytic properties and to further investigate the tentative reaction mechanisms. This review comprehensively summarizes the very latest advances in the catalytic applications of the Au nanoclusters for the C−C cross-coupling reactions, e.g., Ullmann, Sonogashira, Suzuki cross-couplings, and A3−coupling reactions. It is found that the proposed catalytically active sites are associated with the exposure of gold atoms on the surface of the metal core when partial capping organic ligands are selectively detached under the reaction conditions. Finally, the tentative catalytic mechanisms over the ligand-capped Au nanoclusters and the relationship of structure and catalytic performances at the atomic level using computational methods are explored in detail.

1. Introduction

Since the work of Haruta’s group in the late 1980s [1], supported gold nanoparticles with a particle size in the range of 3–20 nm have played a central role in a variety of reactions such as selective oxidation [2,3,4,5], hydrogenation [6,7,8], and photocatalysis [9,10]. These conventional gold catalysts have been realized via deposition–precipitation and co-precipitation impregnation with oxides by controlling the pH of the synthetic system. These obtained gold nanoparticles are usually polydisperse, which is a major issue in fundamental catalysis and investigations [11]. For example, the size-hierarchy of Au nanoparticles are often averaged out in polydispersion. It is difficult to correlate the relationship between the catalytic properties and the structure of the nanoparticles. Therefore, developing nanostructured catalysts with specific morphology (e.g., nanosheet, nanocube, and nanorod) is highly desirable to overcome this issue, further promoting the development of crystal-identified model catalysts [12,13].
On the other hand, the remarkable developments in the synthesis of atomically precise gold nanoclusters have been achieved in recent decades, which opens a new burgeoning area in nanoscience [14,15]. These gold nanoclusters are comprised of a few dozen to a few hundred gold atoms and protecting organic ligands (e.g., thiolate, phosphine, and alkyne), and the size is ultra-small, usually ca. 0.6–2 nm. Some gold nanoclusters are identifiable by X-ray crystallography technology. It offers a big opportunity for in-depth understanding of the relationship of the catalytic properties and the active-site structure at the atomic level [16].
It is observed that these gold nanoclusters exhibit good catalytic performance in heterogeneous catalysis (e.g., selective oxidation and hydrogenation) [16,17,18]. Sometime, the gold clusters show better catalytic behavior (e.g., activity and product selectivity) than the corresponding Au nanoparticles, because of their high surface-to-volume ratio (reaches up to ~100%), surface geometric effect (e.g., low-coordinated Auδ+ atoms, 0 < δ < 1), and the unique electronic properties and the quantum size effect. Furthermore, the protecting organic ligands can improve the product selectivity due to their electronic factors and steric hindrance and weak interaction (e.g., π-π interaction) between the reactants and cluster surface ligands during the catalysis process. In the recent decade, these gold nanoclusters exhibited good catalytic activity in the cross C–C couplings, e.g., Ullmann hetero-coupling, Suzuki and Sonogashira coupling, and A3-coupling [19,20]. Traditionally, these catalyzed C–C coupling reactions are over the Cu, Pd, and Pt complexes and particles in the previous literatures [21,22,23].
In this review, we aim to provide an overview focused on the Au nanocluster-catalyzed coupling reactions, e.g., Ullmann hetero-coupling of Ar-I, Suzuki cross-coupling of PhB(OH)2 and Ph-I (IB), Sonogashira cross-coupling of IB and Ph–C≡C–H (PA), and A3-coupling (Scheme 1). The pathway of the selective detachment of the surface protecting ligands (under the reaction conditions), giving the catalytically active sites, is well discussed. Moreover, the proposed reaction pathway and mechanisms of these carbon–carbon coupling reactions were thoroughly summarized based on the precise framework of gold nanoclusters (e.g., two Au25) as the mode theoretical calculations.

2. Activation of Ph–B(OH)2, Ph–I, and C≡C–H Bonds over Au: Theoretical Simulation

The activation of the Ph–B(OH)2, Ph–I, and C≡C–H bonds over the gold clusters plays an important role and step in the cross-coupling reactions, e.g., Ullmann hetero-coupling, Sonogashira coupling, Suzuki coupling, and A3−coupling. The adsorption and activation process of the reactants over the different well-defined gold facets and sites are distinct [24,25,26,27]. The density functional theory (DFT) calculations should be a feasible and fast method to explore the activation of Ph–B(OH)2, C–I, and C≡C–H bond by gold nanoclusters.
Firstly, the shape controlled Au nanoparticles (e.g., Au nanorod) with well-defined surfaces and morphologies are well investigated [28,29], providing a platform to establish site–activity relationships and to pursue the understanding of heterogeneous processes. Therefore, the gold nanorod is chosen for the theoretical simulation of the activation of Ph–I and C≡C–H bonds. DFT studies showed that the Ph-I reactant adsorbs onto the Au(100) and Au(111) with Au–I distance 2.95 and 2.86 Å. The C–I bond is more elongated on the Au(111) compared to Au(100) (2.17 Å vs. 2.13 Å) [26]; the C–I length of the free IB is 2.09 Å. Furthermore, the phenyl group is located on an Au atom. The iodine atom is strongly chemisorbed on the bridging and hollow sites. Regard to the activation of alkyne, the Ph–C≡C– fragment is also adsorbed on bridging and 3-fold hollow sites of Au(100) and Au(111). The coordination number of one Ph–C≡C– unit is 3 and 4 for Au(100) and Au(111) facets during the cross-coupling reaction, respectively, Figure 1, and the activation energy of cross-coupling is comparable [26].
Next, the activation of Ph−B(OH)2, Ph−I, and C≡C−H also was explored on a model cluster (Au38 and partially oxidized Au38O2) [30,31,32]. It is worthy to note that the model clusters of the Au38 and Au38O2 are somewhat different from the real catalyst in terms of the real structure. The Au38O2 cluster contains metallic Au0 and cationic Auδ+ species, and each O atom bonds to three Au atoms. In the activation process of phenylboronate, DFT showed that the phenylboronate reactant is preferentially adsorbed on the Au0 species rather than on the Auδ+ sites [30], leading to formation of the final product- biphenyl (Figure 2B). Meanwhile, DFT simulation also found that the interaction between the Auδ+ sites and Ph−I is weaker and the adsorption of C≡C−H on Auδ+ sites is relatively strong. Further, the proton of alkyne was detached with aid of O atom in the deprotonation step (Figure 2A). Therefore, a very low activation energy is required when the cross-coupling reactions occurred over the Au38O2 cluster. Of note, the cross-coupling step of the activated PA and –Ph on the Au38O2 should be the rate-determining step [32].

3. Physical Property of Au Nanoclusters

3.1. Framework

The gold clusters with crystal structure can be employed as the practical simulation models for mechanism study. In this Review, we only focus on the two cluster structures of the nanorod-shaped [Au25(PPh3)10L5Cl2]2+ (L = −SR and PA) and the nanosphere [Au25(SR)18]x (x = −1, 0, +1, etc.), used as the real model for DFT studies (vide infra). The Au25(SR)18 cluster comprises an Au13 core [33] and six staples of Au2(SR)3 [34]. [Au25(PPh3)10L5Cl2]2+ is composed of two Au13 cores by sharing one common vertex to form the waist sites [35], connected by thiolate or alkyne ligands [36] (Figure 3).

3.2. Redox Property of Au cluster

Au38S2(SAdm)20 (SAdm = adamantanethiolate) nanoclusters exerted photosensitizing properties to give singlet oxygen (1O2) under visible light irradiation (e.g., 532 and 650 nm) [37]. The Au38S2(SAdm)20 is intact during the whole photocatalysis process, evidenced by UV-vis tracing and mass spectroscopy analysis. The cyclic voltammetry analysis showed that the Au38S2(SAdm)20 cluster had good charge transfer capacity to the redox K3Fe(CN)6 probe, Figure 4 [38]. However, the redox property of the cluster disappeared when β-cyclodextrins (β-CDs) was introduced in the THF solution. After detailed analysis, the huge β-CDs “umbrella” can trap the adamantane groups and then completely cover windows of the Au38S2(SAdm)20 nanoclusters, thereby blocking direct interaction with foreign molecules and then quenching the charge transfer process (Figure 4B). It indicated that these gold nanoclusters have good redox properties and electron transfer (ET) capacity during the catalytic reactions [39].
Further, Kumar and coworkers studied the Au25(SG)18 catalyst in an electrochemical oxidation [40]. The Au25(SG)18 on the electrode gave good electro-activity during the oxidation of ascorbic acid and dopamine over a wide linear range from 0.71 to 44.4 µM. And pH dependent electrocatalytic activity was observed, attributed to the consequence of pH-dependent electrostatic attraction/repulsion between the charged Au25(SG)18 clusters and the charged analytes. Moreover, an amperometric sensing method for other compounds was developed. Next, Kauffman et al. investigated the electron transfer between CO2 and Au25(PET)18 in solution [41]. Upon the DMF solution (containing Au25(SR)18) was saturated with CO2 gas, the optical absorbance features showed the oxidized state of Au25(SR)18. Meanwhile, the photoluminescence increases and blue-shift. The CO2 induced optical changes can be simply reversed by purging the solution with N2 gas to remove the CO2, indicating an interaction between Au25(SR)18 and CO2. DFT calculations revealed that the CO2 molecule interacts with three S atoms of the Au3 site of the Au25(SR)18 cluster, prompting the CO2 electrochemical reduction. These observed unique interactions and electron transfers between gold clusters and reactants make the clusters promising catalysts during the organic transformations.

4. Catalytic Properties

4.1. Ullmann Coupling

At the beginning, the catalytic activity of the Au nanoclusters was examined in the Ullmann homo-coupling reactions of aryl iodides, which are generally catalyzed by palladium, nickel, and copper catalysts [42]. The supported gold cluster catalysts were simply prepared by a vortex-mixing of supports and a solution containing gold clusters at room temperature, and an annealing at 150 °C. The supported Au clusters were intact after the 150 °C annealing process (higher than the reaction temperatures), evidenced by UV−vis and scanning transmission electron microscopy (STEM) [43]. The X-ray photoelectron spectroscopy (XPS) analysis shows the chemical state of Au species in the oxide-supported cluster catalysts is positively charged (Auδ+) [44], where 0 < δ < 1, consistent with the free gold nanoclusters. The catalytic processes were carried out at 130 °C in the presence of base, which is similar with these catalyzed by Pd/Cu complexes or nanoparticles. The Au25(SR)18/CeO2 showed the best catalytic activity, and the test was then expanded to a serial of substituents with functional side-groups (Table 1) [45]. Of note, the efficiency of gold nanoclusters was not as good as the palladium, nickel, and copper nanocomposites in the Ullmann homo-coupling reactions of aryl chlorides and aryl bromides.
Later, these Au cluster were studied in the Ullmann hetero-coupling reactions. The catalytic conditions over the Au25(SR)18/CeO2 catalysts were the same with the homo-coupling reactions (Table 1 vs. Table 2). The aromatic and aliphatic thiolate-capped Au25 nanoclusters (e.g., naphthalenethiolate (-SNap), benzenethiolate (-SPh), hexanethiolate (-SC6H13), and 2-phenylethanethiolate (PET)) were chosen for comparison and exploration in the Ullmann hetero-coupling of 4-MeC6H4I and 4-NO2C6H4I [46]. Intriguingly, the aromatic thiolate ligated Au25 clusters gave much better catalytic performance (both the conversion of NO2C6H4I and selectivity for the hetero-coupling product (4-methyl-4′-nitrobiphenyl) than these protected by alkyl thiolate ligands. The Au25(SNap)18 cluster gave an 82% selectivity toward the hetero-coupling product, which was much higher than the Cu, Pd, and Au complexes (the selectivity: <30%, Table 2). Unfortunately, both of the conversion and selectivity decreased in the 2nd and 3rd cycles, which was due to the removal of the capping surface ligands and hence the decomposition of Au clusters, evidenced by the TEM images. These large gold nanoparticles jeopardized the catalytic performance in this coupling reaction. Thus, the protecting ligands on the clusters’ surface play a key influence on their catalytic properties.
DFT simulations were applied to explain the catalytic results. It is worthy to note that the reactants of both 4-MeC6H4I and 4-NO2C6H4I cannot interact well with the intact Au25(SR)18 clusters, because of the steric effect of the protecting thiolate ligands on the clusters’ surface. In the first step, one “-SR” unit on the Au25(SR)18 cluster was surmised to be detached under the reaction conditions in the presence of a K2CO3 base. Then the gold atoms on the motif were exposed to reactants and were associated with the catalytic sites [44]. Further, the activation energy for the homo- and hetero-couplings over the Au25 protected by “-SCH3” thiolate were compared by the nudged elastic band (NEB) approach (Figure 5). Intriguingly, the activation energy in the hetero-coupling was less than in the homo-coupling in the case of Au25−SNap clusters, (Figure 4). It implied that the aromatic thiolate-capped gold cluster can not only improve the conversion rate but can also favor the hetero-coupled process [46].
DFT simulations were applied to explain the catalytic results. It is worthy to note that the reactants of both 4-MeC6H4I and 4-NO2C6H4I cannot interact well with the intact Au25(SR)18 clusters, because of the steric effect of the protecting thiolate ligands on the clusters’ surface. In the first step, one “-SR” unit on the Au25(SR)18 cluster was surmised to be detached under the reaction conditions in the presence of K2CO3 base. Then the gold atoms on the motif were exposed to reactants and were associated with the catalytic sites [44]. Further, the activation energy for the homo- and hetero-couplings over the Au25 protected by “-SCH3” thiolate are comparable by the nudged elastic band (NEB) approach (Figure 5). Intriguingly, the activation energy in the hetero-coupling less than in the homo-coupling in the case of Au25−SNap clusters, (Figure 4). It implied that the aromatic thiolate-capped gold cluster not only can improve the conversion rate but also can favor the hetero-coupled process [46].

4.2. Suzuki Coupling

Further, the titania-supported Au25 clusters are studied in the Suzuki coupling in the presence of ionic liquids (ILs), which are catalyzed over palladium catalysts [42]. The Suzuki cross-coupling run at 90 °C using different solvents (e.g., ethanol, xylene, toluene, N,N′-dimethylformamide (DMF), ILs, etc.). The imidazolium-based ILs exerted a large effect on the MeOC6H4I conversion to the desire products. A very low conversion (<5%) is observed when using the ethanol, toluene, o-xylene, and DMF as solvents in the Au25/TiO2 catalyzed coupling reactions (Table 3). Interestingly, the iodoanisole conversion over Au25/TiO2 drastically increased to 89%–99% when BMIM·X (BMIM: 1-butyl-3-methylimidazolium, X = Br or Cl or BF4) solvents are introduced to the reaction system (Table 3). The catalytic results indicate that the imidazolium-based ILs acts as a promoter for the cross-coupling reactions [47]. Of note, only the BMIM cation (i.e., the acidic proton at position 2 of the imidazolium ions) play an important role during the reactions, as no activity is found in the presence of BDiMIM·BF4 (BDiMIM: 1-butyl-2,3-dimethylimidazolium) solvent, which is further supported by the DFT calculations. It is worthy to note that the efficiency of gold nanoclusters was not as good as the palladium nanocomposites, however, the gold nanoclusters exhibited much better selectivity toward the target cross-coupling products.
To explore the active species during the coupling reactions, the free Au25(PET)18 was mixed with the BMIM·BF4 under the same reaction conditions [43]. Except the molecular peak of Au25(PET)18 cluster, four new mass peaks are clearly detected in the matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS). These new appeared mass peaks belonged to the Au25−n(SR)18−n (where, n = 1–4) species (Figure 6). Of note, these new species are not the fragments caused by laser of the MALDI method. These species also were observed in the ESI-MS method [48]. The imidazolium-based ILs indeed assist the yield of Au25−n(SR)18−n species under the reaction conditions, which may be the active sites for the cross-coupling reactions. The other explanation is that the Au-NHC complex (NHC: N-heterocyclic carbene) with the Au25−n(SR)18−n species could be responsible for the active sites during the Suzuki cross-coupling reactions, although it needs further investigation. Of note, the Au-NHC complex was the product of the reaction of BMIM cations with the gold nanoclusters.

4.3. Sonogashira Coupling

As the IB and alkyne can be activated over gold clusters, hence, the catalytic performance of the gold nanoclusters may extend to Sonogashira cross-coupling reactions, often catalyzed over palladium catalysts [42]. The catalytic performance of the Au25(PET)18 cluster (supported on oxides) in the Sonogashira cross-coupling reaction was studied [49]. The supported catalyst was prepared by impregnating oxide powders (such as TiO2, CeO2, SiO2, and MgO) in a CH2Cl2 solution of Au25(PET)18 (~1 wt % loading) with a 150 °C annealing. STEM and TG analyses showed that the protecting thiolate ligands were intact on the surface of gold clusters after thermal treatment. Then these Au25/oxide catalysts were applied to the Sonogashira cross-coupling reaction. The optimized reaction conditions were using DMF as a solvent and K2CO3 as a base under an N2 atmosphere at 160 °C, which is harsher than those for the above Suzuki and Ullmann couplings. The Au25/CeO2 catalyst showed the best activity (96.1% iodoanisole conversion with 88.1% selectivity toward the target product) (Table 4). The solvent and base can also influence the product selectivity. The size-dependent catalytic performance also was studied.
The catalytic performance of small-sized Au25(PET)18 cluster catalysts was much better than large-sized of gold clusters of 2–3 nm and Au/CeO2 (~20 nm). Support effects were studied in the cross-coupling, and no distinct effect of the oxide supports was observed (i.e., CeO2, SiO2, TiO2, and MgO). The conversion was no obvious decrease, but the selectivity decreased from 88.1% to 64.5% after 5 cycles. It is noteworthy that TEM analysis shows that the gold clusters grow into larger nanoparticles (>3 nm), meaning that the gold clusters capped by organic ligands cannot stay intact under harsh reaction conditions (160 °C in the presence of a base). The gradual degradation of gold clusters leads to a decrease in selectivity, as the larger Au clusters showed a much lower selectivity. It is worthy to note that the efficiency of gold nanoclusters is much worse than the palladium-based catalysts, and the selectivity for the cross-coupling products over Au clusters is also worse.
DFT calculation found that the reactants (i.e., IB and PA) prefer to adsorb on the open facet (Au3) of the Au25 cluster with the phenyl ring facing a surface Au atom (Figure 7). A total adsorption energy reaches −0.90 eV when the two reactants co-adsorb on the Au25(SR)18 catalyst. While, the IB/IB pair has an adsorption energy of −1.05 eV, indicating that the IB/IB pair interacts strongly with the cluster and the homocoupling of IBs is the dominant side-reaction competing with the cross-coupling between IB and PA. DFT results suggested that the catalytic active sites is associated with the Au25(SR)18 clusters, which is consistent with the experimental results.
The structure of the 25-atom cluster is similar [50], but the electronic property and the catalytic activity of the bimetallic clusters can be largely regulated by the foreign dopants [50,51,52,53,54,55]. Recently, Li et al. [56] studied the doping effects of the Au25(SR)18 nanoclusters in the Sonogashira cross-coupling reaction base on the experiment and DFT simulations. The obtained results suggested that the Cu and Ag atoms are preferentially occupied at the cluster’s kernel (Au13) rather than the Au2(SR)3 staple motif, while a single Pt atom only can be doped individually and locates in the center of the cluster. The overall performance of AgxAu25−x(SR)18 was similar to that of Au25(SR)18 and Pt1Au24(SR)18, which showed a decrease in catalytic activity (Table 5). The catalytic activity was from AgxAu25−x(SR)18 ≈ Au25(SR)18 > CuxAu25−x(SR)18 > Pt1Au24(SR)18. Interestingly, the CuxAu25−x(SR)18 produced a homo-coupling product base on the Ullmann homo-coupling pathway, which is contrary to the other three cluster catalysts. However, DFT calculations showed that the adsorption energy of one PA molecule on the Pt1Au24(SR)18, Cu1/2Au24/23(SR)18, and Au25(SR)18 nanoclusters was very similar (−0.50 to −0.52 eV, Table 6). The adsorption energy of one IB molecule onto the Pt1Au24(SR)18, Ag1/2Au24/23(SR)18 and Au25(SR)18 was also very similar (−0.59 to −0.61 eV, Table 6). These results suggested that the adsorption process of the PA and IB onto the alloy clusters is not the key step during the coupling reactions. Generally, the catalytic activity is largely affected by the electronic effect in the core of bimetallic clusters (i.e., Pt1Au12, CuxAu13-x, AgxAu13-x, and Au13), and the selectivity of product is primarily turned by the atomic type on the shell of MxAu12-x [51,57].
The structure of the 25-atom cluster was similar [50], but the electronic property and the catalytic activity of the bimetallic clusters could be largely regulated by the foreign dopants [50,51,52,53,54,55]. Recently, Li et al. [56] studied the doping effects of the Au25(SR)18 nanoclusters in a Sonogashira cross-coupling reaction based on an experiment and DFT simulations. The obtained results suggested that the Cu and Ag atoms were preferentially occupied at the cluster’s kernel (Au13) rather than the Au2(SR)3 staple motif, while a single Pt atom only can be doped individually and locates in the center of the cluster. The overall performance of AgxAu25−x(SR)18 was similar to that of Au25(SR)18 and Pt1Au24(SR)18, which showed a decrease in catalytic activity (Table 5). The catalytic activity was AgxAu25−x(SR)18 ≈ Au25(SR)18 > CuxAu25−x(SR)18 > Pt1Au24(SR)18. Interestingly, the CuxAu25−x(SR)18 produced a homo-coupling product base on the Ullmann homo-coupling pathway, which was contrary to the other three cluster catalysts. However, DFT calculations showed that the adsorption energy of one PA molecule on the Pt1Au24(SR)18, Cu1/2Au24/23(SR)18, and Au25(SR)18 nanoclusters was very similar (−0.50 to −0.52 eV, Table 6). The adsorption energy of one IB molecule onto the Pt1Au24(SR)18, Ag1/2Au24/23(SR)18, and Au25(SR)18 was also very similar (−0.59 to −0.61 eV, Table 6). These results suggested that the adsorption process of the PA and IB onto the alloy clusters was not the key step during the coupling reactions. Generally, the catalytic activity was largely affected by the electronic effect in the core of bimetallic clusters (i.e., Pt1Au12, CuxAu13-x, AgxAu13-x, and Au13), and the selectivity of the product is primarily turned by the atomic type on the shell of MxAu12-x [51,57].

4.4. A3-Coupling

A3-coupling reactions, where three reactants (aldehydes, amines, and alkynes) react each other in one-pot to yield only one product, have attracted overwhelming interest in the past decade. The A3-coupling is favorable from environmental and economic perspectives: more efficient and less waste [58,59,60,61]. The A3-coupling reactions involve new C–C and C–N bond formation in one procedure. The alkyne activation was deemed as the key step for the A3-coupling; the aldehydes could react with amines spontaneously. Hence, the gold clusters should be active in this reaction, because the cluster catalyst has exhibited the capacity of the alkyne activation in the semi-hydrogenation of terminal alkynes [36].
The catalytic performance of the Au25(PPh3)10(PA)5X2 was investigated in different solvents and reaction temperatures [62]. The most prominent feature was that the polarity of the solvent had a great influence on catalytic activity. The TiO2-supported Au25 catalyst gave higher activity in the polar solvents. The gold clusters also showed good recyclability. Interestingly, the cluster catalyst showed no conversion using ketones as reactants (Figure 8), which was completely different from the catalytic behaviors of the gold complexes and bare gold nanoparticles. Therefore, the electronic factors and steric hindrance of the substituents had significant effects on the reaction conversion rate.
An induction period (0 to 3 h) appeared and the conversion slightly increased during the induction period. After the induction period, the reaction conversion of gold clusters significantly increased in the time evolution for A3-coupling [62]. The results showed that some phosphine ligands are removed to generate catalytic active sites, which associates with the surface gold atoms. Further research found that the capped phosphine ligands can be selectively removed in the case of Au11(PPh3)7X3 with the aid of base (e.g., pyridine), evidenced by UV-vis and ESI-MS analyses [63].
Finally, the catalytic mechanism over the gold clusters was studied by DFT calculations. Firstly, the phosphine ligand is detached in the presence of the reaction system, and then the uncovered Au atoms are exposed to the reactants. Next, the PA molecules are adsorbed onto the M1 site via the interaction of Au···whole triple bond, Figure 9. Further, a terminal hydrogen deprotonation occurrs in the presence of amine (e.g., HN(CH3)2). The iminium ion (H2C=N(CH3)2+) interacts with the PhC≡C– on site M1 and finally give rise to the final product (i.e., propargylamine) [62].
Further, Jin et al. reported Au38 nanocluster-catalyzed the A3-coupling reaction [64]. They argued that the synergistic effect of the partial positive charged Au surface (Auδ+, 0 < δ < 1) and the electron-rich Au23 kernel were responsible for the catalytic behaviors. Li et al. found that cadmium doped Au13 nanocluster also showed catalytic activity in the A3-coupling [65]. The cooperation between the exerted cadmium atoms and the neighbor gold atoms on the surface of Au13 icosahedron are tentatively deemed as the active sites in the cross-coupling reactions.

5. Summary and Outlook

In the past decade years, remarkable advances have been made in developing catalytic applications of gold nanoclusters, which is a new perspective for gold nanocatalysis, especially in carbon–carbon cross-coupling reactions. These gold nanoclusters were more efficient in the cross C-C coupling reactions in this Review. The higher catalytic activity was mainly because of the distinctive frame structure and electronic properties of the gold nanoclusters and the protecting ligands. However, some of the gold nanoclusters with the protection of thiolate/phosphine could not stay intact under the harsh reaction conditions during the catalysis processes, leading to a decrease or even disappearance of catalytic activity due to the increasing size of formed particles. How to maintain the stability of gold nanoclusters during the catalysis has become a major challenge and subject for future research.
The present review demonstrated that CeO2 and TiO2 oxides have been observed to be excellent supports for gold nanoclusters [66]. Few studies have been done on other oxides. Developing new types of supports is another significant issue for substances such as zeolite [67,68], carbon materials (e.g., graphene or graphite oxide) [69,70], and MOFs [71]. For example, mesoporous materials, e.g., MOFs and zeolite, exhibited regular tunnels and cages and improved the stability of the Au clusters during the catalytic processes [72]. The tunnels and cages, by enriching concentration of the reactants, can also improve the catalytic activity. The steric effects of MOFs organic linker can improve the selectivity of products. In addition, the sites of Lewis acids and Bronstein acids in zeolites also can change the catalytic performance of the catalysts.
In a word, gold nanoclusters will be expected to be used into carbon–heteroatom cross-coupling reactions, which are not yet well investigated. The formed C–O and C–N bonds are very useful drug intermediates [73]. Recently, some attempts have been reported in these categories, such as the cross-aldol condensation and Michael addition (forming C=C bonds) and photo-oxidation of amines to imines (C=N bonds) [10,74]. Future research on gold nanoclusters will contribute to fundamental researches and provide clues for the new design of efficient catalysts for other specific chemical processes.

Author Contributions

Investigation, Q.S. and Z.Q. (resources, Q.S. and Z.Q.; writing—original draft preparation, Q.S. and Z.Q.; writing—review and editing, H.X. and G.L.; supervision, G.L.; project administration, G.L.

Funding

This research was funded by National Natural Science Foundation of China (21665020), the Natural Science Foundation of Inner Mongolia Autonomous Region (2018BS02004), and the Program of Higher-level Talents of Inner Mongolia Agricultural University (NDYB2016-03) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. Gold catalysts prepared by coprecipitation for low-temperature oxidation of hydrogen and of carbon-monoxide. J. Catal. 1989, 115, 301–309. [Google Scholar] [CrossRef]
  2. Tsukuda, T.; Tsunoyama, H.; Sakurai, H. Aerobic oxidations catalyzed by colloidal nanogold. Chem. Asian J. 2011, 6, 736–748. [Google Scholar] [CrossRef]
  3. Della, P.C.; Falletta, E.; Rossi, M. Update on selective oxidation using gold. Chem. Soc. Rev. 2012, 41, 350–369. [Google Scholar]
  4. Li, G.; Qian, H.F.; Jin, R.C. Gold nanocluster-catalyzed selective oxidation of sulfide to sulfoxide. Nanoscale 2012, 4, 6714–6717. [Google Scholar] [CrossRef]
  5. Liu, C.; Yan, C.Y.; Lin, J.Z.; Yu, C.L.; Huang, J.H.; Li, G. One-pot synthesis of Au144(SCH2Ph)60 nanoclusters and their catalytic application. J. Mater. Chem. A 2015, 3, 20167–20173. [Google Scholar] [CrossRef]
  6. Hashmi, A.S.K.; Hutchings, G.J. Gold catalysis. Angew. Chem. Int. Ed. 2006, 45, 7896–7936. [Google Scholar] [CrossRef]
  7. Li, G.; Zeng, C.J.; Jin, R.C. Thermally robust Au99(SPh)42 nanoclusters for chemoselective hydrogenation of nitrobenzaldehyde derivatives in water. J. Am. Chem. Soc. 2014, 136, 3673–3679. [Google Scholar] [CrossRef]
  8. Li, G.; Jiang, D.E.; Kumar, S.; Chen, Y.X.; Jin, R.C. Size dependence of atomically precise gold nanoclusters in chemoselective hydrogenation and active site structure. ACS Catal. 2014, 4, 2463–2469. [Google Scholar] [CrossRef]
  9. Naya, S.; Kimura, K.; Tada, H. One-step selective aerobic oxidation of amines to imines by gold nanoparticle-loaded Rutile titanium(IV) oxide plasmon photocatalyst. ACS Catal. 2013, 3, 10–13. [Google Scholar] [CrossRef]
  10. Chen, H.J.; Liu, C.; Wang, M.; Zhang, C.F.; Luo, N.C.; Wang, Y.H.; Abroshan, H.; Li, G.; Wang, F. Visible light gold nanocluster photocatalyst: Selective aerobic oxidation of amines to imines. ACS Catal. 2017, 7, 3632–3638. [Google Scholar] [CrossRef]
  11. Taketoshi, A.; Haruta, M. Size- and structure-specificity in catalysis by gold clusters. Chem. Lett. 2014, 43, 380–387. [Google Scholar] [CrossRef]
  12. Lohse, S.E.; Murphy, C.J. The quest for shape control: A history of gold nanorod synthesis. Chem. Mater. 2013, 25, 1250–1261. [Google Scholar] [CrossRef]
  13. Lu, C.L.; Prasad, K.S.; Wu, H.L.; Ho, J.A.A.; Huang, M.H. Au nanocube-directed fabrication of Au-Pd core-shell nanocrystals with tetrahexahedral, concave octahedral, and octahedral structures and their electrocatalytic activity. J. Am. Chem. Soc. 2010, 132, 14546–14553. [Google Scholar] [CrossRef]
  14. Zhang, J.; Li, Z.; Zheng, K.; Li, G. Synthesis and characterization of size-controlled atomically-precise gold clusters. Phys. Sci. Rev. 2018, 3. [Google Scholar] [CrossRef]
  15. Jin, R.C. Atomically precise metal nanoclusters: Stable sizes and optical properties. Nanoscale 2015, 7, 1549–1565. [Google Scholar] [CrossRef]
  16. Li, G.; Jin, R.C. Atomically precise gold nanoclusters as new model catalysts. Acc. Chem. Res. 2013, 46, 1749–1758. [Google Scholar] [CrossRef]
  17. Yamazoe, S.; Koyasu, K.; Tsukuda, T. Nonscalable oxidation catalysis of gold clusters. Acc. Chem. Res. 2014, 47, 816–824. [Google Scholar] [CrossRef] [PubMed]
  18. Li, Z.M.; Abroshan, H.; Liu, C.; Li, G. A Critical review on the catalytic applications of non-metallic gold nanoclusters: Selective oxidation, hydrogenation, and coupling reactions. Curr. Org. Chem. 2017, 21, 476–488. [Google Scholar] [CrossRef]
  19. Li, G.; Jin, R.C. Catalysis by gold NPs: Carbon-carbon coupling reactions. Nanotechnol. Rev. 2013, 5, 529–545. [Google Scholar]
  20. Zhou, Y.; Li, G. A Critical review on carbon-carbon coupling over ultra-small gold nanoclusters. Acta Phys.-Chim. Sin. 2017, 33, 1297–1309. [Google Scholar]
  21. Alonso, F.; Beletskaya, I.P.; Yus, M. Non-conventional methodologies for transition-metal catalysed carbon-carbon coupling: A critical overview. Part 2: The Suzuki reaction. Tetrahedron 2008, 64, 3047–3101. [Google Scholar] [CrossRef]
  22. Seechurn, C.; Kitching, M.O.; Colacot, T.J.; Snieckus, V. Palladium-catalyzed cross-coupling: A historical contextual perspective to the 2010 Nobel Prize. Angew. Chem. Int. Ed. 2012, 51, 5062–5085. [Google Scholar] [CrossRef] [PubMed]
  23. Fihri, A.; Bouhrara, M.; Nekoueishahraki, B.; Polshettiwar, V. Nanocatalysts for Suzuki cross-coupling reactions. Chem. Soc. Rev. 2011, 40, 5181–5203. [Google Scholar] [CrossRef]
  24. Kanuru, V.K.; Kyriakou, G.; Beaumont, S.K.; Papageorgiou, A.C.; Watson, D.J.; Lambert, R.M. Sonogashira coupling on an extended gold surface in vacuo: Reaction of phenylacetylene with iodobenzene on Au(111). J. Am. Chem. Soc. 2010, 132, 8081–8086. [Google Scholar] [CrossRef]
  25. González-Arellano, C.; Abad, A.; Corma, A.; García, H.; Iglesias, M.; Sánchez, F. Catalysis by gold(I) and gold(III): A parallelism between homo- and heterogeneous catalysts for copper-free Sonogashira cross-coupling reactions. Angew. Chem. Int. Ed. 2007, 46, 1536–1538. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, J.Z.; Abroshan, H.; Liu, C.; Zhu, M.Z.; Li, G.; Haruta, M. Sonogashira cross-coupling on the Au(111) and Au(100) facets of gold nanorod catalysts: Experimental and computational investigation. J. Catal. 2015, 330, 354–361. [Google Scholar] [CrossRef]
  27. Li, G.; Zeng, C.J.; Jin, R.C. Chemoselective hydrogenation of nitrobenzaldehyde to nitrobenzyl alcohol with unsupported au nanorod catalysts in water. J. Phys. Chem. C 2015, 119, 11143–11147. [Google Scholar] [CrossRef]
  28. Murphy, C.J.; Gole, A.M.; Hunyadi, S.E.; Stone, J.W.; Sisco, P.N.; Alkilany, A.; Kinard, B.E.; Hankins, P. Chemical sensing and imaging with metallic nanorods. Chem. Commun. 2008, 5, 544–547. [Google Scholar] [CrossRef]
  29. Bai, X.T.; Gao, Y.N.; Liu, H.G.; Zheng, L.Q. Synthesis of amphiphilic ionic liquids terminated gold nanorods and their superior catalytic activity for the reduction of nitro compounds. J. Phys. Chem. C 2009, 113, 17730–17736. [Google Scholar] [CrossRef]
  30. Boronat, M.; Combita, D.; Concepción, P.; Corma, A.; García, H.; Juárez, R.; Laursen, S.; López-Castro, J.D. Making C-C bonds with gold: Identification of selective gold sites for homo- and cross-coupling reactions between iodobenzene and alkynes. J. Phys. Chem. C 2012, 116, 24855–24867. [Google Scholar] [CrossRef]
  31. Boronat, M.; Corma, A. Molecular approaches to catalysis naked gold NPs as quasi-molecular catalysts for green processes. J. Catal. 2011, 284, 138–147. [Google Scholar] [CrossRef]
  32. Corma, A.; Juárez, R.; Boronat, M.; Sánchez, F.; Iglesiasc, M.; García, H. Gold catalyzes the Sonogashira coupling reaction without the requirement of palladium impurities. Chem. Commun. 2011, 47, 1446–1448. [Google Scholar] [CrossRef] [PubMed]
  33. Zhang, J.; Zhou, Y.; Zheng, K.; Abroshan, H.; Kauffman, D.R.; Sun, J.; Li, G. Diphosphine-induced chiral propeller arrangement of gold nanoclusters for singlet oxygen photogeneration. Nano Res. 2018, 11, 5787–5798. [Google Scholar] [CrossRef]
  34. Zhu, M.; Aikens, C.M.; Hollander, F.J.; Schatz, G.C.; Jin, R.C. Correlating the crystal structure of a thiol-protected Au25 cluster and optical properties. J. Am. Chem. Soc. 2008, 130, 5883–5885. [Google Scholar] [CrossRef]
  35. Zheng, K.; Zhang, J.W.; Zhao, D.; Yang, Y.; Li, Z.M.; Li, G. Motif Mediated Au25(SPh)5(PPh3)10X2 Nanorod of Conjugated Electron Delocalization. Nano Res. 2019, 12, 501–507. [Google Scholar] [CrossRef]
  36. Li, G.; Jin, R.C. Gold nanocluster-catalyzed semihydrogenation: A unique activation pathway for terminal alkynes. J. Am. Chem. Soc. 2014, 136, 11347–11354. [Google Scholar] [CrossRef]
  37. Li, Z.M.; Liu, C.; Abroshan, H.; Kauffman, D.R.; Li, G. Au38S2(SAdm)20 Photocatalyst for One-Step Selective Aerobic Oxidations. ACS Catal. 2017, 7, 3368–3374. [Google Scholar] [CrossRef]
  38. Yan, C.Y.; Liu, C.; Abroshan, H.; Li, Z.M.; Qiu, R.; Li, G. Surface modification of adamantane-terminated gold nanoclusters using cyclodextrins. Phys. Chem. Chem. Phys. 2016, 18, 23358–23364. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, C.; Chen, Y.; Wang, H.; Li, Z.; Zheng, K.; Li, S.; Li, G. Transition-metal-mediated catalytic properties of CeO2-supported gold clusters in aerobic alcohol oxidation. Nano Res. 2018, 11, 2139–2148. [Google Scholar] [CrossRef]
  40. Kumar, S.S.; Kwak, K.; Lee, D. Amperometric sensing based on glutathione protected Au25 nanoparticles and their pH dependent electrocatalytic activity. Electroanalysis 2011, 23, 2116–2124. [Google Scholar] [CrossRef]
  41. Kauffman, D.R.; Alfonso, D.; Matranga, C.; Qian, H.; Jin, R. Experimental and computational investigation of Au25 clusters and CO2: A unique interaction and enhanced electrocatalytic activity. J. Am. Chem. Soc. 2012, 134, 10237–10243. [Google Scholar] [CrossRef] [PubMed]
  42. Monnier, F.; Taillefer, M. Catalytic C-C, C-N, and C-O Ullmann-type coupling reactions. Angew. Chem. Int. Ed. 2009, 48, 6954–6971. [Google Scholar] [CrossRef] [PubMed]
  43. Yu, C.L.; Li, G.; Kumar, K.S.; Kawasaki, H.; Jin, R.C. Stable Au25(SR)18/TiO2 Composite nanostructure with enhanced visible light photocatalytic activity. J. Phys. Chem. Lett. 2013, 4, 2847–2852. [Google Scholar] [CrossRef]
  44. Chen, H.J.; Liu, C.; Wang, M.; Zhang, C.F.; Li, G.; Wang, F. Thermally robust silica-enclosed Au25 nanocluster and its catalysis. Chin. J. Catal. 2016, 37, 1787–1793. [Google Scholar] [CrossRef]
  45. Li, G.; Liu, C.; Lei, Y.; Jin, R.C. Au25 nanocluster-catalyzed Ullmann-type homocoupling reaction of aryl iodides. Chem. Commun. 2012, 48, 12005–12007. [Google Scholar] [CrossRef] [PubMed]
  46. Li, G.; Abroshan, H.; Liu, C.; Zhuo, S.; Li, Z.M.; Xie, Y.; Kim, H.J.; Rosi, N.L.; Jin, R.C. Tailoring the electronic and catalytic properties of Au25 nanoclusters via ligand engineering. ACS Nano 2016, 10, 7998–8005. [Google Scholar] [CrossRef] [PubMed]
  47. Abroshan, H.; Li, G.; Lin, J.Z.; Kim, H.J.; Jin, R.C. Molecular mechanism for the activation of Au25(SCH2CH2Ph)18 nanoclusters by imidazolium-based ionic liquids for catalysis. J. Catal. 2016, 337, 72–79. [Google Scholar] [CrossRef]
  48. Li, G.; Abroshan, H.; Chen, Y.X.; Jin, R.C.; Kim, H.J. Experimental and mechanistic understanding of aldehyde hydrogenation using Au25 nanoclusters with Lewis acids: Unique sites for catalytic reactions. J. Am. Chem. Soc. 2015, 137, 14295–14304. [Google Scholar] [CrossRef]
  49. Li, G.; Jiang, D.E.; Liu, C.; Yu, C.L.; Jin, R.C. Oxide-supported atomically precise gold nanocluster for catalyzing Sonogashira cross-coupling. J. Catal. 2013, 306, 177–183. [Google Scholar] [CrossRef]
  50. Kumara, C.; Aikens, C.M.; Dass, A. X-ray crystal structure and theoretical analysis of Au25−xAgx(SCH2CH2Ph)18- alloy. J. Phys. Chem. Lett. 2014, 5, 461–466. [Google Scholar] [CrossRef]
  51. Li, W.L.; Liu, C.; Abroshan, H.; Ge, Q.J.; Yang, X.J.; Xu, H.Y.; Li, G. Catalytic CO oxidation using bimetallic MxAu25-x clusters: A combined experimental and computational study on doping effects. J. Phys. Chem. C 2016, 120, 10261–10267. [Google Scholar] [CrossRef]
  52. Jiang, D.E.; Whetten, R.L. Magnetic doping of a thiolated-gold superatom: First-principles density functional theory calculations. Phys. Rev. B 2009, 80, 115402. [Google Scholar] [CrossRef] [Green Version]
  53. Qian, H.F.; Jiang, D.E.; Li, G.; Gayathri, C.; Das, A.; Gil, R.R.; Jin, R.C. Monoplatinum Doping of Gold Nanoclusters and Catalytic Application. J. Am. Chem. Soc. 2012, 134, 16159–16162. [Google Scholar] [CrossRef]
  54. Xie, S.H.; Tsunoyama, H.; Kurashige, W.; Negishi, Y.; Tsukuda, T. Enhancement in aerobic alcohol oxidation catalysis of Au25 clusters by single Pd atom doping. ACS Catal. 2012, 2, 1519–1523. [Google Scholar] [CrossRef]
  55. Li, G.; Jin, R.C. Atomic level tuning of the catalytic properties: Doping effects of 25-atom bimetallic nanoclusters on styrene oxidation. Catal. Today 2016, 278, 187–191. [Google Scholar] [CrossRef] [Green Version]
  56. Li, Z.M.; Yang, X.J.; Liu, C.; Wang, J.; Li, G. Effects of doping in 25-atom bimetallic nanocluster catalysts for carbon-carbon coupling reaction of iodoanisole and phenylacetylene. Proc. Nat. Sci. Mater. Int. 2016, 26, 477–482. [Google Scholar] [CrossRef]
  57. Qin, Z.; Zhao, D.; Zhao, L.; Xiao, Q.; Wu, T.; Zhang, J.; Wan, C.-Q.; Li, G. Tailoring the Stability, Photocatalysis and Photoluminescence Properties of Au11 Nanocluster via Doping Engineering. Nanoscale Adv. 2019, 2. [Google Scholar] [CrossRef]
  58. Dömling, A. Recent developments in isocyanide based multicomponent reactions in applied chemistry. Chem. Rev. 2006, 106, 17–89. [Google Scholar] [CrossRef]
  59. Ganem, B. Strategies for innovation in multicomponent reaction design. Acc. Chem. Res. 2009, 42, 463–472. [Google Scholar] [CrossRef]
  60. Climent, M.J.; Corma, A.; Iborra, S. Homogeneous and heterogeneous catalysts for multicomponent reactions. RSC Adv. 2012, 2, 16–58. [Google Scholar] [CrossRef]
  61. Wei, C.M.; Li, C.J. A highly efficient three-component coupling of aldehyde, alkyne, and amines via C-H activation catalyzed by gold in water. J. Am. Chem. Soc. 2003, 125, 9584–9585. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, Y.D.; Liu, C.; Abroshan, H.; Li, Z.M.; Wang, J.; Li, G.; Haruta, M. Phosphine/phenylacetylide-ligated Au clusters for multicomponent coupling reactions. J. Catal. 2016, 337, 287–294. [Google Scholar] [CrossRef]
  63. Liu, C.; Abroshan, H.; Yan, C.Y.; Li, G.; Haruta, M. One-pot synthesis of Au11(PPh2Py)7Br3 for the highly chemoselective hydrogenation of nitrobenzaldehyde. ACS Catal. 2016, 6, 92–99. [Google Scholar] [CrossRef]
  64. Li, Q.; Das, A.; Wang, S.X.; Chen, Y.X.; Jin, R.C. Highly efficient three-component coupling reaction catalysed by atomically precise ligand-protected Au38(SC2H4Ph)24 nanoclusters. Chem. Commun. 2016, 52, 14298–14301. [Google Scholar] [CrossRef]
  65. Li, M.; Tian, S.; Wu, Z. Improving the catalytic activity of Au25 nanocluster by peeling and doping. Chin. J. Chem. 2017, 35, 567–571. [Google Scholar] [CrossRef]
  66. Wen, Z.Y.; Li, Z.M.; Ge, Q.J.; Zhou, Y.; Sun, J.; Li, G. Robust nickel cluster@Mes-HZSM-5 composite nanostructure with enhanced catalytic activity in the DTG reaction. J. Catal. 2018, 363, 26–33. [Google Scholar] [CrossRef]
  67. Li, Z.M.; Li, W.L.; Abroshan, H.; Ge, Q.J.; Zhou, Y.; Zhang, C.L.; Li, G.; Jin, R.C. Dual Effects of Water Vapor over Ceria-Supported Gold Clusters. Nanoscale 2018, 10, 6558–6565. [Google Scholar] [CrossRef]
  68. Wang, F.; Wen, Z.; Fang, Q.; Ge, Q.; Sun, J.; Li, G. Manganese cluster induce the control synthesis of RHO- and CHA-type silicoaluminaphosphates for dimethylether to light olefin conversion. Fuel 2019, 244, 104–109. [Google Scholar] [CrossRef]
  69. Guo, S.; Zhang, S.H.; Fang, Q.H.; Abroshan, H.; Kim, H.; Haruta, M.; Li, G. Gold-palladium nanoalloys supported by graphene oxide and lamellar TiO2 for direct synthesis of hydrogen peroxide. ACS Appl. Mater. Interfaces 2018, 10, 40599–40607. [Google Scholar] [CrossRef]
  70. Liu, C.; Zhang, J.Y.; Huang, J.H.; Zhang, C.L.; Hong, F.; Zhou, Y.; Li, G.; Haruta, M. Efficient aerobic oxidation of glucose to gluconic acid over activated carbon-supported gold clusters. ChemSuSChem 2017, 10, 1976–1980. [Google Scholar] [CrossRef]
  71. Vilhelmsen, L.B.; Walton, K.S.; Sholl, D.S. Structure and mobility of metal clusters in MOFs: Au, Pd, and AuPd clusters in MOF-74. J. Am. Chem. Soc. 2012, 134, 12807–12816. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, H.; Li, Z.; Qin, Z.; Kim, H.J.; Abroshan, H.; Li, G. Silica-encapsulated gold nanoclusters for efficient acetylene hydrogenation to ethylene. ACS Appl. Nano Mater. 2019. [Google Scholar] [CrossRef]
  73. Ciobanu, M.; Cojocaru, B.; Teodorescu, C.; Vasiliu, F.; Coman, S.M.; Leitner, W.; Parvulescu, V.I. Heterogeneous amination of bromobenzene over titania-supported gold catalysts. J. Catal. 2012, 296, 43–54. [Google Scholar] [CrossRef]
  74. Fang, Q.; Qin, Z.; Shi, Y.; Liu, F.; Barkaoui, S.; Abroshan, H.; Li, G. Au/NiO composite: A catalyst for one-pot cascade conversion of furfural. ACS Appl. Energy Mater. 2019, 2, 2654–2661. [Google Scholar] [CrossRef]
Scheme 1. Cross-coupling reactions over gold cluster in this review.
Scheme 1. Cross-coupling reactions over gold cluster in this review.
Nanomaterials 09 00838 sch001
Figure 1. Proposed pathway for the cross-coupling on the (A) Au(100) and (B) Au(111). Reproduced with permission from [26]. Elsevier, 2015.
Figure 1. Proposed pathway for the cross-coupling on the (A) Au(100) and (B) Au(111). Reproduced with permission from [26]. Elsevier, 2015.
Nanomaterials 09 00838 g001
Figure 2. Calculated energy of the dissociation of PhB(OH)3 anion (A) and deprotonation of PA (B) on the Au38O2 cluster. Reproduced with permission from [30]. American Chemical Society, 2012.
Figure 2. Calculated energy of the dissociation of PhB(OH)3 anion (A) and deprotonation of PA (B) on the Au38O2 cluster. Reproduced with permission from [30]. American Chemical Society, 2012.
Nanomaterials 09 00838 g002
Figure 3. Framework of Au25(SR)18 (A) and Au25(PPh3)10L5Cl2 (B). The orange areas in are the Au3 and the waist active sites during the catalysis. Reproduced with permission from [36]. American Chemical Society, 2008 and Springer, 2019.
Figure 3. Framework of Au25(SR)18 (A) and Au25(PPh3)10L5Cl2 (B). The orange areas in are the Au3 and the waist active sites during the catalysis. Reproduced with permission from [36]. American Chemical Society, 2008 and Springer, 2019.
Nanomaterials 09 00838 g003
Figure 4. (A) Cyclic voltammograms (CV) for the redox reaction in potassium ferricyanide solution on Au38S2(SAdm)20 clusters and Au38S2(SAdm)20-(β-CD)2 conjugates. (B) Schematic illustrations of the Au38S2(SAdm)20-(β-CD)2 formation and the charge transfer between Fe(CN)63− and the Au clusters and conjugates. Reproduced with permission from [38]. The Royal Society of Chemistry, 2016.
Figure 4. (A) Cyclic voltammograms (CV) for the redox reaction in potassium ferricyanide solution on Au38S2(SAdm)20 clusters and Au38S2(SAdm)20-(β-CD)2 conjugates. (B) Schematic illustrations of the Au38S2(SAdm)20-(β-CD)2 formation and the charge transfer between Fe(CN)63− and the Au clusters and conjugates. Reproduced with permission from [38]. The Royal Society of Chemistry, 2016.
Nanomaterials 09 00838 g004
Figure 5. Interaction of 4-NO2C6H4I with the exposed Au atoms on the Au25(SCH3)2(SH)15 (A) and Au25(SNap)2(SH)15 (B). Energy vs. reaction coordinate with “-SCH3” (C) and “-SNap” ligands (D) during the homo- and hetero-coupling reactions. Reproduced with permission from [46]. American Chemical Society, 2016.
Figure 5. Interaction of 4-NO2C6H4I with the exposed Au atoms on the Au25(SCH3)2(SH)15 (A) and Au25(SNap)2(SH)15 (B). Energy vs. reaction coordinate with “-SCH3” (C) and “-SNap” ligands (D) during the homo- and hetero-coupling reactions. Reproduced with permission from [46]. American Chemical Society, 2016.
Nanomaterials 09 00838 g005
Figure 6. Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) of the fresh Au25(PET)18 clusters and samples treated with the ionic liquid (IL) of BMIM·BF4; the (n, m) presents the number of the Aun(PET)m species. Reproduced with permission from [47]. American Chemical Society, 2015.
Figure 6. Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) of the fresh Au25(PET)18 clusters and samples treated with the ionic liquid (IL) of BMIM·BF4; the (n, m) presents the number of the Aun(PET)m species. Reproduced with permission from [47]. American Chemical Society, 2015.
Nanomaterials 09 00838 g006
Figure 7. (A) Top view of one of the two open facets of Au25(PET)18 clusters where three external Au atoms are exposed; the other facet is on the back side. (B) Side view of the two facets which are at top and bottom. (C) Top view and (D) side view of the co-adsorption of PA and IB on the surface of the Au25(PET)18 clusters. Color code: Au, yellow; S, blue; C, gray; H, white; I, green. Reproduced with permission from [49]. Elsevier, 2013.
Figure 7. (A) Top view of one of the two open facets of Au25(PET)18 clusters where three external Au atoms are exposed; the other facet is on the back side. (B) Side view of the two facets which are at top and bottom. (C) Top view and (D) side view of the co-adsorption of PA and IB on the surface of the Au25(PET)18 clusters. Color code: Au, yellow; S, blue; C, gray; H, white; I, green. Reproduced with permission from [49]. Elsevier, 2013.
Nanomaterials 09 00838 g007
Figure 8. Catalytic results over the Au25(PPh3)10(PA)5X2 ([Au25]/TiO2) catalysts in the A3-coupling reaction. Reaction conditions: 100 mg catalyst (1 wt.% cluster loading), 1.0 mmol benzaldehyde, 1.2 mmol piperidine, 1.3 mmol phenylacetylene, 5 mL water, 100 °C, 18 h, under a N2 atmosphere.
Figure 8. Catalytic results over the Au25(PPh3)10(PA)5X2 ([Au25]/TiO2) catalysts in the A3-coupling reaction. Reaction conditions: 100 mg catalyst (1 wt.% cluster loading), 1.0 mmol benzaldehyde, 1.2 mmol piperidine, 1.3 mmol phenylacetylene, 5 mL water, 100 °C, 18 h, under a N2 atmosphere.
Nanomaterials 09 00838 g008
Figure 9. Proposed mechanism for the [Au25(PH3)10(PA)5Cl2]2+-catalyzed A3-coupling reaction. Reproduced with permission from [62]. Elsevier, 2016.
Figure 9. Proposed mechanism for the [Au25(PH3)10(PA)5Cl2]2+-catalyzed A3-coupling reaction. Reproduced with permission from [62]. Elsevier, 2016.
Nanomaterials 09 00838 g009
Table 1. The catalytic results over the Au25(PET)18/CeO2 catalysts in Ullmann homo-coupling. Reaction conditions: 0.2 mmol iodobenzene, 0.6 mmol K2CO3, 100 mg Au25(PET)18/CeO2 (1 wt% cluster loading), 1 mL DMF, 130 °C, 2 day.
Table 1. The catalytic results over the Au25(PET)18/CeO2 catalysts in Ullmann homo-coupling. Reaction conditions: 0.2 mmol iodobenzene, 0.6 mmol K2CO3, 100 mg Au25(PET)18/CeO2 (1 wt% cluster loading), 1 mL DMF, 130 °C, 2 day.
Nanomaterials 09 00838 i011
EntrySubstrateProductConversion (%)
1 Nanomaterials 09 00838 i001 Nanomaterials 09 00838 i00299.8
2 Nanomaterials 09 00838 i003 Nanomaterials 09 00838 i00499.5
3 Nanomaterials 09 00838 i005 Nanomaterials 09 00838 i00667.5
4 Nanomaterials 09 00838 i007 Nanomaterials 09 00838 i00878.2
5 Nanomaterials 09 00838 i009 Nanomaterials 09 00838 i01099.7
Table 2. The catalytic results over the Au25(SR)18/CeO2 catalysts in Ullmann hetero-coupling. Reaction conditions: 0.06 mmoL 4-methyl iodobenzene, 0.05 mmoL 4-nitroiodobenzene, 0.3 mmol K2CO3, 100 mg Au25(SR)18/CeO2 (~1 wt % cluster loading), 1 mL DMF, 130 °C, 24 h.
Table 2. The catalytic results over the Au25(SR)18/CeO2 catalysts in Ullmann hetero-coupling. Reaction conditions: 0.06 mmoL 4-methyl iodobenzene, 0.05 mmoL 4-nitroiodobenzene, 0.3 mmol K2CO3, 100 mg Au25(SR)18/CeO2 (~1 wt % cluster loading), 1 mL DMF, 130 °C, 24 h.
Nanomaterials 09 00838 i012
CatalystConversion (%)Selectivity (%)
Au25(SC6H13)186916
Au25(PET)187219
Au25(SPh)188050
Au25(SNap)189182
Table 3. Catalytic results over the Au25/TiO2 catalysts in the Suzuki cross-coupling reactions of iodoanisole and phenylboronic acid using the imidazolium-based ionic liquids. Reaction conditions: 0.1 mmol iodoanisole, 0.12 mmol phenylboronic acid, 0.36 mmol K2CO3, 100 mg Au25(SR)18/TiO2 (~1 wt % cluster loading), solvents (1 mL EtOH or EtOH:H2O (10:1, v/v), and 0.2 mL IL or other organic solvents), 90 °C, 18 h.
Table 3. Catalytic results over the Au25/TiO2 catalysts in the Suzuki cross-coupling reactions of iodoanisole and phenylboronic acid using the imidazolium-based ionic liquids. Reaction conditions: 0.1 mmol iodoanisole, 0.12 mmol phenylboronic acid, 0.36 mmol K2CO3, 100 mg Au25(SR)18/TiO2 (~1 wt % cluster loading), solvents (1 mL EtOH or EtOH:H2O (10:1, v/v), and 0.2 mL IL or other organic solvents), 90 °C, 18 h.
Nanomaterials 09 00838 i013
SolventConversion (%)
o-xylene + EtOH:H2O<0.5
DMF + EtOH:H2O4
BMIM·Br + EtOH89
BMIM·Cl + EtOH90
BMIM·BF4 + EtOH94
BMIM·BF4 + EtOH:H2O EtOH:H2O>99
BDiMIM·BF4 + EtOH<0.5
Table 4. The catalytic performance of Au25(SR)18/oxides in the Sonogashira cross-coupling reaction of p-iodoanisole and phenylacetylene. DMBP and MPEB stand for homo-coupling product of 4,4′-dimethoxy-1,1′-biphenyl and cross-coupling product of 1-methoxy-4-(2-phenylethynyl)benzene, respectively. Reaction conditions: 100 mg catalyst (1 wt % Au25(SR)18 or AuNC 2–3 nm (SC6H13) loading), 0.1 mmol p-iodoanisole, 0.15 mmol phenylacetylene, 0.3 mmol K2CO3, 1 mL DMF, 160 °C, 40 h. n.r. = no reaction. Conv. = conversion.
Table 4. The catalytic performance of Au25(SR)18/oxides in the Sonogashira cross-coupling reaction of p-iodoanisole and phenylacetylene. DMBP and MPEB stand for homo-coupling product of 4,4′-dimethoxy-1,1′-biphenyl and cross-coupling product of 1-methoxy-4-(2-phenylethynyl)benzene, respectively. Reaction conditions: 100 mg catalyst (1 wt % Au25(SR)18 or AuNC 2–3 nm (SC6H13) loading), 0.1 mmol p-iodoanisole, 0.15 mmol phenylacetylene, 0.3 mmol K2CO3, 1 mL DMF, 160 °C, 40 h. n.r. = no reaction. Conv. = conversion.
EntryCatalystConv. (%)Selectivity (%)
MPEBDMBP
1Au25(SR)18/CeO296.188.111.9
2Au25(SR)18/TiO292.882.917.1
3Au25(SR)18/SiO290.879.320.7
4Au25(SR)18/MgO93.380.619.4
5AuNC 2–3 nm (SC6H13)/CeO265.557.242.8
6CeO2n.r.
7TiO2n.r.
Table 5. Catalytic performance of TiO2-supported MxAu25−x(SR)18 catalysts in the C–C coupling reaction between MeOC6H4I and PA. Reaction conditions: 100 mg catalyst, 1 wt % MxAu25−x(SR)18 loading, 0.1 mmol MeOC6H4I, 0.1 mmol PA, 0.3 mmol K2CO3, 1 mL DMF, 160 °C, 40 hr.
Table 5. Catalytic performance of TiO2-supported MxAu25−x(SR)18 catalysts in the C–C coupling reaction between MeOC6H4I and PA. Reaction conditions: 100 mg catalyst, 1 wt % MxAu25−x(SR)18 loading, 0.1 mmol MeOC6H4I, 0.1 mmol PA, 0.3 mmol K2CO3, 1 mL DMF, 160 °C, 40 hr.
Nanomaterials 09 00838 i014
EntryCatalystsConversion (%)Selectivity (%)
MPEBDMBP
1Au25(SR)1879.565.734.3
2AgxAu25−x(SR)1883.055.444.6
3CuxAu25−x(SR)1852.428.371.7
4Pt1Au24(SR)1848.567.232.9
Table 6. Adsorption energy of the PA and IB on the Au3 site of the M1Au24(SR)18 (M: Pt, Ag, and Cu, and noted as M1Au24) and M2Au23(SR)18 (M: Ag and Cu, M2Au23) cluster models, respectively.
Table 6. Adsorption energy of the PA and IB on the Au3 site of the M1Au24(SR)18 (M: Pt, Ag, and Cu, and noted as M1Au24) and M2Au23(SR)18 (M: Ag and Cu, M2Au23) cluster models, respectively.
Adsorption Energy (eV)MxAu25−x(SR)18 Cluster
Au25Pt1Au24Ag1Au24Ag2Au23Cu1Au24Cu2Au23
One PA−0.51−0.50−0.61−0.60−0.52−0.52
One IB−0.60−0.61−0.59−0.59−0.54−0.58
“PA + IB” pair−1.11−1.11−1.20−1.19−1.06−1.10
“IB + IB” pair−1.20−1.22−1.18−1.18−1.08−1.16

Share and Cite

MDPI and ACS Style

Shi, Q.; Qin, Z.; Xu, H.; Li, G. Heterogeneous Cross-Coupling over Gold Nanoclusters. Nanomaterials 2019, 9, 838. https://doi.org/10.3390/nano9060838

AMA Style

Shi Q, Qin Z, Xu H, Li G. Heterogeneous Cross-Coupling over Gold Nanoclusters. Nanomaterials. 2019; 9(6):838. https://doi.org/10.3390/nano9060838

Chicago/Turabian Style

Shi, Quanquan, Zhaoxian Qin, Hui Xu, and Gao Li. 2019. "Heterogeneous Cross-Coupling over Gold Nanoclusters" Nanomaterials 9, no. 6: 838. https://doi.org/10.3390/nano9060838

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop