Next Article in Journal
Tailorable Synthesis of Highly Oxidized Graphene Oxides via an Environmentally-Friendly Electrochemical Process
Next Article in Special Issue
Origin of the Photoluminescence of Metal Nanoclusters: From Metal-Centered Emission to Ligand-Centered Emission
Previous Article in Journal
Ultra-Uniform and Very Thin Ag Nanowires Synthesized via the Synergy of Cl, Br and Fe3+ for Transparent Conductive Films
Previous Article in Special Issue
A New Lamellar Gold Thiolate Coordination Polymer, [Au(m-SPhCO2H)]n, for the Formation of Luminescent Polymer Composites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gold Nanoclusters as Electrocatalysts for Energy Conversion

by
Tokuhisa Kawawaki
1,2 and
Yuichi Negishi
1,2,*
1
Department of Applied Chemistry, Faculty of Science, Tokyo University of Science, 1–3 Kagurazaka, Shinjuku-ku, Tokyo 162–8601, Japan
2
Photocatalysis International Research Center, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba 278−8510, Japan
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(2), 238; https://doi.org/10.3390/nano10020238
Submission received: 11 January 2020 / Revised: 23 January 2020 / Accepted: 27 January 2020 / Published: 29 January 2020

Abstract

:
Gold nanoclusters (Aun NCs) exhibit a size-specific electronic structure unlike bulk gold and can therefore be used as catalysts in various reactions. Ligand-protected Aun NCs can be synthesized with atomic precision, and the geometric structures of many Aun NCs have been determined by single-crystal X-ray diffraction analysis. In addition, Aun NCs can be doped with various types of elements. Clarification of the effects of changes to the chemical composition, geometric structure, and associated electronic state on catalytic activity would enable a deep understanding of the active sites and mechanisms in catalytic reactions as well as key factors for high activation. Furthermore, it may be possible to synthesize Aun NCs with properties that surpass those of conventional catalysts using the obtained design guidelines. With these expectations, catalyst research using Aun NCs as a model catalyst has been actively conducted in recent years. This review focuses on the application of Aun NCs as an electrocatalyst and outlines recent research progress.

1. Introduction

Gold nanoclusters (Aun NCs) have physical/chemical properties that differ from those of bulk Au owing to their size-specific electrical/geometrical structure [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22]. Therefore, Aun NCs have been actively studied since the 1960s from the viewpoints of both basic science and application. Since Brust et al., discovered a method for synthesizing Aun NCs protected by thiolate (Aun(SR)m) in 1994 [1], researches on Aun NCs in particular have grown [6]. Aun(SR)m NCs exhibit high stability both in solution and in the solid state because Au forms a strong bond with SR. In addition, Aun(SR)m NCs can be synthesized by simply mixing reagents under the ambient atmosphere. Aun(SR)m NCs with these unique characteristics have a low handling threshold even for researchers unfamiliar with the chemical synthesis of metal clusters. Aun(SR)m NCs are thus currently one of the most studied metal NCs [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18]. For these Aun(SR)m NCs, it became possible to synthesize a series of Aun(SR)m NCs with atomic precision in 2005 [19]. In addition, since 2007, the geometric structures of many Aun(SR)m NCs have been determined through single-crystal X-ray diffraction (SC-XRD) analysis [20]. Since 2009, partial replacement of the Au atoms of Aun(SR)m NCs with other elements such as silver (Ag), copper (Cu), platinum (Pt), palladium (Pd), cadmium (Cd), and mercury (Hg) has also been realized [3,4,5,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44].
In parallel to these synthesis and structural analysis studies, studies on the functions of Aun NCs have also been actively conducted. Aun NCs have been observed to possess catalytic activity for several reactions, including carbon monoxide oxidation [45,46,47,48,49,50,51,52,53,54,55], alcohol oxidation [56,57,58,59,60,61,62,63,64,65], styrene oxidation [66,67,68,69,70], aromatic compound oxidation [71,72], sulfide oxidation [73,74,75], and carbon dioxide reduction [76,77,78,79,80,81,82,83]. One of the reasons for these active studies on the catalysis of Aun NCs is that their electronic and geometric structures are well understood. Thus, if the obtained catalytic properties are compared with the electronic/geometrical structures of Aun(SR)m NCs, information on active sites, mechanisms, and key factors for high activation in catalytic reactions can be obtained. With these expectations, Aun(SR)m NCs have received great attention as model catalysts [45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83].
In addition, several studies on Aun(SR)m NCs as electrocatalysts have also been performed recently. To prevent serious environmental issues including the depletion of fossil fuels and global warming, the establishment of a system in which hydrogen (H2) is generated from water and solar energy using a photocatalyst is desired, with the generated H2 used for the generation of electricity using fuel cells [84,85]. Once such an energy conversion system is established, it will be possible to circulate an energy medium (H2) in addition to obtaining electricity only from solar energy and abundant water resources. However, realization of such an ultimate energy conversion system requires further improvement of the reaction efficiency of each half reaction of water splitting and fuel cells, including the hydrogen evolution reaction (HER), oxygen evolution reaction (OER), hydrogen oxidation reaction (HOR), and oxygen reduction reaction (ORR; Figure 1A).
To improve the reactivity per unit volume, it is necessary to increase the specific surface area of the active sites and increase the reaction rate at the active sites. For the former, size reduction of the catalyst is one effective method. However, the latter is strongly related to the adsorption energy of reactive molecules on the catalyst surface. The activity of the chemical reaction on the catalyst surface is the highest when the Gibbs energy of adsorption between the catalyst and reactant is moderate according to the Sabatier principle [86]. This is because the reaction does not occur without the adsorption of reactants but is inhibited by the strong adsorption of reactants. Therefore, the relationship between the reaction efficiency and the Gibbs energy for the adsorption of reactants follows a curved line called an activity volcano plot [87]. Fine nanoparticle catalysts suitable for the HER [88,89,90,91,92], OER [93,94,95], and ORR [96,97,98,99,100,101] have been developed based on theoretical predictions of activity volcano plots using various metals and alloy nanoparticles (NPs). Aun NCs have recently been observed to possess catalytic activity for the HER, OER, and ORR [77,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116] (Figure 1). Therefore, Aun NCs are expected to become a model catalyst even in such an energy conversion system. A better understanding of the correlation between electronic/geometrical structures and the catalytic activity of the HER, OER, and ORR in Aun NCs might lead to the discovery of new key factors for achieving high activation. Furthermore, because Aun NCs are composed of several tens of atoms or less, the use of fine Aun NCs as a catalyst is also effective in reducing the consumption of expensive noble metals. Thus, it may be possible to create HER, OER, and ORR catalysts with properties that surpass those of conventional catalysts using these unique characteristics of Aun NCs. With these expectations, several groups are conducting research on the application of Aun NCs as electrocatalysts. This article reviews the basic theory of electrocatalysts and recent research on HER, OER, and ORR catalysts using Aun NCs and their alloy NCs.

2. Electrocatalytic Reaction in Water Splitting

H2 is expected to be an important energy source to support a sustainable energy society. Currently, H2 is generated as a by-product during steam reforming or coke production. However, if a water-splitting reaction using an electrocatalyst can be applied for hydrogen production, the large-scale facility of the current system would not be required. In addition, it would be possible to produce H2 only with water and electricity using the surplus power from a power plant. Therefore, water electrolysis is considered one of the cleanest energy production reactions for a sustainable energy society.
The water-splitting reaction consists of two half reactions, the HER and OER. When a voltage is applied to the metal electrode, a reduction reaction proceeds at the cathode and an oxidation reaction proceeds at the anode, resulting in the decomposition of water molecules into H2 and O2 at each electrode. However, the reactions do not proceed even if a potential equal to or higher than both the oxidation and reduction potentials in each reaction (HER: 0 V vs. SHE, OER: 1.23 V vs. SHE; SHE = standard hydrogen electrode) is applied to the electrode. This is because the activation energy of each reaction is too high. Therefore, noble metal NPs are used as a catalyst to reduce the activation energy of the reaction.

2.1. Hydrogen Evolution Reaction

In the HER, metal surface atoms of the catalyst form bonding orbitals with protons (H+) through the Volmer–Heyrovsky or Volmer–Tafel mechanism, producing molecular hydrogen [117].
Under acidic conditions, the following reactions occur:
Volmer reaction: M + H+ + e → M–H
Heyrovsky reaction: M–H + H+ + e → M–H2
Tafel reaction: 2M–H → 2M + H2
However, under alkaline conditions, the following reactions occur:
Volmer reaction: 2M + 2H2O + 2e → 2M–H + 2OH
Heyrovsky reaction: M–H + H2O + e → M–H2 + OH
Tafel reaction: 2M–H → 2M + H2
Bulk Au possesses almost no HER activity, whereas Aun(SR)m NCs possess HER activity. In addition, their activity can be further improved by doping Aun(SR)m NCs with appropriate heterogeneous elements. These effects were reported by Lee and Jiang et al., in 2017 [102]. They evaluated the HER activity using linear sweep voltammetry (LSV) in tetrahydrofuran (THF) solution with 1.0 M trifluoroacetic acid (TFA) and 0.1 M tetrabutylammonium hexafluorophosphate (Bu4NPF6) in the absence (black) and presence of Au25(SC6H13)18 or Au24Pt(SC6H13)18 (SC6H13 = 1-hexanethiolate) on a glassy carbon electrode (GCE). The onset potential of the HER (Figure 1B(a)) occurred at −1.25 V for the GCE blank (Figure 2A, black line), whereas it occurred at −1.1 V for the GCE with Au25(SC6H13)18 (Figure 2A, red line). In addition, for the GCE with Au24Pt(SC6H13)18, the onset potential of the HER was further reduced to −0.89 V (Figure 2A, blue line). These findings indicated that Aun(SR)m NCs has catalytic activity for the HER and that the HER activity can be further improved by substituting one Au atom of the Aun(SR)m NCs with a Pt atom (Table 1). They estimated the HER energies of Au25(SCH3)18 and Au24Pt(SCH3)18 (SCH3 = methanethiolate) using density functional theory (DFT) calculations to elucidate the reasons for this behavior (Figure 2C). In these DFT calculations, H+ solvated by two THF molecules was used as H+. The resulting energy change in the Volmer step was 0.539 eV for [Au25(SCH3)18], indicating that this reaction is endothermic. However, the energy change in the Volmer step was −0.059 eV for [Au24Pt(SCH3)18]2−, indicating that there is almost no energy change (Figure 2C, step 1). The higher HER activity of Au24Pt(SC6H13)18 was explained by these differences in the energy barriers in the reaction. In addition, Au24Pt(SC6H13)18 possessed higher HER activity even compared with Pt NPs, which are highly active materials for the HER (Figure 2B).
Lee and Jiang et al., observed that a high HER activity and a high catalyst turnover frequency (TOF) can be achieved by doping Au25(SC6H13)18 with not only Pt but also Pd (Au24Pt(SC6H13)18 > Au24Pd(SC6H13)18 > Au25(SC6H13)18) [103]. They reported that TOF values of Au25(SC6H13)18, Au24Pd(SC6H13)18, and Au24Pt(SC6H13)18 were 8.2, 13.0, and 33.3 mol H2 (mol catalyst)−1 s−1 at −0.60 V vs. the reversible hydrogen electrode (RHE), respectively. In addition, it was revealed that the doping of Au38(SR)24 with different elements results in a similar activity enhancement effect with Au25(SC6H13)18 (Au36Pt2(SC6H13)24 > Au36Pd2(SC6H13)24 > Au38(SC6H13)24) [103]. These results are in good agreement with the DFT calculation results. In addition to these studies, Jiang et al., also investigated the doping effects of various elements (Pt, Pd, Ag, Cu, Hg, and Cd) in Au25(SCH3)18 using DFT calculations [105]. The results predicted that Au24Pt(SCH3)18, Au24Pd(SCH3)18, and Au24Cu(SCH3)18, in which the heteroatom (Pt, Pd, or Cu) is located at the center of the metal core, have a higher HER activity than Au25(SCH3)18. Zhu et al., reported that another fine alloy NC, Au2Pd6(S4(PPh3)4(PhF2S)6) (PPh3 = triphenylphosphine, PhF2S = 3,4-difluorobenzenethiolate), also exhibits HER activity (Table 1) [106]. These studies revealed that Aun(SR)m and their alloy NCs have HER activity and it can be improved by controlling the electronic structure of Aun NCs through heteroatom doping.
The HER activity varies depending not only on the chemical composition of the metal core but also on the properties of the ligand. In 2018, Teranishi and Sakamoto et al., used Aun NCs coordinated with SR-containing porphyrin (porphyrin SCxP). They investigated the effects of the ligand structure on the HER activity of Aun(SR)m NCs [107]. In these clusters, the porphyrin ring coordinates horizontally to the gold core. Then, the distance between the porphyrin ring and the Au surface was controlled by changing the length of the alkyl chain between the porphyrin ring and the acetylthio group (Figure 3A,C) [118,119]. The alkyl chain is a methylene chain for porphyrin SC1P and an ethylene chain for porphyrin SC2P. The distance between the porphyrin ring and the acetylthio group was determined to be 3.4 Å for porphyrin SC1P and 4.9 Å for porphyrin SC2P by SC-XRD analysis. The researchers synthesized three sizes of Aun NCs with a core size of approximately 1.3, 2.2, or 3.8 nm using porphyrin SC1P, porphyrin SC2P, or a common protective ligand, 2-phenylethanethiolate (PET). Transmission electron microscope (TEM) images of the synthesized Aun(SR)m NCs (SR = porphyrin SC1P, porphyrin SC2P, or PET) with a core size of approximately 1.3 nm are presented in Figure 3B,D,F, respectively. Among these products, matrix-assisted laser desorption/ionization mass spectrometry indicated that Aun(porphyrin SC1P)m NCs consisted of 77 Au atoms and 8 porphyrin SC1P molecules and Aun(porphyrin SC2P)m NCs consisted of 75 Au atoms and 11 porphyrin SC2P molecules. The effects of the ligand structure and Au core size on the HER activity of Aun(SR)m NCs were investigated using the obtained nine types of Aun(SR)m NCs. As a result, in Aun(SR)m NCs with a core size of approximately 1.3 nm, Aun(porphyrin SC1P)m and Aun(porphyrin SC2P)m NCs exhibited higher current densities of the HER than Aun(PET)m NCs (Table 1). For instance, Aun(porphyrin SC1P)m NCs resulted in a 4.6 times higher current density of the HER than Aun(PET)m NCs at −0.4 V vs. RHE. In addition, using Aun(porphyrin SC1P)m NCs, the HER occurred at a smaller overvoltage than using Aun(porphyrin SC2P)m NCs. These results indicate that the HER activity of Aun NCs depends on the type of ligand and the distance between the ligand and the metal core in Aun NCs [107]. In this work, the Aun(SR)m NCs with a core size of approximately 2.2 nm showed higher catalytic activity than those with a core size of approximately 1.3 nm (Figure 3G,H). This size dependence of the catalytic activity is a little strange considering the surface area of the metal core because a reduction of a core size of Aun(SR)m NCs typically leads to the increase in the surface area of Au metal core, which are active sites in HER. The authors have not discussed the details on this point in this paper probably due to the difficulty in precisely estimating the surface area of each Aun(SR)m NCs.
The property of the ligand also strongly affects the interaction between Aun(SR)m NCs and the electrode as well as the affinity between Aun(SR)m NCs and water molecules. Lee and Jiang et al., synthesized Aun(SR)m NCs with SC6H13, 3-mercaptopropionic acid (MPA), or 3-mercapto-1-propanesulfonic acid (MPS; Figure 4B) as a ligand (Au25(SC6H13)18, Au25(MPA)18, and Au25(MPS)18) and used them to investigate the effect of ligand properties on the HER activity [109]. In the experiment, Au25(SC6H13)18, Au25(MPA)18, or Au25(MPS)18 was dissolved at a concentration of 1 mM in 0.1 M KCl aqueous solution, and LSV measurements were performed using a GCE (50 mV s−1). Although the blank current was 0.01 mA at −0.7 V vs. RHE (Figure 4C, black line), the HER current of the sample including Au25(MPA)18 increased up to 0.13 mA at −0.7 V vs. RHE (Figure 4C, red line). When Au25(MPS)18 was used, a higher HER current of 1.0 mA was observed at −0.7 V vs. RHE (Figure 4C, blue line). MPS and MPA have a hydrophilic functional group (sulfonic acid or carboxylic acid group, respectively) unlike SC6H13. These hydrophilic functional groups have the property of releasing H+ in an aqueous solution. In addition, the sulfonic acid group of MPS (pKa < 1) is expected to have higher H+ releasing ability than the carboxylic acid group of MPA (pKa = 3.7). For these reasons, it was interpreted that the difference in the HER activity described above is largely related to the difference in the H+ releasing ability of these ligands (Table 1). It was speculated that the energy barrier associated with the intermolecular and intramolecular H+ transfer steps is lowered by H+ relay in Aun NCs with high HER activity (Figure 4A). In this paper, they also reported that the use of Au24Pt(MPS)18, in which Au25(MPS)18 is replaced with Pt, results in even higher HER activity than Au25(MPS)18 (Figure 4D and Table 1). They descried that the TOF value of Au24Pt(MPS)18 was 127 mol H2 (mol catalyst)−1 s−1, which was 4 times higher than that of Au25(MPS)18 at −0.7 V vs. RHE.
An electronic interaction also occurs between the Aun(SR)m NCs and a catalytic support. This phenomenon was revealed by Jin et al., by measuring the HER activity of MoS2 nanosheets (catalytic support) carrying Au25(PET)18 (Au25(PET)18/MoS2) [108]. In this experiment, Au25(PET)18/MoS2 was prepared by mixing the MoS2 nanosheets synthesized by the hydrothermal method and Au25(PET)18 in dichloromethane for 1 h and drying the obtained products under nitrogen atmosphere. High-angle annular dark-field scanning TEM (HAADF-STEM) images confirmed that Au25(PET)18 was uniformly supported on MoS2 (Figure 5A). Au25(PET)18/MoS2 was then loaded on a GCE, and the HER polarization curve of Au25(PET)18/MoS2 was obtained by scanning the potential in a 0.5 M H2SO4 aqueous solution using the rotating disk electrode (RDE) method (Figure 5B,D). MoS2 without Au25(PET)18 exhibited a HER overvoltage of 0.33 V at a current density of 10 mA cm−2, whereas Au25(PET)18/MoS2 exhibited a smaller HER overvoltage of approximately −0.28 V at the same current density. In addition, Au25(PET)18/MoS2 (59.3 mA cm−2) exhibited a 1.79 times higher current density than that of MoS2 (33.2 mA cm−2) at an applied voltage of −0.4 V vs. RHE. Thus, the HER activity of the MoS2 nanosheets was greatly improved by carrying Au25(PET)18 (Table 1). This improvement of the HER activity was interpreted to be greatly related to the electronic interaction between Au25(PET)18 and MoS2. In fact, X-ray photoelectron spectroscopy (XPS) analysis confirmed that the binding energy of MoS2 in the Mo 3 d orbit was negatively shifted by 0.4 eV after Au25(PET)18 was loaded (Figure 5C). It was assumed that the charge transfer from Au25(PET)18 to MoS2 occurred in Au25(PET)18/MoS2, causing a high HER activity of Au25(PET)18/MoS2. In this study, the HER activity of MoS2 nanosheets carrying Au25(SePh)18 (SePh = phenylselenolate) (Au25(SePh)18/MoS2) was also investigated. Au25(SePh)18/MoS2 was shown to also exhibit higher HER activity than MoS2 nanosheets (Table 1). However, the improvement of the activity was smaller than that when carrying Au25(PET)18 (Figure 5D). This difference was attributed to the difference in the electron interaction and electron relay between Au cores of Aun NCs and the MoS2 nanosheet depending on the ligands. In this way, the HER activity of the Aun NCs-loaded catalyst was shown to depend on the electronic interaction between the Aun NCs and the catalytic support.

2.2. Oxygen Evolution Reaction

The OER is a multi-step four-electron reaction in which the reaction proceeds along different reaction paths depending on the binding energy between the metal and the OER intermediate (O, OH, and OOH).
Under acidic conditions, the following reactions occur:
M + H2O → M−OH + H+ + e
M−OH → M−O + H+ + e
2(M−O) → 2M + O2
M−O + H2O → M−OOH + H+ + e
M−OOH → M + O2 + H+ + e
However, under alkaline conditions, the following reactions occur:
M + OH → M−OH + e
M−OH + OH → M−O + H2O + e
2(M−O) → 2M + O2
M−O + OH → M−OOH + e
M−OOH + OH → M + O2 + H2O + e
As described above, because the reaction route of OER depends on the intermediates (O, OH, and OOH) on the surface of catalyst, the OER activity of the catalyst also depends on these intermediates. Catalysts that have neither too high nor too low binding energy with oxygen species are suitable for the OER. Previous studies have demonstrated that iridium oxide and ruthenium oxide have such desirable properties. Therefore, miniaturization of these metal oxides and prediction of their physical properties by theoretical calculation have been actively performed [120,121,122,123]. However, because these precious metals are expensive and have the problem of depletion, a search for low-cost catalysts is also being conducted. Related studies have shown that cobalt (Co)-based materials (oxides, hydroxides, selenides, and phosphides) can be used as good OER catalysts. Furthermore, it has been reported that when Au NPs are composited with such Co materials, the OER performance is greatly enhanced as a result of the improved electron conductivity and preferential formation of OOH intermediates on the surface of the catalyst [124,125,126].
Jin et al., have shown that these mixing effects also occur when Aun NCs are used instead of Au NPs [110]. In this study, the Au25(PET)18-loaded CoSe2 nanosheet (Au25(PET)18/CoSe2) was prepared by stirring Au25(PET)18 and CoSe2 nanosheets in dichloromethane for 1 h. HAADF-STEM analysis confirmed that Au25(PET)18 was uniformly supported on the CoSe2 nanosheets (Figure 6A,B). Au25(PET)18/CoSe2 was loaded on the GCE, and their OER polarization curves were obtained by scanning the applied potential (5 mV s−1) in 0.1 M KOH aqueous solution. The CoSe2 nanosheets without Au25(PET)18 exhibited an OER overvoltage of 0.52 V at a current density of 10 mA cm−2 (Figure 1B(b)), whereas Au25(PET)18/CoSe2 exhibited a smaller OER overvoltage of 0.43 V at the same current density (Figure 6C). XPS (Figure 6E) and Raman spectroscopy (Figure 6F) analyses revealed that the electronic interaction occurred between the Au25(PET)18 and CoSe2 nanosheet even in such a composite catalyst. Furthermore, DFT calculation revealed that the formation of the intermediate via OH is more advantageous by 0.21 eV mol−1 at the interface of Co–Au than at the surface of Co. It was thus interpreted that Au25(PET)18/CoSe2 exhibited higher OER activity than the CoSe2 nanosheets because Au25(PET)18/CoSe2 stabilized the generation of an OOH intermediate compared with only the CoSe2 nanosheet (Table 2). This study also revealed that the OER activity increases with the core size of Aun(SR)m NCs (Figure 6D).

3. Electrocatalytic Reactions in Fuel Cells

To establish a circulating energy system that does not use fossil fuels and only produces water and a small amount of carbon dioxide as waste, it is essential to further improve the functions of fuel cells. Fuel cells can be roughly classified into those using hydrogen and those using alcohol as a fuel. In fuel cells using hydrogen as a fuel, the HOR and ORR are involved in the system. The HOR is a one-electron reaction, and generally an HER-active catalyst is also useful for the HOR. However, the ORR is a four-electron reaction, and the reaction process is complicated. In addition, the OER is a reaction under oxidizing conditions, whereas the ORR is a reaction under reducing conditions. The surface state of the catalyst and the accompanying binding to the reactants also differ greatly between the OER and ORR. Therefore, catalysts that are active for OER are not necessarily useful for the ORR. Because the ORR is rate-limiting step in a fuel cell, controlling the ORR is important for further development of fuel cells. The ORR pathways under acidic and alkaline conditions are as follows [94].
Under acidic conditions:
O2 + 4H+ + 4e → 2H2O
O2 + 2H+ + 2e → H2O2
H2O2 + 2H+ + 2e → 2H2O
Under alkaline conditions:
O2 + 2H2O + 4e → 4OH
O2 + H2O + 2e → OOH + OH
OOH + H2O + 2e → 3OH
Equations (17) and (20) are four-electron reactions, and Equations (18), (19), (21), and (22) are two-electron reactions. For both sets of reactions, the reactions start with the breaking of the O−O bond. The theoretical redox potential is 1.23 V vs. SHE in the direct four-electron path and 0.68 V vs. SHE in the indirect two-electron path. Therefore, a higher energy conversion efficiency can be achieved using the direct four-electron path, and this reaction path is thus more desirable for fuel cells [81]. Although Pt is a useful catalyst for such a reaction pathway, it is expected to be replaced with another metal element because of the high cost of Pt and the resource depletion issue. In addition, synthesis methods of Ptn NCs in ambient atmosphere with atomic precision are limited, and therefore, it is difficult to study the ORR mechanism using Ptn NCs as model catalysts. However, for Aun NCs, there are many examples of synthesis with atomic precision, and these catalysts are stable in ambient atmosphere. In addition, theoretical calculations [127,128] and experimental results [65,129] have predicted that O2 molecules can be highly activated on the surface of Aun NCs. For these reasons, several studies have also been performed on the application of Aun NCs as ORR catalysts.
In 2009, Chen et al., evaluated the ORR catalytic activity of Au11(PPh3)8Cl3, Au25(PET)18, Au55(PPh3)12Cl6, and Au140(SC6H13)53 (Cl = chlorine) [111]. In this experiment, after a series of Aun NCs were loaded on the GCE, the ORR activity was measured by scanning the potential using the RDE method in a 0.1 M KOH aqueous solution filled with O2. When Au11(PPh3)8Cl3 was used as the Aun NCs, the onset potential of the ORR (Figure 1B(c)) was about −0.08 V, and the peak current density was 2.4 mA cm−2 (Figure 7A). However, when Au140(SC6H13)53 was used as the Aun NCs, the onset potential shifted to the more cathodic −0.22 V and the reduction peak current decreased to less than 1.0 mA cm−2. These results and those for the other two Aun NCs indicated that the ORR activity increased with decreasing Au core size (Au11(PPh3)8Cl3 > Au25(PET)18 > Au55(PPh3)12Cl6 > Au140(SC6H13)53) (Figure 7A,B and Table 3). From estimation of the number of electrons for the ORR from a Koutecky–Levich plot [85], it was observed that the relatively small size of Aun NCs (Au11(PPh3)8Cl3, Au25(PET)18, and Au55(PPh3)12Cl6) resulted in the occurrence of the four-electron reaction, whereas Au140(SC6H13)53 tended to follow the two-electron reaction pathway (Figure 7C,D). Later, these researchers also synthesized a series of Aun(SR)m NCs (Au25(PET)18, Au38(PET)24, and Au144(PET)60) with PET ligands and measured their ORR activities. The results revealed that a smaller core size was associated with higher ORR activity: Au25(PET)18 > Au38(PET)24 > Au144(PET)60 (Table 3) [112]. As the core size decreased, the ratio of low-coordinated surface atoms increased and the d-band center of the Fermi level changed. It was interpreted that smaller Aun(SR)m NCs exhibited higher ORR activity because the promotion of oxygen adsorption on the gold core surface was accelerated by miniaturization of the metal core.
On the other hand, Dass et al., studied the dependence of the ORR activity on the core size using Aun NCs protected by 4-tert-butylbenzenethiolate (TBBT), whose structure differs significantly from that of PET [113]. In this experiment, single-walled carbon nanotubes (SWNTs) carrying Aun(TBBT)m NCs (n = 28, 36, 133, and 279; Figure 8A; Aun(TBBT)m NCs/SWNTs) were loaded onto the GCE. The ORR actives were measured by scanning the potential using the RDE method in a 0.1 M KOH aqueous solution filled with O2 (Figure 8B). The overvoltage of the ORR was smaller in the order of Au36(TBBT)24 > Au133(TBBT)52 > Au279(TBBT)84 > Au28(TBBT)20. However, the selectivity of the four-electron reduction reaction was superior in the order of Au36(TBBT)24 ≈ Au133(TBBT)52 > Au279(TBBT)84 > Au28(TBBT)20 [113] (Figure 8C). Notably, this trend was similar to that of the size dependence of the stability of Aun(TBBT)m NCs itself. The same group performed similar studies using tert-butylthiolate (S-tBu) instead of TBBT as a ligand [114]. S-tBu has a bulky framework and when this ligand is used in the synthesis of Aun(SR)m NCs, the ratio of the metal atom and the ligand in the generated Aun(SR)m NCs is different from that in Aun(SR)m NCs synthesized using another ligand. Such Aun(S-tBu)m NCs exhibit a unique size dependency for ORR activity (Au65(S-tBu)29 > Au46(S-tBu)24 > Au30(S-tBu)18 > Au23(S-tBu)16) [114].
In addition to these effects of core sizes and ligands, the ORR activity also depended on the charge state of Aun(SR)m NCs. Chen et al., carried [Au25(SC12H25)18], [Au25(SC12H25)18]0, and [Au25(SC12H25)18]+ (SC12H25 = 1-dodecanethiolate) on the GCE, and their ORR activities were evaluated by scanning the potential in a 0.1 M KOH aqueous solution using a rotating ring-disk electrode (RRDE) filled with O2 [115]. In addition, the generation of H2O2 was evaluated from the RRDE current at a fixed ring potential (0.5 V vs. saturated calomel electrode (SCE)). When [Au25(SC12H25)18], [Au25(SC12H25)18]0, and [Au25(SC12H25)18]+ were used, the efficiencies of H2O2 were 86%, 82%, and 72%, respectively. In addition, the number of electrons for the ORR was estimated to be 2.28 ([Au25(SC12H25)18]), 2.35 ([Au25(SC12H25)18]0), and 2.56 ([Au25(SC12H25)18]+; Figure 9A–C). For [Au25(SC12H25)18], which showed the highest production rate of H2O2, the activity decreased only 9% even after 1000 cycles (Figure 9D). These results indicate that [Au25(SC12H25)18] has high H2O2 generating ability (Table 3) [115]. Since H2O2 is a useful raw material for chemical products, the development of their highly selective production reactions is important. Jin et al., also studied the dependence of the ORR activity on the charge state of Aun(SR)m NCs using [Au25(PET)18], [Au25(PET)18]0, and [Au25(PET)18]+. They reported that too strong of an OH adsorbing ability of [Au25(PET)18]+ reduces the ORR activity [77]. Thus, it has been clarified that the charge state of Aun(SR)m NCs also has a significant effect on the ORR activity of Aun(SR)m NCs.

4. Conclusions

A system for the generation of a fuel such as hydrogen or methanol using natural energy (e.g., solar cells or photocatalytic water splitting) and the production of electricity by fuel cells using these fuels would be one of the ultimate energy conversion systems for our society. To realize such a system, high activation of the HER, OER, HOR, and ORR is indispensable. Recently, Aun NCs have attracted considerable attention as model catalysts for these reactions. In this review, recent works on these materials were summarized. The overall characteristics of the HER, OER, and ORR can be summarized as follows.
1) Since the core size, doping metal, ligand structure, and charge state affect the electronic and geometrical structures of Aun NCs, these parameters also have a great effect on the catalytic activity of Aun NCs.
2) Although these three reactions proceed via different mechanisms, reducing the core size of Aun NCs and improving the ligand conductivity tend to improve the activities.
3) When Aun NCs are carried on a conventional catalytic support, their electronic structure changes and thus their catalytic activity also changes. Therefore, Aun NCs are also useful for improving the catalytic activity of conventional catalytic materials.

5. Perspectives

Until recently, the materials with relatively high activity for all of HER, OER, and ORR are considered to be limited to Ir, Rh, Ru, and Pt [84,85]. However, the recent studies demonstrated that these properties could also be caused in Au by the discretization of the band structure (e.g., shift of d-band center [107,111]). For Aun NCs, it is possible to precisely control the electronic/geometrical structures and thereby to elucidate the correlation between catalytic activity and electronic/geometrical structure. In addition, the use of fine Aun NCs as a catalyst is effective in reducing the consumption of expensive noble metals. It is expected that the studies on the catalytic activities of Aun NCs lead to solve the mechanism in catalytic reactions on the metal surface and create the amazing catalysts we have never seen.
However, to create such HER, OER, and ORR catalysts using Aun NCs and their alloy NCs, further studies are required. Previous studies have shown that doping with Group 10 elements (Pt and Pd) induces high activation. Thus, a method for increasing the doping concentration of these elements is expected to be developed in the future. In addition, regarding the HER and OER, in spite of decomposing water, most studies thus far have used hydrophobic ligands that are not compatible with water. This may be related to the fact that the synthesis of hydrophobic Aun NCs is easier than that of hydrophilic Aun NCs. In particular, it is difficult to selectively synthesize a group-10-element-doped cluster using a hydrophilic ligand using the conventional synthesis method. However, as shown in this review, it is more appropriate to use hydrophilic Aun NCs as HER and OER catalysts. Therefore, in the future, additional research on hydrophilic Aun NCs is expected to increase the types of ligands and core sizes of hydrophilic Aun NCs. Such studies are expected to lead to the creation of highly active HER, OER, and ORR catalysts and eventually to the development of design guidelines for establishing ultimate energy conversion systems.

Author Contributions

T.K. and Y.N. developed the idea for this review article and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Japan Society for the Promotion of Science (JSPS) KAKENHI (grant number JP18K14074 and JP18H01953), Scientific Research on Innovative Areas “Coordination Asymmetry” (grant number 17H05385) and Scientific Research on Innovative Areas “Innovations for Light-Energy Conversion” (grant number 18H05178).

Conflicts of Interest

There are no conflicts to declare.

References

  1. Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D.J.; Whyman, R. Synthesis of Thiol-Derivatised Gold Nanoparticles in a Two-Phase Liquid–Liquid System. J. Chem. Soc. Chem. Commun. 1994, 801–802. [Google Scholar] [CrossRef]
  2. Jin, R.; Zeng, C.; Zhou, M.; Chen, Y. Atomically Precise Colloidal Metal Nanoclusters and Nanoparticles: Fundamentals and Opportunities. Chem. Rev. 2016, 116, 10346–10413. [Google Scholar] [CrossRef] [PubMed]
  3. Kurashige, W.; Niihori, Y.; Sharma, S.; Negishi, Y. Precise Synthesis, Functionalization and Application of Thiolate-Protected Gold Clusters. Coord. Chem. Rev. 2016, 320, 238–250. [Google Scholar] [CrossRef]
  4. Hossain, S.; Niihori, Y.; Nair, L.V.; Kumar, B.; Kurashige, W.; Negishi, Y. Alloy Clusters: Precise Synthesis and Mixing Effects. Acc. Chem. Res. 2018, 51, 3114–3124. [Google Scholar] [CrossRef] [PubMed]
  5. Gan, Z.; Xia, N.; Wu, Z. Discovery, Mechanism; Application of Antigalvanic Reaction. Acc. Chem. Res. 2018, 51, 2774–2783. [Google Scholar] [CrossRef] [PubMed]
  6. Chakraborty, I.; Pradeep, T. Atomically Precise Clusters of Noble Metals: Emerging Link between Atoms and Nanoparticles. Chem. Rev. 2017, 117, 8208–8271. [Google Scholar] [CrossRef]
  7. Yao, Q.; Chen, T.; Yuan, X.; Xie, J. Toward Total Synthesis of Thiolate-Protected Metal Nanoclusters. Acc. Chem. Res. 2018, 51, 1338–1348. [Google Scholar] [CrossRef]
  8. Qian, H.; Zhu, M.; Wu, Z.; Jin, R. Quantum Sized Gold Nanoclusters with Atomic Precision. Acc. Chem. Res. 2012, 45, 1470–1479. [Google Scholar] [CrossRef]
  9. Whetten, R.L.; Weissker, H.-C.; Pelayo, J.J.; Mullins, S.M.; López-Lozano, X.; Garzón, I.L. Chiral-Icosahedral (I) Symmetry in Ubiquitous Metallic Cluster Compounds (145A,60X): Structure and Bonding Principles. Acc. Chem. Res. 2019, 52, 34–43. [Google Scholar] [CrossRef] [Green Version]
  10. Aikens, C.M. Electronic and Geometric Structure, Optical Properties, and Excited State Behavior in Atomically Precise Thiolate-Stabilized Noble Metal Nanoclusters. Acc. Chem. Res. 2018, 51, 3065–3073. [Google Scholar] [CrossRef]
  11. Nieto-Ortega, B.; Bürgi, T. Vibrational Properties of Thiolate-Protected Gold Nanoclusters. Acc. Chem. Res. 2018, 51, 2811–2819. [Google Scholar] [CrossRef] [PubMed]
  12. Agrachev, M.; Ruzzi, M.; Venzo, A.; Maran, F. Nuclear and Electron Magnetic Resonance Spectroscopies of Atomically Precise Gold Nanoclusters. Acc. Chem. Res. 2019, 52, 44–52. [Google Scholar] [CrossRef] [PubMed]
  13. Pei, Y.; Wang, P.; Ma, Z.; Xiong, L. Growth-Rule-Guided Structural Exploration of Thiolate-Protected Gold Nanoclusters. Acc. Chem. Res. 2019, 52, 23–33. [Google Scholar] [CrossRef] [PubMed]
  14. Ghosh, A.; Mohammed, O.F.; Bakr, O.M. Atomic-Level Doping of Metal Clusters. Acc. Chem. Res. 2018, 51, 3094–3103. [Google Scholar] [CrossRef] [Green Version]
  15. Bigioni, T.P.; Whetten, R.L.; Dag, Ö. Near-Infrared Luminescence from Small Gold Nanocrystals. J. Phys. Chem. B 2000, 104, 6983–6986. [Google Scholar] [CrossRef]
  16. Yan, J.; Teo, B.K.; Zheng, N. Surface Chemistry of Atomically Precise Coinage–Metal Nanoclusters: From Structural Control to Surface Reactivity and Catalysis. Acc. Chem. Res. 2018, 51, 3084–3093. [Google Scholar] [CrossRef]
  17. Sakthivel, N.A.; Dass, A. Aromatic Thiolate-Protected Series of Gold Nanomolecules and a Contrary Structural Trend in Size Evolution. Acc. Chem. Res. 2018, 51, 1774–1783. [Google Scholar] [CrossRef]
  18. Tang, Q.; Hu, G.; Fung, V.; Jiang, D.-E. Insights into Interfaces, Stability, Electronic Properties, and Catalytic Activities of Atomically Precise Metal Nanoclusters from First Principles. Acc. Chem. Res. 2018, 51, 2793–2802. [Google Scholar] [CrossRef]
  19. Negishi, Y.; Nobusada, K.; Tsukuda, T. Glutathione-Protected Gold Clusters Revisited:  Bridging the Gap between Gold(I)−Thiolate Complexes and Thiolate-Protected Gold Nanocrystals. J. Am. Chem. Soc. 2005, 127, 5261–5270. [Google Scholar] [CrossRef]
  20. Jadzinsky, P.D.; Calero, G.; Ackerson, C.J.; Bushnell, D.A.; Kornberg, R.D. Structure of a Thiol Monolayer-Protected Gold Nanoparticle at 1.1 Å Resolution. Science 2007, 318, 430–433. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, S.; Li, Q.; Kang, X.; Zhu, M. Customizing the Structure, Composition, and Properties of Alloy Nanoclusters by Metal Exchange. Acc. Chem. Res. 2018, 51, 2784–2792. [Google Scholar] [CrossRef] [PubMed]
  22. Higaki, T.; Li, Q.; Zhou, M.; Zhao, S.; Li, Y.; Li, S.; Jin, R. Toward the Tailoring Chemistry of Metal Nanoclusters for Enhancing Functionalities. Acc. Chem. Res. 2018, 51, 2764–2773. [Google Scholar] [CrossRef] [PubMed]
  23. Negishi, Y.; Kurashige, W.; Niihori, Y.; Iwasa, T.; Nobusada, K. Isolation, Structure, and Stability of a Dodecanethiolate-Protected Pd1Au24 Cluster. Phys. Chem. Chem. Phys. 2010, 12, 6219–6225. [Google Scholar] [CrossRef]
  24. Negishi, Y.; Iwai, T.; Ide, M. Continuous Modulation of Electronic Structure of Stable Thiolate-Protected Au25 Cluster by Ag Doping. Chem. Commun. 2010, 46, 4713–4715. [Google Scholar] [CrossRef] [PubMed]
  25. Negishi, Y.; Igarashi, K.; Munakata, K.; Ohgake, W.; Nobusada, K. Palladium Doping of Magic Gold Cluster Au38(SC2H4Ph)24: Formation of Pd2Au36(SC2H4Ph)24 with Higher Stability than Au38(SC2H4Ph)24. Chem. Commun. 2012, 48, 660–662. [Google Scholar] [CrossRef] [PubMed]
  26. Negishi, Y.; Munakata, K.; Ohgake, W.; Nobusada, K. Effect of Copper Doping on Electronic Structure, Geometric Structure, and Stability of Thiolate-Protected Au25 Nanoclusters. J. Phys. Chem. Lett. 2012, 3, 2209–2214. [Google Scholar] [CrossRef]
  27. Niihori, Y.; Kurashige, W.; Matsuzaki, M.; Negishi, Y. Remarkable Enhancement in Ligand-Exchange Reactivity of Thiolate-Protected Au25 Nanoclusters by Single Pd Atom Doping. Nanoscale 2013, 5, 508–512. [Google Scholar] [CrossRef]
  28. Negishi, Y.; Kurashige, W.; Kobayashi, Y.; Yamazoe, S.; Kojima, N.; Seto, M.; Tsukuda, T. Formation of a Pd@Au12 Superatomic Core in Au24Pd1(SC12H25)18 Probed by 197Au Mössbauer and Pd K-edge EXAFS Spectroscopy. J. Phys. Chem. Lett. 2013, 4, 3579–3583. [Google Scholar] [CrossRef]
  29. Negishi, Y.; Kurashige, W.; Niihori, Y.; Nobusada, K. Toward the Creation of Stable, Functionalized Metal Clusters. Phys. Chem. Chem. Phys. 2013, 15, 18736–18751. [Google Scholar] [CrossRef]
  30. Niihori, Y.; Matsuzaki, M.; Uchida, C.; Negishi, Y. Advanced Use of High-Performance Liquid Chromatography for Synthesis of Controlled Metal Clusters. Nanoscale 2014, 6, 7889–7896. [Google Scholar] [CrossRef]
  31. Yamazoe, S.; Kurashige, W.; Nobusada, K.; Negishi, Y.; Tsukuda, T. Preferential Location of Coinage Metal Dopants (M = Ag or Cu) in [Au25-xMx(SC2H4Ph)18] (x ~ 1) as Determined by Extended X-ray Absorption Fine Structure and Density Functional Theory Calculations. J. Phys. Chem. C 2014, 118, 25284–25290. [Google Scholar] [CrossRef]
  32. Sharma, S.; Kurashige, W.; Nobusada, K.; Negishi, Y. Effect of Trimetallization in Thiolate-Protected Au24−nCunPd Clusters. Nanoscale 2015, 7, 10606–10612. [Google Scholar] [CrossRef] [PubMed]
  33. Niihori, Y.; Eguro, M.; Kato, A.; Sharma, S.; Kumar, B.; Kurashige, W.; Nobusada, K.; Negishi, Y. Improvements in the Ligand-Exchange Reactivity of Phenylethanethiolate-Protected Au25 Nanocluster by Ag or Cu Incorporation. J. Phys. Chem. C 2016, 120, 14301–14309. [Google Scholar] [CrossRef]
  34. Niihori, Y.; Hossain, S.; Kumar, B.; Nair, L.V.; Kurashige, W.; Negishi, Y. Perspective: Exchange Reactions in Thiolate-Protected Metal Clusters. APL Mater. 2017, 5, 053201. [Google Scholar] [CrossRef]
  35. Niihori, Y.; Hossain, S.; Sharma, S.; Kumar, B.; Kurashige, W.; Negishi, Y. Understanding and Practical Use of Ligand and Metal Exchange Reactions in Thiolate-Protected Metal Clusters to Synthesize Controlled Metal Clusters. Chem. Rec. 2017, 17, 473–484. [Google Scholar] [CrossRef] [PubMed]
  36. Niihori, Y.; Shima, D.; Yoshida, K.; Hamada, K.; Nair, L.V.; Hossain, S.; Kurashige, W.; Negishi, Y. High-Performance Liquid Chromatography Mass Spectrometry of Gold and Alloy Clusters Protected by Hydrophilic Thiolates. Nanoscale 2018, 10, 1641–1649. [Google Scholar] [CrossRef]
  37. Hossain, S.; Ono, T.; Yoshioka, M.; Hu, G.; Hosoi, M.; Chen, Z.; Nair, L.V.; Niihori, Y.; Kurashige, W.; Jiang, D.-E.; et al. Thiolate-Protected Trimetallic Au~20Ag~4Pd and Au~20Ag~4Pt Alloy Clusters with Controlled Chemical Composition and Metal Positions. J. Phys. Chem. Lett. 2018, 9, 2590–2594. [Google Scholar] [CrossRef]
  38. Yokoyama, T.; Hirata, N.; Tsunoyama, H.; Negishi, Y.; Nakajima, A. Characterization of Floating-gate Memory Device with Thiolate-Protected Gold and Gold-Palladium Nanoclusters. AIP Adv. 2018, 8, 065002. [Google Scholar] [CrossRef] [Green Version]
  39. Niihori, Y.; Koyama, Y.; Watanabe, S.; Hashimoto, S.; Hossain, S.; Nair, L.V.; Kumar, B.; Kurashige, W.; Negishi, Y. Atomic and Isomeric Separation of Thiolate-Protected Alloy Clusters. J. Phys. Chem. Lett. 2018, 9, 4930–4934. [Google Scholar] [CrossRef]
  40. Niihori, Y.; Hashimoto, S.; Koyama, Y.; Hossain, S.; Kurashige, W.; Negishi, Y. Dynamic Behavior of Thiolate-Protected Gold–Silver 38-Atom Alloy Clusters in Solution. J. Phys. Chem. C 2019, 123, 13324–13329. [Google Scholar] [CrossRef]
  41. Kurashige, W.; Hayashi, R.; Wakamatsu, K.; Kataoka, Y.; Hossain, S.; Iwase, A.; Kudo, A.; Yamazoe, S.; Negishi, Y. Atomic-Level Understanding of the Effect of Heteroatom Doping of the Cocatalyst on Water-Splitting Activity in AuPd or AuPt Alloy Cluster-Loaded BaLa4Ti4O15. ACS Appl. Energy Mater. 2019, 2, 4175–4187. [Google Scholar] [CrossRef]
  42. Hossain, S.; Imai, Y.; Suzuki, D.; Choi, W.; Chen, Z.; Suzuki, T.; Yoshioka, M.; Kawawaki, T.; Lee, D.; Negishi, Y. Elucidating Ligand Effects in Thiolate-Protected Metal Clusters Using Au24Pt(TBBT)18 as a Model Cluster. Nanoscale 2019, 11, 22089–22098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kawawaki, T.; Negishi, Y.; Kawasaki, H. Photo/electrocatalysis and Photosensitization Using Metal Nanoclusters for Green Energy and Medical Applications. Nanoscale Adv. 2020, 2, 17–36. [Google Scholar] [CrossRef] [Green Version]
  44. Hossain, S.; Imai, Y.; Motohashi, Y.; Chen, Z.; Suzuki, D.; Suzuki, T.; Kataoka, Y.; Hirata, M.; Ono, T.; Kurashige, W.; et al. Understanding and Designing One-Dimensional Assemblies of Ligand-Protected Metal Nanoclusters. Mater. Horiz. in press. [CrossRef] [Green Version]
  45. Haruta, M. Spiers Memorial Lecture Role of Perimeter Interfaces in Catalysis by Gold Nanoparticles. Faraday Discuss. 2011, 152, 11–32. [Google Scholar] [CrossRef] [PubMed]
  46. Nie, X.; Qian, H.; Ge, Q.; Xu, H.; Jin, R. CO Oxidation Catalyzed by Oxide-Supported Au25(SR)18 Nanoclusters and Identification of Perimeter Sites as Active Centers. ACS Nano 2012, 6, 6014–6022. [Google Scholar] [CrossRef]
  47. Nie, X.; Zeng, C.; Ma, X.; Qian, H.; Ge, Q.; Xu, H.; Jin, R. CeO2-Supported Au38(SR)24 Nanocluster Catalysts for CO Oxidation: A Comparison of Ligand-on and -off Catalysts. Nanoscale 2013, 5, 5912–5918. [Google Scholar] [CrossRef]
  48. Wu, Z.; Jiang, D.-e.; Mann, A.K.P.; Mullins, D.R.; Qiao, Z.-A.; Allard, L.F.; Zeng, C.; Jin, R.; Overbury, S.H. Thiolate Ligands as a Double-Edged Sword for CO Oxidation on CeO2 Supported Au25(SCH2CH2Ph)18 Nanoclusters. J. Am. Chem. Soc. 2014, 136, 6111–6122. [Google Scholar] [CrossRef]
  49. Li, W.; Ge, Q.; Ma, X.; Chen, Y.; Zhu, M.; Xu, H.; Jin, R. Mild Activation of CeO2-Supported Gold Nanoclusters and Insight into the Catalytic Behavior in CO Oxidation. Nanoscale 2016, 8, 2378–2385. [Google Scholar] [CrossRef]
  50. Gaur, S.; Miller, J.T.; Stellwagen, D.; Sanampudi, A.; Kumar, C.S.S.R.; Spivey, J.J. Synthesis, Characterization, and Testing of Supported Au Catalysts Prepared from Atomically-Tailored Au38(SC12H25)24 Clusters. Phys. Chem. Chem. Phys. 2012, 14, 1627–1634. [Google Scholar] [CrossRef]
  51. Wu, Z.; Hu, G.; Jiang, D.-e.; Mullins, D.R.; Zhang, Q.-F.; Allard, L.F.; Wang, L.-S.; Overbury, S.H. Diphosphine-Protected Au22 Nanoclusters on Oxide Supports Are Active for Gas-Phase Catalysis without Ligand Removal. Nano Lett. 2016, 16, 6560–6567. [Google Scholar] [CrossRef]
  52. Wu, Z.; Mullins, D.R.; Allard, L.F.; Zhang, Q.; Wang, L. CO Oxidation over Ceria Supported Au22 Nanoclusters: Shape Effect of the Support. Chin. Chem. Lett. 2018, 29, 795–799. [Google Scholar] [CrossRef]
  53. Lin, J.; Li, W.; Liu, C.; Huang, P.; Zhu, M.; Ge, Q.; Li, G. One-Phase Controlled Synthesis of Au25 Nanospheres and Nanorods from 1.3 nm Au : PPh3 Nanoparticles: The Ligand Effects. Nanoscale 2015, 7, 13663–13670. [Google Scholar] [CrossRef] [PubMed]
  54. Li, W.; Liu, C.; Abroshan, H.; Ge, Q.; Yang, X.; Xu, H.; 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]
  55. Good, J.; Duchesne, P.N.; Zhang, P.; Koshut, W.; Zhou, M.; Jin, R. On the Functional Role of the Cerium Oxide Support in the Au38(SR)24/CeO2 Catalyst for CO Oxidation. Catal. Today 2017, 280, 239–245. [Google Scholar] [CrossRef] [Green Version]
  56. Du, Y.; Sheng, H.; Astruc, D.; Zhu, M. Atomically Precise Noble Metal Nanoclusters as Efficient Catalysts: A Bridge between Structure and Properties. Chem. Rev. 2020, 120, 526–622. [Google Scholar] [CrossRef]
  57. Xie, S.; 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]
  58. Yoskamtorn, T.; Yamazoe, S.; Takahata, R.; Nishigaki, J.-i.; Thivasasith, A.; Limtrakul, J.; Tsukuda, T. Thiolate-Mediated Selectivity Control in Aerobic Alcohol Oxidation by Porous Carbon-Supported Au25 Clusters. ACS Catal. 2014, 4, 3696–3700. [Google Scholar] [CrossRef]
  59. Lavenn, C.; Demessence, A.; Tuel, A. Au25(SPh-pNH2)17 Nanoclusters Deposited on SBA-15 as Catalysts for Aerobic Benzyl Alcohol Oxidation. J. Catal. 2015, 322, 130–138. [Google Scholar] [CrossRef]
  60. Deng, H.; Wang, S.; Jin, S.; Yang, S.; Xu, Y.; Liu, L.; Xiang, J.; Hu, D.; Zhu, M. Active Metal (Cadmium) Doping Enhanced the Stability of Inert Metal (Gold) Nanocluster under O2 Atmosphere and the Catalysis Activity of Benzyl Alcohol Oxidation. Gold Bull. 2015, 48, 161–167. [Google Scholar] [CrossRef] [Green Version]
  61. Li, L.; Dou, L.; Zhang, H. Layered Double Hydroxide Supported Gold Nanoclusters by Glutathione-Capped Au Nanoclusters Precursor Method for Highly Efficient Aerobic Oxidation of Alcohols. Nanoscale 2014, 6, 3753–3763. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, S.; Yin, S.; Chen, G.; Li, L.; Zhang, H. Nearly Atomic Precise Gold Nanoclusters on Nickel-Based Layered Double Hydroxides for Extraordinarily Efficient Aerobic Oxidation of Alcohols. Catal. Sci. Technol. 2016, 6, 4090–4104. [Google Scholar] [CrossRef]
  63. Yin, S.; Li, J.; Zhang, H. Hierarchical Hollow Nanostructured Core@Shell Recyclable Catalysts γ-Fe2O3@LDH@Au25-X for Highly Efficient Alcohol Oxidation. Green Chem. 2016, 18, 5900–5914. [Google Scholar] [CrossRef]
  64. Lee, K.E.; Shivhare, A.; Hu, Y.; Scott, R.W.J. Supported Bimetallic AuPd Clusters Using Activated Au25 Clusters. Catal. Today 2017, 280, 259–265. [Google Scholar] [CrossRef] [Green Version]
  65. Tsunoyama, H.; Ichikuni, N.; Sakurai, H.; Tsukuda, T. Effect of Electronic Structures of Au Clusters Stabilized by Poly(N-vinyl-2-pyrrolidone) on Aerobic Oxidation Catalysis. J. Am. Chem. Soc. 2009, 131, 7086–7093. [Google Scholar] [CrossRef]
  66. Zhu, Y.; Qian, H.; Zhu, M.; Jin, R. Thiolate-Protected Aun Nanoclusters as Catalysts for Selective Oxidation and Hydrogenation Processes. Adv. Mater. 2010, 22, 1915–1920. [Google Scholar] [CrossRef]
  67. Zhu, Y.; Qian, H.; Jin, R. An Atomic-Level Strategy for Unraveling Gold Nanocatalysis from the Perspective of Aun(SR)m Nanoclusters. Chem. Eur. J. 2010, 16, 11455–11462. [Google Scholar] [CrossRef]
  68. Wang, S.; Jin, S.; Yang, S.; Chen, S.; Song, Y.; Zhang, J.; Zhu, M. Total Structure Determination of Surface Doping [Ag46Au24(SR)32](BPh4)2 Nanocluster and Its Structure-Related Catalytic Property. Science Adv. 2015, 1, e1500441. [Google Scholar] [CrossRef] [Green Version]
  69. Chai, J.; Chong, H.; Wang, S.; Yang, S.; Wu, M.; Zhu, M. Controlling the Selectivity of Catalytic Oxidation of Styrene over Nanocluster Catalysts. RSC Adv. 2016, 6, 111399–111405. [Google Scholar] [CrossRef]
  70. Turner, M.; Golovko, V.B.; Vaughan, O.P.H.; Abdulkin, P.; Berenguer-Murcia, A.; Tikhov, M.S.; Johnson, B.F.G.; Lambert, R.M. Selective Oxidation with Dioxygen by Gold Nanoparticle Catalysts Derived from 55-Atom Clusters. Nature 2008, 454, 981–983. [Google Scholar] [CrossRef]
  71. Zhang, B.; Kaziz, S.; Li, H.; Hevia, M.G.; Wodka, D.; Mazet, C.; Bürgi, T.; Barrabés, N. Modulation of Active Sites in Supported Au38(SC2H4Ph)24 Cluster Catalysts: Effect of Atmosphere and Support Material. J. Phys. Chem. C 2015, 119, 11193–11199. [Google Scholar] [CrossRef]
  72. Liu, Y.; Tsunoyama, H.; Akita, T.; Xie, S.; Tsukuda, T. Aerobic Oxidation of Cyclohexane Catalyzed by Size-Controlled Au Clusters on Hydroxyapatite: Size Effect in the Sub-2 nm Regime. ACS Catal. 2011, 1, 2–6. [Google Scholar] [CrossRef]
  73. Liu, C.; Yan, C.; Lin, J.; Yu, C.; Huang, J.; 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]
  74. Li, G.; Qian, H.; Jin, R. Gold Nanocluster-Catalyzed Selective Oxidation of Sulfide to Sulfoxide. Nanoscale 2012, 4, 6714–6717. [Google Scholar] [CrossRef]
  75. Chen, Y.; Wang, J.; Liu, C.; Li, Z.; Li, G. Kinetically Controlled Synthesis of Au102(SPh)44 Nanoclusters and Catalytic Application. Nanoscale 2016, 8, 10059–10065. [Google Scholar] [CrossRef] [PubMed]
  76. Kauffman, D.R.; Alfonso, D.; Matranga, C.; Ohodnicki, P.; Deng, X.; Siva, R.C.; Zeng, C.; Jin, R. Probing Active Site Chemistry with Differently Charged Au25q Nanoclusters (q = −1, 0, +1). Chem. Sci. 2014, 5, 3151–3157. [Google Scholar] [CrossRef]
  77. 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]
  78. Zhao, S.; Jin, R.; Jin, R. Opportunities and Challenges in CO2 Reduction by Gold- and Silver-Based Electrocatalysts: From Bulk Metals to Nanoparticles and Atomically Precise Nanoclusters. ACS Energy Lett. 2018, 3, 452–462. [Google Scholar] [CrossRef]
  79. Andrews, E.; Katla, S.; Kumar, C.; Patterson, M.; Sprunger, P.; Flake, J. Electrocatalytic Reduction of CO2 at Au Nanoparticle Electrodes: Effects of Interfacial Chemistry on Reduction Behavior. J. Electrochem. Soc. 2015, 162, F1373–F1378. [Google Scholar] [CrossRef]
  80. Kauffman, D.R.; Thakkar, J.; Siva, R.; Matranga, C.; Ohodnicki, P.R.; Zeng, C.; Jin, R. Efficient Electrochemical CO2 Conversion Powered by Renewable Energy. ACS Appl. Mater. Inter. 2015, 7, 15626–15632. [Google Scholar] [CrossRef]
  81. Zhao, S.; Austin, N.; Li, M.; Song, Y.; House, S.D.; Bernhard, S.; Yang, J.C.; Mpourmpakis, G.; Jin, R. Influence of Atomic-Level Morphology on Catalysis: The Case of Sphere and Rod-Like Gold Nanoclusters for CO2 Electroreduction. ACS Catal. 2018, 8, 4996–5001. [Google Scholar] [CrossRef]
  82. Jupally, V.R.; Dharmaratne, A.C.; Crasto, D.; Huckaba, A.J.; Kumara, C.; Nimmala, P.R.; Kothalawala, N.; Delcamp, J.H.; Dass, A. Au137(SR)56 Nanomolecules: Composition, Optical Spectroscopy, Electrochemistry and Electrocatalytic Reduction of CO2. Chem. Commun. 2014, 50, 9895–9898. [Google Scholar] [CrossRef] [PubMed]
  83. Alfonso, D.R.; Kauffman, D.; Matranga, C. Active Sites of Ligand-Protected Au25 Nanoparticle Catalysts for CO2 Electroreduction to CO. J. Chem. Phys. 2016, 144, 184705. [Google Scholar] [CrossRef] [PubMed]
  84. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining Theory and Experiment in Electrocatalysis: Insights into Materials Design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Design of Electrocatalysts for Oxygen- and Hydrogen-Involving Energy Conversion Reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef]
  86. Sabatier, P. Hydrogénations Et Déshydrogénations Par Catalyse. Ber. Dtsch. Chem. Ges. 1911, 44, 1984–2001. [Google Scholar] [CrossRef] [Green Version]
  87. Parsons, R. The Rate of Electrolytic Hydrogen Evolution and the Heat of Adsorption of Hydrogen. Trans. Faraday Soc. 1958, 54, 1053–1063. [Google Scholar] [CrossRef]
  88. Benck, J.D.; Hellstern, T.R.; Kibsgaard, J.; Chakthranont, P.; Jaramillo, T.F. Catalyzing the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide Nanomaterials. ACS Catal. 2014, 4, 3957–3971. [Google Scholar] [CrossRef]
  89. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [Green Version]
  90. Cao, B.; Veith, G.M.; Neuefeind, J.C.; Adzic, R.R.; Khalifah, P.G. Mixed Close-Packed Cobalt Molybdenum Nitrides as Non-Noble Metal Electrocatalysts for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 19186–19192. [Google Scholar] [CrossRef]
  91. Popczun, E.J.; McKone, J.R.; Read, C.G.; Biacchi, A.J.; Wiltrout, A.M.; Lewis, N.S.; Schaak, R.E. Nanostructured Nickel Phosphide as an Electrocatalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 9267–9270. [Google Scholar] [CrossRef] [PubMed]
  92. Kibler, L.A.; El-Aziz, A.M.; Hoyer, R.; Kolb, D.M. Tuning Reaction Rates by Lateral Strain in a Palladium Monolayer. Angew. Chem. Int. Ed. 2005, 44, 2080–2084. [Google Scholar] [CrossRef] [PubMed]
  93. McCrory, C.C.L.; Jung, S.; Peters, J.C.; Jaramillo, T.F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 16977–16987. [Google Scholar] [CrossRef] [PubMed]
  94. Stoerzinger, K.A.; Qiao, L.; Biegalski, M.D.; Shao-Horn, Y. Orientation-Dependent Oxygen Evolution Activities of Rutile IrO2 and RuO2. J. Phys. Chem. Lett. 2014, 5, 1636–1641. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, B.; Zheng, X.; Voznyy, O.; Comin, R.; Bajdich, M.; García-Melchor, M.; Han, L.; Xu, J.; Liu, M.; Zheng, L.; et al. Homogeneously Dispersed Multimetal Oxygen-Evolving Catalysts. Science 2016, 352, 333–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Gasteiger, H.A.; Kocha, S.S.; Sompalli, B.; Wagner, F.T. Activity Benchmarks and Requirements for Pt, Pt-Alloy, and Non-Pt Oxygen Reduction Catalysts for PEMFCs. Appl. Catal. B 2005, 56, 9–35. [Google Scholar] [CrossRef]
  97. Gasteiger, H.A.; Marković, N.M. Just a Dream−or Future Reality? Science 2009, 324, 48–49. [Google Scholar] [CrossRef]
  98. Siahrostami, S.; Verdaguer-Casadevall, A.; Karamad, M.; Deiana, D.; Malacrida, P.; Wickman, B.; Escudero-Escribano, M.; Paoli, E.A.; Frydendal, R.; Hansen, T.W.; et al. Enabling Direct H2O2 Production through Rational Electrocatalyst Design. Nat. Mater. 2013, 12, 1137–1143. [Google Scholar] [CrossRef] [Green Version]
  99. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  100. Huang, X.; Zhao, Z.; Cao, L.; Chen, Y.; Zhu, E.; Lin, Z.; Li, M.; Yan, A.; Zettl, A.; Wang, Y.M.; et al. High-Performance Transition Metal–Doped Pt3Ni Octahedra for Oxygen Reduction Reaction. Science 2015, 348, 1230–1234. [Google Scholar] [CrossRef] [Green Version]
  101. Peng, Z.; Yang, H. Synthesis and Oxygen Reduction Electrocatalytic Property of Pt-on-Pd Bimetallic Heteronanostructures. J. Am. Chem. Soc. 2009, 131, 7542–7543. [Google Scholar] [CrossRef] [PubMed]
  102. Kwak, K.; Choi, W.; Tang, Q.; Kim, M.; Lee, Y.; Jiang, D.-e.; Lee, D. A Molecule-Like PtAu24(SC6H13)18 Nanocluster as an Electrocatalyst for Hydrogen Production. Nat. Commun. 2017, 8, 14723. [Google Scholar] [CrossRef] [PubMed]
  103. Choi, W.; Hu, G.; Kwak, K.; Kim, M.; Jiang, D.-e.; Choi, J.-P.; Lee, D. Effects of Metal-Doping on Hydrogen Evolution Reaction Catalyzed by MAu24 and M2Au36 Nanoclusters (M = Pt, Pd). ACS Appl. Mater. Inter. 2018, 10, 44645–44653. [Google Scholar] [CrossRef] [PubMed]
  104. Kwak, K.; Lee, D. Electrochemistry of Atomically Precise Metal Nanoclusters. Acc. Chem. Res. 2019, 52, 12–22. [Google Scholar] [CrossRef] [PubMed]
  105. Hu, G.; Tang, Q.; Lee, D.; Wu, Z.; Jiang, D.-e. Metallic Hydrogen in Atomically Precise Gold Nanoclusters. Chem. Mater. 2017, 29, 4840–4847. [Google Scholar] [CrossRef]
  106. Du, Y.; Xiang, J.; Ni, K.; Yun, Y.; Sun, G.; Yuan, X.; Sheng, H.; Zhu, Y.; Zhu, M. Design of Atomically Precise Au2Pd6 Nanoclusters for Boosting Electrocatalytic Hydrogen Evolution on MoS2. Inorg. Chem. Front. 2018, 5, 2948–2954. [Google Scholar] [CrossRef]
  107. Eguchi, D.; Sakamoto, M.; Teranishi, T. Ligand Effect on the Catalytic Activity of Porphyrin-Protected Gold Clusters in the Electrochemical Hydrogen Evolution Reaction. Chem. Sci. 2018, 9, 261–265. [Google Scholar] [CrossRef] [Green Version]
  108. Zhao, S.; Jin, R.; Song, Y.; Zhang, H.; House, S.D.; Yang, J.C.; Jin, R. Atomically Precise Gold Nanoclusters Accelerate Hydrogen Evolution over MoS2 Nanosheets: The Dual Interfacial Effect. Small 2017, 13, 1701519. [Google Scholar] [CrossRef]
  109. Kwak, K.; Choi, W.; Tang, Q.; Jiang, D.-e.; Lee, D. Rationally Designed Metal Nanocluster for Electrocatalytic Hydrogen Production from Water. J. Mater. Chem. A 2018, 6, 19495–19501. [Google Scholar] [CrossRef]
  110. Zhao, S.; Jin, R.; Abroshan, H.; Zeng, C.; Zhang, H.; House, S.D.; Gottlieb, E.; Kim, H.J.; Yang, J.C.; Jin, R. Gold Nanoclusters Promote Electrocatalytic Water Oxidation at the Nanocluster/CoSe2 Interface. J. Am. Chem. Soc. 2017, 139, 1077–1080. [Google Scholar] [CrossRef]
  111. Chen, W.; Chen, S. Oxygen Electroreduction Catalyzed by Gold Nanoclusters: Strong Core Size Effects. Angew. Chem. Int. Ed. 2009, 48, 4386–4389. [Google Scholar] [CrossRef] [PubMed]
  112. Wang, L.; Tang, Z.; Yan, W.; Yang, H.; Wang, Q.; Chen, S. Porous Carbon-Supported Gold Nanoparticles for Oxygen Reduction Reaction: Effects of Nanoparticle Size. ACS Appl. Mater. Inter. 2016, 8, 20635–20641. [Google Scholar] [CrossRef] [PubMed]
  113. Sumner, L.; Sakthivel, N.A.; Schrock, H.; Artyushkova, K.; Dass, A.; Chakraborty, S. Electrocatalytic Oxygen Reduction Activities of Thiol-Protected Nanomolecules Ranging in Size from Au28(SR)20 to Au279(SR)84. J. Phys. Chem. C 2018, 122, 24809–24817. [Google Scholar] [CrossRef]
  114. Jones, T.C.; Sumner, L.; Ramakrishna, G.; Hatshan, M.b.; Abuhagr, A.; Chakraborty, S.; Dass, A. Bulky t-Butyl Thiolated Gold Nanomolecular Series: Synthesis, Characterization, Optical Properties, and Electrocatalysis. J. Phys. Chem. C 2018, 122, 17726–17737. [Google Scholar] [CrossRef]
  115. Lu, Y.; Jiang, Y.; Gao, X.; Chen, W. Charge State-Dependent Catalytic Activity of [Au25(SC12H25)18] Nanoclusters for the Two-Electron Reduction of Dioxygen to Hydrogen Peroxide. Chem. Commun. 2014, 50, 8464–8467. [Google Scholar] [CrossRef]
  116. Kwak, K.; Azad, U.P.; Choi, W.; Pyo, K.; Jang, M.; Lee, D. Efficient Oxygen Reduction Electrocatalysts Based on Gold Nanocluster–Graphene Composites. ChemElectroChem 2016, 3, 1253–1260. [Google Scholar] [CrossRef]
  117. Skúlason, E.; Tripkovic, V.; Björketun, M.E.; Gudmundsdóttir, S.; Karlberg, G.; Rossmeisl, J.; Bligaard, T.; Jónsson, H.; Nørskov, J.K. Modeling the Electrochemical Hydrogen Oxidation and Evolution Reactions on the Basis of Density Functional Theory Calculations. J. Phys. Chem. C 2010, 114, 18182–18197. [Google Scholar] [CrossRef]
  118. Sakamoto, M.; Tanaka, D.; Tsunoyama, H.; Tsukuda, T.; Minagawa, Y.; Majima, Y.; Teranishi, T. Platonic Hexahedron Composed of Six Organic Faces with an Inscribed Au Cluster. J. Am. Chem. Soc. 2012, 134, 816–819. [Google Scholar] [CrossRef]
  119. Tanaka, D.; Inuta, Y.; Sakamoto, M.; Furube, A.; Haruta, M.; So, Y.-G.; Kimoto, K.; Hamada, I.; Teranishi, T. Strongest Π–Metal Orbital Coupling in a Porphyrin/Gold Cluster System. Chem. Sci. 2014, 5, 2007–2010. [Google Scholar] [CrossRef]
  120. Li, P.; Wang, M.; Duan, X.; Zheng, L.; Cheng, X.; Zhang, Y.; Kuang, Y.; Li, Y.; Ma, Q.; Feng, Z.; et al. Boosting Oxygen Evolution of Single-Atomic Ruthenium through Electronic Coupling with Cobalt-Iron Layered Double Hydroxides. Nat. Commun. 2019, 10, 1711. [Google Scholar] [CrossRef] [Green Version]
  121. Lee, Y.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao-Horn, Y. Synthesis and Activities of Rutile IrO2 and RuO2 Nanoparticles for Oxygen Evolution in Acid and Alkaline Solutions. J. Phys. Chem. Lett. 2012, 3, 399–404. [Google Scholar] [CrossRef] [PubMed]
  122. Mattioli, G.; Giannozzi, P.; Amore Bonapasta, A.; Guidoni, L. Reaction Pathways for Oxygen Evolution Promoted by Cobalt Catalyst. J. Am. Chem. Soc. 2013, 135, 15353–15363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Frydendal, R.; Paoli, E.A.; Knudsen, B.P.; Wickman, B.; Malacrida, P.; Stephens, I.E.L.; Chorkendorff, I. Benchmarking the Stability of Oxygen Evolution Reaction Catalysts: The Importance of Monitoring Mass Losses. ChemElectroChem 2014, 1, 2075–2081. [Google Scholar] [CrossRef] [Green Version]
  124. Zhuang, Z.; Sheng, W.; Yan, Y. Synthesis of Monodispere Au@Co3O4 Core-Shell Nanocrystals and Their Enhanced Catalytic Activity for Oxygen Evolution Reaction. Adv. Mater. 2014, 26, 3950–3955. [Google Scholar] [CrossRef]
  125. Zhao, X.; Gao, P.; Yan, Y.; Li, X.; Xing, Y.; Li, H.; Peng, Z.; Yang, J.; Zeng, J. Gold Atom-Decorated CoSe2 Nanobelts with Engineered Active Sites for Enhanced Oxygen Evolution. J. Mater. Chem. A 2017, 5, 20202–20207. [Google Scholar] [CrossRef]
  126. Li, Z.-y.; Ye, K.-h.; Zhong, Q.-s.; Zhang, C.-j.; Shi, S.-t.; Xu, C.-w. Au–Co3O4/C as an Efficient Electrocatalyst for the Oxygen Evolution Reaction. ChemPlusChem 2014, 79, 1569–1572. [Google Scholar] [CrossRef]
  127. Mills, G.; Gordon, M.S.; Metiu, H. Oxygen Adsorption on Au Clusters and a Rough Au(111) Surface: The Role of Surface Flatness, Electron Confinement, Excess Electrons, and Band Gap. J. Chem. Phys. 2003, 118, 4198–4205. [Google Scholar] [CrossRef] [Green Version]
  128. Okumura, M.; Kitagawa, Y.; Kawakami, T.; Haruta, M. Theoretical Investigation of the Hetero-Junction Effect in PVP-Stabilized Au13 Clusters. The Role of PVP in Their Catalytic Activities. Chem. Phys. Lett. 2008, 459, 133–136. [Google Scholar] [CrossRef]
  129. Yin, H.; Tang, H.; Wang, D.; Gao, Y.; Tang, Z. Facile Synthesis of Surfactant-Free Au Cluster/Graphene Hybrids for High-Performance Oxygen Reduction Reaction. ACS Nano 2012, 6, 8288–8297. [Google Scholar] [CrossRef]
Figure 1. (A) Schematic illustration of gold nanoclusters (Aun NCs) for an electrocatalytic reaction in water splitting (hydrogen evolution reaction (HER) and oxygen evolution reaction (OER)) and fuel cells (oxygen reduction reaction (ORR)). (B) Current–potential characteristics for (a) HER, (b) OER, and (c) ORR.
Figure 1. (A) Schematic illustration of gold nanoclusters (Aun NCs) for an electrocatalytic reaction in water splitting (hydrogen evolution reaction (HER) and oxygen evolution reaction (OER)) and fuel cells (oxygen reduction reaction (ORR)). (B) Current–potential characteristics for (a) HER, (b) OER, and (c) ORR.
Nanomaterials 10 00238 g001
Figure 2. (A) HER polarization curves of Au25(SC6H13)18- or Au24Pt(SC6H13)18-adsorbed glassy carbon electrode (GCE), or GCE. (B) H2 production rates per mass of metals in the catalyst of Au24Pt(SC6H13)18/C (blue circles) and Pt/C (black triangles) electrodes. (C) DFT calculation results for Au24Pt(SCH3)18. Color code: golden = Au core; olive = Au shell; purple = Pt; green = adsorbed H from the liquid medium; grey = S. Panels (AC) are reproduced with permission from reference [102]. Copyright Springer Nature, 2017.
Figure 2. (A) HER polarization curves of Au25(SC6H13)18- or Au24Pt(SC6H13)18-adsorbed glassy carbon electrode (GCE), or GCE. (B) H2 production rates per mass of metals in the catalyst of Au24Pt(SC6H13)18/C (blue circles) and Pt/C (black triangles) electrodes. (C) DFT calculation results for Au24Pt(SCH3)18. Color code: golden = Au core; olive = Au shell; purple = Pt; green = adsorbed H from the liquid medium; grey = S. Panels (AC) are reproduced with permission from reference [102]. Copyright Springer Nature, 2017.
Nanomaterials 10 00238 g002
Figure 3. (A,C,E) Schematic illustration of coordination of ligands: (A) porphyrin SC1P, (C) porphyrin SC2P, and (E) PET. (B,D,F) TEM images of Au NCs with a core size of approximately 1.3 nm protected by porphyrin SC1P, porphyrin SC2P, or PET, respectively. (G) Comparison of overpotential at −10 mA cm−2 and (H) current density at −0.4 V of each size of Au NCs protected with each ligand. Panels (AH) are reproduced with permission from reference [107]. Copyright Royal Society of Chemistry, 2018.
Figure 3. (A,C,E) Schematic illustration of coordination of ligands: (A) porphyrin SC1P, (C) porphyrin SC2P, and (E) PET. (B,D,F) TEM images of Au NCs with a core size of approximately 1.3 nm protected by porphyrin SC1P, porphyrin SC2P, or PET, respectively. (G) Comparison of overpotential at −10 mA cm−2 and (H) current density at −0.4 V of each size of Au NCs protected with each ligand. Panels (AH) are reproduced with permission from reference [107]. Copyright Royal Society of Chemistry, 2018.
Nanomaterials 10 00238 g003
Figure 4. (A) Schematic illustration of proton relay mechanism of Au24Pt(SR)18 nanocluster for formation of H2 and (B) ligand structures: SC6H13, MPA, and MPS. Color codes: blue = Pt; golden = core Au; red = shell Au; and green = S. (C) HER polarization curves in 0.1 M KCl aqueous solution containing 180 mM acetic acid for MPA-Au25 (red) or MPS-Au25 (blue). (D) turnover frequencies (TOFs) obtained at various potentials in water (3.0 M KCl) containing 180 mM HOAc for MPA-Au25 (red), MPS-Au25 (blue), or MPS-Au24Pt (green). Panels (AD) are reproduced with permission from reference [109]. Copyright Royal Society of Chemistry, 2018.
Figure 4. (A) Schematic illustration of proton relay mechanism of Au24Pt(SR)18 nanocluster for formation of H2 and (B) ligand structures: SC6H13, MPA, and MPS. Color codes: blue = Pt; golden = core Au; red = shell Au; and green = S. (C) HER polarization curves in 0.1 M KCl aqueous solution containing 180 mM acetic acid for MPA-Au25 (red) or MPS-Au25 (blue). (D) turnover frequencies (TOFs) obtained at various potentials in water (3.0 M KCl) containing 180 mM HOAc for MPA-Au25 (red), MPS-Au25 (blue), or MPS-Au24Pt (green). Panels (AD) are reproduced with permission from reference [109]. Copyright Royal Society of Chemistry, 2018.
Nanomaterials 10 00238 g004
Figure 5. (A) High-angle annular dark-field scanning TEM (HAADF-STEM) images, (B) HER polarization curves, and (C) Mo 3d X-ray photoelectron spectroscopy (XPS) spectra of Au25(PET)18/MoS2. (D) HER polarization curves of Au25(SePh)18/MoS2. Panels (AD) are reproduced with permission from reference [108]. Copyright Wiley-VCH, 2017.
Figure 5. (A) High-angle annular dark-field scanning TEM (HAADF-STEM) images, (B) HER polarization curves, and (C) Mo 3d X-ray photoelectron spectroscopy (XPS) spectra of Au25(PET)18/MoS2. (D) HER polarization curves of Au25(SePh)18/MoS2. Panels (AD) are reproduced with permission from reference [108]. Copyright Wiley-VCH, 2017.
Nanomaterials 10 00238 g005
Figure 6. (A,B) HAADF-STEM images of Au25(PET)18/CoSe2 composite at different magnifications. (C,D) OER polarization curves of CoSe2, Au10(SPh-tBu)10/CoSe2, Au25(PET)18/CoSe2, Au144(PET)60/CoSe2, Au333(PET)79/CoSe2, and PtNP/CB (CB = carbon black). (E) Co 2p XPS spectra and (F) Raman spectra of CoSe2 and Au25(PET)18/CoSe2 composite. Panels (AF) are reproduced with permission from reference [110]. Copyright American Chemical Society, 2017.
Figure 6. (A,B) HAADF-STEM images of Au25(PET)18/CoSe2 composite at different magnifications. (C,D) OER polarization curves of CoSe2, Au10(SPh-tBu)10/CoSe2, Au25(PET)18/CoSe2, Au144(PET)60/CoSe2, Au333(PET)79/CoSe2, and PtNP/CB (CB = carbon black). (E) Co 2p XPS spectra and (F) Raman spectra of CoSe2 and Au25(PET)18/CoSe2 composite. Panels (AF) are reproduced with permission from reference [110]. Copyright American Chemical Society, 2017.
Nanomaterials 10 00238 g006
Figure 7. (A) Cyclic voltammograms of Aun(SR)m/GCE (n = 11, 25, 55, and 140) saturated with O2 and Au11(PPh3)8Cl3/GCE saturated with N2 (thin solid curve). (B) Current density and overpotential of ORR activity with each size of Aun NCs. (C) Koutecky–Levich plots at different applied potentials of a GCE modified with Au11(PPh3)8Cl3. (D) Rotating-disk voltammograms (rotation rate: 3600 rpm) of various Aun(SR)m/GCE (n = 11, 25, 55, and 140). Panels (AD) are reproduced with permission from reference [111]. Copyright Wiley-VCH, 2009.
Figure 7. (A) Cyclic voltammograms of Aun(SR)m/GCE (n = 11, 25, 55, and 140) saturated with O2 and Au11(PPh3)8Cl3/GCE saturated with N2 (thin solid curve). (B) Current density and overpotential of ORR activity with each size of Aun NCs. (C) Koutecky–Levich plots at different applied potentials of a GCE modified with Au11(PPh3)8Cl3. (D) Rotating-disk voltammograms (rotation rate: 3600 rpm) of various Aun(SR)m/GCE (n = 11, 25, 55, and 140). Panels (AD) are reproduced with permission from reference [111]. Copyright Wiley-VCH, 2009.
Nanomaterials 10 00238 g007
Figure 8. (A) X-ray crystal structures of Aun(TBBT)m NCs (n = 28, 36, 133, and 279). (B) Rotating-disk voltammograms recorded for the ORR activity of Au36(TBBT)24/GCE at different rotation rates. (C) Reaction rate constant ln(k) vs. overpotential E plots with each size of Aun(TBBT)m (n = 28, 36, 133, and 279). Panels (AC) are reproduced with permission from reference [113]. Copyright American Chemical Society, 2018.
Figure 8. (A) X-ray crystal structures of Aun(TBBT)m NCs (n = 28, 36, 133, and 279). (B) Rotating-disk voltammograms recorded for the ORR activity of Au36(TBBT)24/GCE at different rotation rates. (C) Reaction rate constant ln(k) vs. overpotential E plots with each size of Aun(TBBT)m (n = 28, 36, 133, and 279). Panels (AC) are reproduced with permission from reference [113]. Copyright American Chemical Society, 2018.
Nanomaterials 10 00238 g008
Figure 9. (A) Cyclic voltammograms, (B) electron transfer number (n), and (C) percentage of H2O2 of the ORR on Au25(SC12H25)18 with different charge states ([Au25(SC12H25)18], [Au25(SC12H25)18]0, and [Au25(SC12H25)18]+) in 0.1 M KOH aq saturated with O2. (D) Accelerated durability tests of [Au25(SC12H25)18] performed for 1000 cycles. Panels (AD) are reproduced with permission from reference [115]. Copyright Royal Society of Chemistry, 2014.
Figure 9. (A) Cyclic voltammograms, (B) electron transfer number (n), and (C) percentage of H2O2 of the ORR on Au25(SC12H25)18 with different charge states ([Au25(SC12H25)18], [Au25(SC12H25)18]0, and [Au25(SC12H25)18]+) in 0.1 M KOH aq saturated with O2. (D) Accelerated durability tests of [Au25(SC12H25)18] performed for 1000 cycles. Panels (AD) are reproduced with permission from reference [115]. Copyright Royal Society of Chemistry, 2014.
Nanomaterials 10 00238 g009
Table 1. Representative references on HER activity of Aun NCs and related alloy NCs.
Table 1. Representative references on HER activity of Aun NCs and related alloy NCs.
LigandSupportExperimental ConditionActivityReference
SC6H131.0 M TFA and 0.1 M Bu4NPF6 in THF cAu24Pt(SC6H13)18 > Au25(SC6H13)18[102]
SC6H13carbon black1 M Britton–Robinson buffer solution in 2 M KCl aq (pH 3) c,dAu24Pt(SC6H13)18 > Au24Pd(SC6H13)18 > Au25(SC6H13)18[103]
SC6H13carbon black1 M Britton–Robinson buffer solution in 2 M KCl aq (pH 3) c,dAu36Pt2(SC6H13)24 > Au36Pd2(SC6H13)24 > Au38(SC6H13)24[103]
PPh3
PPh2 a
Cl b
PhF2S
MoS20.5 M phosphate buffer solution (pH 6.7) c,dAu2Pd6(S4(PPh3)4(PhF2S)6)/MoS2 > Mixture of Au2Cl2C(PPh2)2 and Pd3(Cl(PPh2)2(PPh3)3)/MoS2 > Pd3(Cl(PPh2)2(PPh3)3)/MoS2 > Au2Cl2C(PPh2)2/MoS2 > MoS2[104]
porphyrin SC1P
porphyrin SC2P
PET
0.5 M H2SO4 aq eAu(1.3 nm)(porphyrin SC1P) > Au(1.3 nm)(porphyrin SC2P) > Au(1.3 nm)(PET)[107]
PET
SePh
MoS20.5 M H2SO4 aq c,dAu25(PET)18/MoS2 > Au25(SePh)18/MoS2 > MoS2[108]
SC6H13
MPA
MPS
0.1 M KCl aq cAu24Pt(MPS)18 > Au25(MPS)18 > Au25(MPA)18 > Au25(SC6H13)18[109]
a Diphenylphosphine. b Chlorine. c WE: Working electrode; GCE. d WE: Containing Nafion. e WE: Carbon tape.
Table 2. Representative reference on OER activity of Aun(SR)m NCs.
Table 2. Representative reference on OER activity of Aun(SR)m NCs.
LigandSupportExperimental ConditionActivityReference
PETCoSe20.l M KOH aq a, bAu25(PET)18/CoSe2 > CoSe2[110]
a WE: GCE. b WE: Containing Nafion.
Table 3. Representative references on ORR activity of Aun NCs.
Table 3. Representative references on ORR activity of Aun NCs.
LigandSupportExperimental ConditionActivityReference
PET
SC6H13
Cl
PPh3
0.1 M KOH aq aAu11(PPh3)8Cl3 > Au25(PET)18 > Au55(PPh3)12Cl6 > Au140(SC6H13)53[101]
PETReduced graphene oxide0.1 M KOH aq a, bAu25(PET)18 > Au38(PET)24 > Au144(PET)60[112]
TBBTSWNTs0.1 M KOH aq a, bAu36(TBBT)24 > Au133(TBBT)52 > Au279(TBBT)84 > Au28(TBBT)20[113]
S-tBuSWNTs0.1 M KOH aq a, bAu65(S-tBu)29 > Au46(S-tBu)24 > Au30(S-tBu)18 > Au23(S-tBu)16[114]
SC12H250.1 M KOH aq a, b[Au25(SC12H25)18] > [Au25(SC12H25)18]0 > [Au25(SC12H25)18]+ c[115]
a WE: GCE. b WE; Containing Nafion. c Tow-electron reduction.

Share and Cite

MDPI and ACS Style

Kawawaki, T.; Negishi, Y. Gold Nanoclusters as Electrocatalysts for Energy Conversion. Nanomaterials 2020, 10, 238. https://doi.org/10.3390/nano10020238

AMA Style

Kawawaki T, Negishi Y. Gold Nanoclusters as Electrocatalysts for Energy Conversion. Nanomaterials. 2020; 10(2):238. https://doi.org/10.3390/nano10020238

Chicago/Turabian Style

Kawawaki, Tokuhisa, and Yuichi Negishi. 2020. "Gold Nanoclusters as Electrocatalysts for Energy Conversion" Nanomaterials 10, no. 2: 238. https://doi.org/10.3390/nano10020238

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