Yuanxin Du*,
Yi Fang,
Pei Wang and
Manzhou Zhu
*
Department of Materials Science and Engineering, Centre for Atomic Engineering of Advanced Materials, Key Laboratory of Structure and Functional Regulation of Hybrid Materials of Ministry of Education, Key Laboratory of Functional Inorganic Material Chemistry of Anhui Province, Anhui University, Hefei 230601, China. E-mail: duyuanxin@ahu.edu.cn; zmz@ahu.edu.cn
First published on 30th April 2025
Utilizing renewable sources to convert small-molecule energy carriers (such as nitrogen, carbon dioxide, water, or oxygen) into high value-added chemicals and fuels is of great significance. Rational design of the catalyst is the key to achieving efficient catalytic performance. Atomically precise metal nanoclusters (NCs) exhibit the advantages of high atomic economy, distinctive discrete electronic energy, and homogeneity in size, composition, structure, and surface environment, not only offering extraordinary catalytic activity but also providing the opportunity to reveal the reaction mechanism. In the metal NC family, Au-based NCs have attracted widespread and sustained interest due to their simple preparation, high stability, easy functionalization, and especially their unique catalytic activity, which once provoked a “gold rush” in academia. The synergistic effect between different metal atoms is regarded as an effective strategy to achieve enhanced catalytic performance, but the underlying mechanism is a puzzle. Recently, abundant, diverse and adjustable atomically precise Au-based bimetallic NCs (doped with Ag, Cu, Pt, Pd, Cd, Hg, Ir etc.) have emerged, which not only provide a bank of materials for highly active catalysts, but also provide feasibility for revealing synergistic effects at the atomic level. This perspective briefly introduces the common synthesis strategy and structural characteristics of atomically precise Au-based bimetallic NCs, summarizes recent advances in their synergistic catalysis in energy-related small-molecule conversion, and proposes insights and advice for future breakthroughs in this field.
In order to improve the reactivity per unit volume of the catalyst, there are two main strategies: one is to increase the number of active sites and the other is to improve the reaction rate. Catalyst miniaturization is an effective method to realize the former strategy.25–29 In recent years, with progress in nanomaterial preparation methods and the rapid development of fine structure characterization techniques, scientists have been able to achieve the controllable preparation of nanoparticles (NPs), from tens of nanometers to precise nanoclusters (NCs) composed of a few to several hundred atoms, and then to single atom catalysts (SACs), maximizing the utilization rate of atoms and achieving atomic economy.30–32 The latter is strongly influenced by the adsorption, activation and desorption behaviours of reactants, intermediates and products on the catalyst surface. By using atom doping to exert a synergistic effect between the components, these behaviours can be regulated and optimized to obtain an appropriate and matchable interaction between the reaction molecules and the catalyst, and then realizing an improvement in the catalytic reaction rate.33–35 Metal NCs, as an emerging type of nanocatalyst, bridging the gap between metal NPs and SACs, has attracted extensive interest due to the characteristic of having a unique molecule-like discrete electronic energy different from those of metal NPs or SACs.36–39 In addition, compared to its large counterpart metal NPs, metal NCs have relatively small size with higher atomic utilization and abundant surface unsaturated coordination active sites.40–42 In comparison to SACs, metal NCs exhibit the possibility of a synergistic effect.43,44 Furthermore, metal NCs show great adjustability in terms of catalytic properties, because at the sub-nanometer scale, even a single atom change will have a huge impact on the properties of the material.45–47
In recent years, research on metal NC catalysis has undergone explosive growth, especially, those metal NCs with precise composition and structure, due to the potential opportunity to investigate the catalytic mechanism based on a clear structure–activity relationship.32,40,48,49 Au has for a long time been considered an unreactive noble metal. In the 1980s, Haruta and Hutching et al. discovered that small Au NPs exhibit extraordinary catalytic activity in CO oxidation and acetylene hydrochlorination, which has caused a boom in research into Au catalysis.50,51 Within the large family of precise NCs, atomically precise Au-based NCs show high stability and ready solubility in various solvents, which is very important for practical catalytic application.52,53 In addition, the advantages of easy preparation, high yield, and versatile surface functionalization are also the foundation of Au-based NCs as excellent catalysts.54–63 For example, the typical representative, Au25 series NCs, can be seen as the standard-bearer of the NC family due to their early discovery (successfully synthesized in 1998 and structurally determined in 2008), and in-depth investigation (properties of optics, chirality, magnetism, electrochemistry, etc.), and wide application (catalysis, chemical sensing, imaging, bio-labeling, etc.).64,65 Furthermore, a variety of derivatives with abundant diversity in terms of properties can be obtained by metal doping, changing ligands, regulating charge states, etc.64–66
Among various adjustment strategies, heteroatom doping is an effective way to significantly expand the diversity of the composition and structure of NCs and provide a potential synergistic effect for improving catalytic performance. Speaking of a synergistic effect, it is a widely known concept among the public. Simply put, in the field of catalysis, a synergistic effect refers to the catalytic performance produced by the combination of multiple components surpassing that of a single component. It is mainly manifested through synergistic effects between metal atoms, between metals and supports, and between metal atoms and ligands. Here, we focus on the synergistic effects between metal atoms, reflected in the interactions between different metal active centers. Taking the ternary catalyst PtPdRh in automobile exhaust treatment as an example, Pt is mainly responsible for catalyzing the oxidation reaction of CO and CHx, converting them into non-toxic CO2 and H2O. Pd also participates in this process, mainly playing a role in heat resistance and improving stability. Rh is mainly responsible for catalyzing the reduction reaction of NOx, converting them into N2 and O2. The synergistic effect among the three achieves efficient catalytic activity and stability.
More importantly, the selective and purposeful introduction of foreign atoms into a monometallic NC can rationally optimize catalytic performance via modulation of electronic structure, additional active sites, and improvement in structural stability.43,67,68 Many researches have reported that Au NCs can be partially substituted with Ag, Cu, Pt, Pd, Cd, Hg, or Ir, and the Au-based alloy NCs perform with superior catalytic activity to their parent Au NCs in numerous reactions due to the synergistic effect.69 Motivated by the current era of rapid development of NC materials and energy catalysis, herein, recent advances in atomically precise Au-based bimetallic NCs in energy-related small-molecule catalytic conversion are reviewed, including a summary of common synthesis strategies and structural characteristic of Au-based bimetallic NCs, and a discussion of the synergistic effect on catalytic activity, selectivity, and stability in HER, OER, ORR, CO2RR, NRR, NO3RR, etc.
Direct co-reduction strategy: It usually takes a metal–ligand complex (i.e. Au–SR, Ag–SR, etc.) as a precursor and directly synthesizes bimetallic NCs with a reducing agent (i.e. NaBH4, CO, etc.); it is also called “in situ” or “one-pot” synthesis.48,69 For example, Au25−xAgx(SR)18 NCs with continuously modulated x have been successfully synthesized by this method.72
Cation-assisted strategy: this refers to the synthesis of bimetallic NCs by the reaction of the existing parent NCs with another metal cation.69 For instance, Murray et al. obtained Au25−xAgx(SR)18 NCs by utilizing Ag+ ions reacting with Au25(SR)18 NCs.73
Anti-galvanic reduction (AGR) strategy: Wu et al. proposed that the redox potential of the order of metals is no longer the only decisive thermodynamic parameter when the size of NCs is less than 3 nm.74 They prepared Au–Ag and Au–Cu bimetallic NCs by doping heteroatoms in Au25(SR)18 on the basis of the AGR method.75,76
Metal-exchange strategy: instead of using an inorganic metal salt, an organic metal–ligand complex is used in the synthesis of bimetallic NCs, and this synthesis process can no longer relate to the rule of metal redox potential.77 For example, our group utilized Au25(SR)18 as a template to synthesize Cd1Au24(SR)18 and Hg1Au24(SR)18 by this method.78
Intercluster reaction strategy: this refers to the preparation of new bimetallic NCs via the reaction between two stable NCs.48,69 For example, (Au25−xAgx)(SR)18 and AuxAg44−x(SR)30 NCs are obtained by spontaneously exchanging metal atoms after mixing Au25(SR)18 and Ag44(SR)30.79
For bimetallic NCs formed by heteroatom doping, the most important considerations are the type, number and position of the doping atoms, because these structural parameters have a great influence on the catalytic performance. Considering that NCs can be viewed as being composed of a metal kernel and a metal–ligand motif, the doping position of the heteroatom can be divided into center doping, kernel doping, kernel surface doping, and motif doping (Fig. 1). For motif doping, the heteroatom is connected with the coordination atom of the ligand, such as S, P, or N atoms. It is located on the outside of the cluster and has the opportunity to make direct contact with the reactant. For kernel doping (including center doping and kernel surface doping simultaneously) and kernel surface doping, the heteroatom is located at the surface of the metal core. When the catalytic reaction occurs, some ligands will detach due to the weak connection between metal atoms and ligands, and the atoms at the surface of the metal core are exposed as active centers to directly participate in the reaction. For center doping, the heteroatom is located at the center of the metal core, and it seems to have no influence on catalysis due to the indirect contact with the reactant. However, it can participate in the catalytic reaction by modulating the electronic structure of the metal core, regulating the interaction strength with the reactant or intermediate, and optimizing the reaction pathway and kinetics. The synthesis methods and characteristic structural parameters of Au-based bimetallic NCs reported in recent years are summarized in Appendix Table 1.
![]() | ||
Fig. 1 Schematic diagram of the doping position of the heteroatom in Au-based bimetallic NCs. Atom colors: Au = orange, heteroatom = blue, S = red. |
At present, Pt is recognized as the best-known efficient HER electrocatalyst; however, its high cost and scarce stocks hamper its large-scale industrial application.84 As the sixth period element in the periodic table next to Pt, the electronic structure of Au is similar to that of Pt, so there is reason to believe that Au NCs have the potential to be an outstanding HER catalyst.85,86 Kwak et al. selectively replaced the center Au atom in Au25 NCs with a single Pt atom and obtained Pt1Au24 bimetallic NCs (Fig. 2a).87 Pt doping has little effect on the geometric structure of the NCs; Pt1Au24 almost maintains the configuration of the original Au25. However, the electronic structure changes a lot, which is reflected in the surface charge state ([Pt1Au24]0 & [Au25]−), superatomic electronic configuration (Au25: 8-electron, Pt1Au24: 6-electron)88–90 and optical absorbance spectra (Au25:
1.8 eV, Pt1Au24
:
1.1 and 2.1 eV). The redox behavior is also drastically altered. As measured by square-wave voltammetry (SWV), the gaps between the first oxidation (O1) and reduction (R1) potential for Au25 and Pt1Au24 are 1.67 and 0.73 V, respectively (Fig. 2b). The HOMO–LUMO (highest occupied-lowest unoccupied molecular orbitals) gaps are 1.32 and 0.29 V for Au25 and Pt1Au24, respectively. It is worth noting that the reduction potential of Pt1Au24 is nearly 1 V more positive than that of Au25, indicating the possibility of a lower overpotential for the electrocatalytic reduction reaction. This conjecture is further confirmed by the result of linear sweep voltammograms (LSVs). In THF (0.1 M Bu4NPF6) solution containing 1.0 M trifluoroacetic acid (TFA), Au25 and Pt1Au24 show onset potentials at −1.10 and −0.89 V, respectively (Fig. 2c). Notably, the overpotential for Pt1Au24 is ∼70 mV (relative to −0.82 V, the thermodynamic reduction potential of a proton in THF with 1.0 M TFA), which is superior to other natural hydrogenase enzymes (∼100 mV).91–93 With an increase in TFA concentration, the current at −0.76 V (the first reduction [Pt1Au24]0/−) of Pt1Au24 shows no significant change, while the current at −1.10 V (the second reduction [Pt1Au24]−/2−) is significantly enhanced, suggesting that Pt1Au24 acts as an electron transfer mediator for HER (Fig. 2d).94,95 The kobs (pseudo-first-order rate constant) of Pt1Au24 is calculated to be 121000 s−1 at η = 650 mV, which is higher than that of Au25 (8000 s−1) or other molecule-like complexes (Co-complex: 700 s−1 at η = 890 mV, Cu-complex: 11000 s−1 at η = 720 mV, Ni-complex: 106000 s−1 at η = 650 mV) (Fig. 2e).96–98
![]() | ||
Fig. 2 (a) Structure of Au25 and Pt1Au24 NC. (b) SWVs of Au25 (red) and Pt1Au24 (blue) in CH2Cl2 (0.1 M Bu4NPF6). (c) LSVs in THF (0.1 M Bu4NPF6) solution containing 1.0 M TFA without a catalyst (black) and with Au25 (red) and Pt1Au24 (blue). (d) LSVs of Pt1Au24 in THF (0.1 M Bu4NPF6) with different concentrations of TFA. (e) Kobs–potential plots for Au25 (red) and Pt1Au24 (blue). (f) Reaction energy calculation for HER on Pt1Au24. Reproduced from ref. 87 with permission from Springer Nature, copyright 2017. |
In addition, Pt1Au24 shows charge-state-dependent catalytic activity, indicating molecule-like catalytic behavior different from that of the larger-sized metal NPs. The currents at potentials negative to the [Pt1Au24]−/2− all exhibit linear correlation with [Pt1Au24] and [TFA]1/2, suggesting a heterolytic HER mechanism, [H–Pt1Au24]− + H+ → [Pt1Au24]0 + H2. In contrast, the current at the potential at which [Pt1Au24]− exists as the predominant form shows linear correlation with [TFA] and [Pt1Au24]3/2, indicating a homolytic HER mechanism, [H–Pt1Au24]0 + [H–Pt1Au24]0 → 2[Pt1Au24]0 + H2. The heterolytic HER mechanism over Pt1Au24 is further confirmed by density functional theory (DFT) calculations with an energy change of −0.155 eV, while for Au25, the thermodynamically favorable HER pathway is the homolytic mechanism (Fig. 2f). Additionally, an H–Pt bond can spontaneously form. The bond length between adsorbed H and central Pt is shorter than that between adsorbed H and surface Au, indicating that the stronger H–Pt interaction is beneficial for HER energetics on [Pt1Au24]2−. As a result, Pt1Au24 exhibits excellent HER activity whether for homogeneous catalysis in non-aqueous solvent or heterogeneous catalysis in aqueous media, compared to Au25, other molecule-like complexes, or commercial Pt/C benchmarking.
Recently, Sun et al. revisited the electrocatalytic HER activity origin of Pt1Au24 NCs by a combination of advanced first-principles calculations and attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) experiments.99 They found that in addition to the central Pt atom, the exposed bridged Au site, due to spontaneous thiolate ligand desorption during the electrochemical process, is also a catalytically active site. The synergistic effect between the two types of active site contributes to the extraordinary HER activity of the Pt1Au24 cluster. As Pt centrally doped Pt1Au24 significantly improves HER activity, it is natural for researchers to ask whether Pd1Au24, which has a nearly isoelectronic structure and similar redox property, would also exhibit similarly high HER activity. Choi et al. systematically compared the HER catalytic activity parameters, such as onset potential (Eonset), current density, and turnover frequency (TOF) of M1Au24 and M2Au36 (M = Pt, Pd).100 Compared to non-doped Au NCs, atom doping not only brings about a change in the electrochemical redox property, but also causes an improvement in HER activity. Eonset is determined by the match between the reduction potential of NC and H+. In both M1Au24 and M2Au36 systems, Pt doping shows higher current density and TOF than Pd doping, which is attributed to the lower H adsorption free energy (ΔGH) in Pt-doped Au NCs (Fig. 3a). Based on the excellent HER performance of Pt1Au24, Negishi et al. further changed the surface ligand to, for example, C6 = 1-hexanethiolates, TBBT = 4-tert-butylbenzenethiolate, PDT = 1,3-propanedithiolate, or PET = 2-phenylethanethiolate, in pursuit of higher catalytic activity.101 Due to the differences in the lengths and orientations of -Au(I)-SR-Au(I)- staples between TDT/PDT- and C6/PET-protected NCs, the Au atoms of the Pt1Au12 metal core in TDT/PDT-protected Pt1Au24 NCs are exposed on the outside, resulting in enhanced HER activity. Although Ag is considered to be inactive for HER, obtaining AuAg alloy NCs with comparable HER activity is highly possible via a reasonable design. Li et al. synthesized Au36Ag2(SR)18 NCs, which can be regarded as having a trimeric structure, that is with three icosahedral (Ih) units face-fused together in a cyclic manner.102 The unique face-fusion mode in Au36Ag2 endows it with unfilled superatomic orbitals, a low ligand-to-metal ratio, and low-coordinated Au atoms, leading to lower ΔGH and higher electron affinity, thus showing improved HER activity over its counterpart monomeric Au25 and dimeric Au38 (Fig. 3b and c).
![]() | ||
Fig. 3 (a) Schematic diagram of M1Au24 (M = Pt, Pd) enhanced HER activity compared to Au25 NCs. Reproduced from ref. 100 with permission from American Chemical Society, copyright 2018. (b) LSVs and (c) ΔGH of Au25, Au38, and Au36Ag2. Reproduced from ref. 102 with permission from American Chemical Society, copyright 2021. (d) LSVs, (e) current density and overpotential comparison, (f) Tafel slope, (g) electrochemical impedance spectroscopy, and (h) electrochemical double-layer capacitance of MoS2, Au2/MoS2, Pd3/MoS2, Au2–Pd3/MoS2, and Au2Pd6/MoS2. (i) Stability test of Au2Pd6/MoS2. Reproduced from ref. 103 with permission from Royal Society of Chemistry, copyright 2018. |
In addition to the synergies between different metal components within clusters, there are also interactions between NCs and functional carriers, which can be combined in one system to maximize synergies. Du et al. designed and synthesized Au2Pd6 NCs, which can be considered to be two Pd3 NCs connected by an Au2 unit.103 On the one hand, after MoS2 couple metal NCs, the number of catalytic active sites is increased, and electron transport is promoted, resulting in enhanced HER activity. On the other hand, compared to monometallic Au2 and Pd3 NCs, and mixed Au2–Pd3 NCs, bimetallic Au2Pd6 shows the best HER activity with the lowest overpotential, the smallest Tafel slope and electrochemical resistance, and the largest electrochemically active surface area and TOF value (Fig. 3d–i). X-ray photoelectron spectroscopy and Raman results indicate the strong electronic interaction between NCs and MoS2. DFT calculations demonstrate that the largest number of catalytic sites with almost nearly zero ΔGH occur on Au2Pd6/MoS2, further explaining the excellent HER activity of Au2Pd6/MoS2 from a theoretical perspective. A similar phenomenon is observed on bimetallic Au4Cu2 NCs, which show better HER activity than that of monometallic Cu6 or Au6 NCs.104 Taking atomically precise alloy NCs as a precursor to proceed with ligand-assisted pyrolysis is a novel method to form nanocluster/single atom (NC/SA) composite systems. Lv et al. obtained a hybrid system containing two different active sites types, AuPd alloy NCs and satellite Pd SAs (AuPdNCs/PdSAs), by controllable thermal treatment of Au4Pd2(SC2H4Ph)8.105 In the hybrid system, the satellite Pd SAs play the role of optimizing the electronic structure, and the Au sites effectively promote the adsorption and dissociation of H2O molecules. Thereby, the synergy between Au and Pd promotes the excellent HER activity of AuPdNCs/PdSAs-600, which shows one order of magnitude higher mass activity and TOF than commercial Pd/C or Pt/C.
In addition to electrocatalytic HER, solar-driven photocatalytic water splitting is a sustainable and clean approach to produce hydrogen. Metal NCs exhibit a molecule-like discrete electronic energy band and strong light absorption in broad spectra, which meet the basic requirements for potential promising photocatalysts.106 Bootharaju et al. synthesized a selenolated [Au12Ag32(SePh)30]4− core–shell cluster based on a templated galvanic exchange strategy, which is comprised of an Au icosahedral core and an Ag12(SePh)30 shell.107 The synergistic effect between the Au core and the Ag shell greatly changes the overall electronic structure, which aligns well with the band structure of TiO2, facilitating photogenerated charge carrier separation. Therefore, Au12Ag32/TiO2 shows a photocatalytic H2 production rate of 6810 μmol g−1 h−1 under solar irradiation, which is 37.8 and 6 times higher than those of TiO2 and Ag44/TiO2, respectively (Fig. 4a and b). Liu et al. obtained [AuxAg25−x(SR)18]− series NCs (1 ≤ x ≤ 3, 3 ≤ x ≤ 8, 19 ≤ x ≤ 23, defined as NC-1, NC-2, NC-3).108 Under visible-light irradiation, the H2 production rate follows the order: NC-3 > NC-2 > NC-1 > Au25 > Ag25. As revealed by DFT calculation, bimetallic AuxAg25−x shows that the better photocatalytic HER activity is due to the optimization of electronic structure caused by the Au–Ag synergy. The narrowed HOMO–LUMO gap in AuxAg25−x partially contributes to the better HER performance. What is more, the microenvironment localized in dual Au–Ag sites benefits the formation of electron-acceptor centers, resulting in photogenerated electrons tending to remain in a cluster to reduce the adsorbed H+ (Fig. 4c). Additionally, the Au–Ag bimetallic synergistic effect upshifts the D-band centers and balances hydrogen adsorption/desorption dynamics (Fig. 4d and e). Kurashige et al. loaded Pd1Au24 and Pt1Au24 on BaLa4Ti4O15 via stepwise ligand-exchange, adsorption, and calcination processes, to investigate the effect of heteroatom doping type on photocatalytic HER activity.109 Although the Pd and Pt atoms are both in the center of M1Au24 when it exists in the form of NCs alone, the positions of Pd and Pt atom are different after depositing them on BaLa4Ti4O15. As can be rationally inferred from extended X-ray absorption fine structure (EXAFS) spectroscopy results, Pd is located at the surface of the NCs, while Pt is at the interface between NCs and BaLa4Ti4O15. Whether Pd or Pt doping is used, the electron density of Au in the NCs is improved. However, for photocatalytic HER activity opposite results are obtained, where Pt doping increases the activity; in contrast, Pd doping causes a decrease in activity, which to a great degree can be attributed to the doping position of the heteroatom (Fig. 4f).
![]() | ||
Fig. 4 (a and b) Comparison of photocatalytic HER activity of TiO2, Ag44/TiO2, and Au12Ag32/TiO2. Reproduced from ref. 107 with permission from Wiley-VCH, copyright 2023. (c) The isosurfaces of charge density difference of H*-adsorbed NCs. (d) D-band center and (e) ΔGH of [AuxAg25−x(SR)18]− series NCs. Reproduced from ref. 108 with permission from Royal Society of Chemistry, copyright 2023. (f) Proposed structures of M1Au24–BaLa4Ti4O15 for M = (Au, Pd, Pt) before (top) and during (bottom) the water splitting reaction. Reproduced from ref. 109 with permission from American Chemical Society, copyright 2019. |
![]() | ||
Fig. 5 (a) Schematic diagram of Au15Ag23 NC as efficient HER and OER catalyst. Reproduced from ref. 112 with permission from Elsevier, copyright 2024. (b) Structural analysis of Au1Ag21 NC. (c) Comparison of ORR performance of Au1Ag21 and Ag22 NC. Reproduced from ref. 124 with permission from Royal Society of Chemistry, copyright 2021. (d) Structural scheme of metal-doped Au25 NC. Comparison of (e) HER, (f) OER, and (g) ORR activity of Au25 and different metal-doped Au25 NCs. Reproduced from ref. 127 with permission from Royal Society of Chemistry, copyright 2020. |
The electrocatalytic OWS performance of Au NCs can be regulated by utilizing the electron-attracting ability of an Fe atom to Au NCs. Sun et al. synthesized a hybrid system of Au–Fe1NCs loaded on Ni foam to act as alkaline HER and OER bifunctional electrocatalysts.113 The electron structure of H adsorption on the Au NC surface is optimized by the introduction of Fe atoms, manifested by a low overpotential of 35.6 mV to reach 10 mA cm−2 current density in HER. In addition, the hybrid system Au–Fe1NCs display good OER activity with a 246 mV low overpotential at 10 mA cm−2 oxidation current density. In a two-electrode alkaline OWS device, the cell voltage to reach 10 mA cm−2 current density is only 1.52 V, and the durability test lasts for 40 h. The synergistic effect in Au–Fe1NCs is elucidated by X-ray photoelectron spectroscopy (XPS) experiments and DFT theoretical calculations, mainly reflected in the promotion of electron delocalization and enhancement of O–H bond activation of Au NCs and an upshift in the D-band center of the Fe atom.
The synergistic effect induced by heteroatom doping has shown great improvement in electrochemical activity; however, the understanding of the effect on photo-/photoelectro-catalytic activity, especially that involving charge transfer characteristics, remains blank. Su et al. took glutathione-protected Ag-doped Au NCs (Au1−xAgx@GSH) and a non-conjugated insulating polymer of poly(diallyl-dimethylammonium chloride) (PDDA), as negatively and positively charged building blocks to elaborately fabricate a spatially multilayered alloy NCs/metal oxide photoanode heterostructure via layer-by-layer (LBL) assembly.114 Au1−xAgx@GSH acts as a photosensitizer that can be instantly photoexcited to generate e− and h+, and PDDA plays the role of an electron-withdrawing mediator to accelerate interfacial charge transfer and form a cascade electron transport channel. As a result, LBL-assembled TiO2/(PDDA-Au1−xAgx)n shows good photoelectrochemical H2O oxidation performance under visible-light irradiation.
According to theoretical calculation, the decrease in the core size of Au NPs will result in a narrowed D-band and shift to the Fermi level, which are considered to be beneficial for O2 adsorption.120,121 At an early stage, Chen et al. utilized two series of Au NCs with different core sizes to investigate the effect of size on ORR activity, and they found the ORR catalytic activity of both follows the trend where a smaller size of NCs results in better ORR reactivity (Au11 > Au25 > Au55 > Au140 and Au25 > Au38 > Au144).122,123 Besides the core size of the NCs, heteroatom doping is another critical factor influencing the ORR activity of Au NCs. Zou et al. synthesized M1Ag21(dppf)3(SAdm)12](BPh4)2 (M = Au/Ag) by introducing 1,1′-bis-(diphenylphosphino)-ferrocene (dppf) as an activating ligand.124 Dppf is a derivative of ferrocene as well as an electron donor, acting as a common protective ligand used to design and synthesize a metal complex with excellent electrochemical activity. Compared to other Au/Ag NCs without a dppf ligand, the two NCs show better ORR activity. In addition, the two NCs exhibit the same icosahedral M13 (M = Au/Ag) kernel with a protective shell of 3 Ag3(SR)4 motifs and the protection of 3 dppf ligands (Fig. 5b). The identical geometric structure provides an ideal platform to reveal the heteroatom doping effect. Au1Ag21 shows higher onset potential and diffusion limited current density, indicating better ORR activity than that of Ag22 (Fig. 5c).
Xu et al. simultaneously introduced a dppf surface ligand and a Cd surface doping atom to perform a surface engineering strategy, and then obtained Au27Cd1(SAdm)14(dppf)Cl NCs.125 Au27Cd1 and Au38 NCs have a similar bi-icosahedral Au23 core, but different surface structures due to surface engineering. The reconstructed surface structure offers enhanced ORR activity for Au27Cd1. The synergistic effect between Au and Cd is not only reflected in a comparison of the ORR catalytic activity of Au27Cd1 and Au38 NCs but is also manifested in the improved ORR activity of Au22Cd1 compared to that of Au24 NCs. Furthermore, based on the Koutecky–Levich plots, the overall number of electrons transferred for Cd-doped Au NCs is higher than that for bare Au NCs (Au38:
2, Au27Cd1: 3.3, Au24
:
2.2, Au22Cd1: 3), indicating that Cd doping will change the ORR reaction pathway of Au NCs. In addition to Ag/Cd doping, AuPd alloy NCs show better ORR activity than Au NCs. Yan et al. synthesized GSH-protected AuPd NCs and tuned the structure and composition with the Pd-to-Au ratio.126 The AuPd NCs are supported on carbon nanosheets and the protective ligand is completely removed by calcination. The hybrid composite with 30% metal mass loading and Pd-to-Au ratio of 1
:
2 exhibits optimized ORR activity.
Although there is lots of research into Aun(SR)m-related alloy NCs in electrocatalytic HER, OER, and ORR, the activities are obtained under different experimental conditions, so it is difficult to make a deep comparison and develop a general rule about the methods needed to achieve high activity. Kumar et al. systematically investigated the effect of heteroatom species on the activity of these reactions.127 Compared to Cu/Ag doping, Pd-doped Au NCs show the highest current density in all three reactions (Fig. 5d–g). Other parameters, such as the number of constituent atoms or ligand functional groups, are also carefully evaluated. The results indicate that the decreased number of constituent atoms and thickness of the ligand layer are helpful for an improvement in activity in all three reactions.
In addition, direct formic acid fuel cells (DFAFCs) fueled by formic acid are considered to be one of the most promising power sources for portable electronic devices in the future due to their small size, low toxicity and low penetration of Nafion membranes.139,140 Pt is generally considered to be the most effective catalyst towards the anodic reaction of DFAFCs, formic acid oxidation (FAO). However, due to its poor resistance to CO toxicity, its current development faces a bottleneck. Lu et al. synthesized single-Pt-atom-doped Au25 (Pt1Au24(SR)18) NCs.141 Although only one Pt replaces the central Au atom, Pt1Au24 shows excellent FAO performance with a high mass activity of 3.7 A mgPt+Au−1, which is 12 and 34 times higher than that of Pt NCs and commercial Pt/C, respectively. Moreover, the unavailability of adjacent Pt atoms in Pt1Au24 NCs effectively suppresses the poisoning of CO intermediates, which is reflected in the outstanding performance in an accelerated durability test and CO stripping experiment. On the basis of in situ electrochemical FTIR observation and thermodynamic and kinetic calculation, the FAO process on Pt1Au24 is speculated to be a direct pathway with COOH* as the preferred reactive intermediate. The introduction of a Pt atom leads to a decreased HOMO–LUMO gap and higher reactivity, and the original Au atomic shell plays a CO anti-poisoning role. The synergistic effect in Pt1Au24 NCs contributes to the high activity and good stability.
At present, the possible main products of CO2 electroreduction include C1 products (e.g., carbon monoxide, formic acid, methane, methanol) and C2+ chemicals (e.g., ethylene, ethanol, ethane, n-propanol).145 Among the different products of CO2RR, the conversion of CO2 to CO is considered to be one of the most promising reactions in the chemical industry due to its technical and economic feasibility.147 Due to the relatively weak binding of *CO intermediates to Au, Au-based materials show particularly high selectivity for the formation of CO. It is believed that alloying or doping can further improve the CO2RR efficiency of Au NCs by optimizing the electronic structure and surface geometry. Zhuang et al. employed Au44(TBBT)28 as the parent NCs and chose Cd2+ as the oxidative ions to prepare Au47Cd2(TBBT)31 NCs by the anti-galvanic reduction (AGR) method in a two-phase system.148 The doped Cd atoms are on the surface and are coordinated by three thiolates. Compared to Au44(TBBT)28, Au47Cd2(TBBT)31 shows higher faradaic efficiency (FE) for CO over a wide potential range from −0.3 V to −0.9 V, displaying FECO up to 96% at −0.57 V (Fig. 6a). DFT calculations indicate that introducing Cd atoms results in a change in the adsorption configuration for the COOH* intermediate. Unlike Au44(TBBT)28, the O atom of COOH* prefers to bind with Cd, leading to the formation of Cd–O–C(OH)–Au and decreased COOH* formation energy, benefiting the enhancement in CO2RR activity.
![]() | ||
Fig. 6 (a) Schematic diagram of the synthesis and comparison of the CO2RR activity of Au44 and Au47Cd2 NCs. Reproduced from ref. 148 with permission from Wiley-VCH, copyright 2020. (b) Schematic diagram of the structural dynamic evolution of Au19Cd2 during CO2RR. Reproduced from ref. 149 with permission from Wiley-VCH, copyright 2021. (c) Synthesis scheme and comparison of the CO2RR activity of Au4Pd6 and Au3AgPd6 NCs. Reproduced from ref. 152 with permission from American Chemical Society, copyright 2022. (d) Schematic diagram of tuning selectivity in CO2 catalytic conversion by Au9 and Au8Pd1 NCs. Reproduced from ref. 153 with permission from Chinese Chemical Society, copyright 2021. |
In addition, Li et al. synthesized Cd-surface-modified Au NCs (Au19Cd2(SR)16, SR = cyclohexanethiolate) by taking Au23(SR)16 as the template.149 Compared to Au23, the kernel of Au19Cd2 is maintained, except that two surface Au atoms are replaced by Cd dopants. The introduction of Cd obviously improves the selectivity and activity of CO formation. The determination of active sites is investigated in detail. Compared to the inactivity of fully ligand-protected NCs, partial ligand detachment is considered a necessity for high catalytic activity. However, it needs to be carefully studied which part (–R or –SR) it is reasonable to remove, and which exposure site (S or Au) is the real catalytic site. Through DFT calculations, for both NCs, removing the –R group is considered to be more feasible than losing the –SR group, and the exposed S site is likely to be the active center for CO2RR. In contrast, if –SR is removed, the exposed Au site is more selective for HER than for CO2RR. Furthermore, for Au19Cd2 NCs, the S active site (Cd–S–Au staple motif) will undergo a dynamic regeneration process during CO formation and desorption, contributing to the different *CO binding mode and leading to the decreased energy for CO formation (Fig. 6b). At the same time, Sun et al. investigated the CO2RR activities of Au25(SR)18, Au24Cd1(SR)18, Au19Cd4(SR)18, and Au38Cd4(SR)30, and found that Cd doping can selectively control the cleavage of the Au–S or C–S bond.150 They reached a similar conclusion: that is, the exposed open S site obtained by cleaving the C–S bond is readily bound to CO2 and favourable for CO2RR, while breakage of the Au–S bond leads to the exposure of the metal site, which is preferable for HER. Recognition of the real catalytic site in the two studies perfectly reflects the advantages of atomically precise NCs in revealing the catalytic mechanism and exploring the relationship between structure and activity.
In a comparison of CO2RR performance between Pd1Au24(SR)18 and Au25(SR)18 NCs (R = –CH2CH2Ph), Li et al. found that Pd1Au24 can greatly improve the CO2RR selectivity, with ∼100% FECO, especially at high potential (up to −1.2 V), while the FECO of Au25 NCs starts to decrease at −0.9 V.151 Theoretical calculations demonstrate that the surface S atom is the active site, and Pd1Au24 can increase the thermodynamic barrier for ligand removal and retain a larger population of S active sites compared to Au25, which contributes to the improved CO2RR selectivity in the extended potential range. As Pd doping can effectively improve the activity of Au NCs in CO2RR, it is natural to think of two-element doping to form trimetallic Au-based NCs. Zhuang et al. utilized the AGR method to synthesize Au3AgPd6(TBBT)12 NCs by introducing a kernel Ag single atom on the basis of Au4Pd6(TBBT)12.152 The trimetallic Au3AgPd6 NCs show higher CO2RR activity and selectivity than Au4Pd6 with a similar structure (Fig. 6c). The introduction of Ag causes a significant change in electronic structure, decreases the d-center, weakens *CO adsorption, and promotes the release of CO, thereby improving the CO2RR activity. Meanwhile, the difficulty of *H desorption is increased, suppressing HER and enhancing CO2RR selectivity. Pd doping can not only enhance the activity of CO2RR but also shows the ability to change product type. Cai et al. prepared Au-based NCs (Au9 and Au8Pd1) confined in a layer of montmorillonite to form NC-based heterogenous catalysts.153 Pd doping can reduce the propensity for structural variation induced by the change in coordination number of the surface Au site during the reaction, bringing about multiple benefits to the catalysis. In a traditional fixed-bed reactor, Au8Pd1 exhibits unique catalytic performance in the CO2 hydrogenation reaction, changing the product type from methane to ethane, and improving the catalytic stability as well (Fig. 6d).
Many reports have also shown that AuAg alloy NCs are a highly efficient electrocatalyst for CO2RR. Seong et al. developed an active-site transplantation strategy to synthesize core−shell AuAg12@Au12(SEtPh)18 NCs (SEtPh = 2-phenylethanethiolate) by replacing the Ag12(SR)18 shell of Ag25(SR)18 NCs with an Au12(SR)18 shell.154 Compared to Ag25(SR)18 NCs, the active-site-engineered AuAg12@Au12(SEtPh)18 NCs show significantly improved CO2RR activity with a 200 mA cm−2 industry current density and a 2.1 V full-cell potential in a zero-gap CO2-to-CO electrolyzer. Lin et al. prepared AuAg26(SAdm)18S− (HSAdm = 1-adamantanethiolate) NCs composed of an AuAg12 icosahedron kernel and an Ag14(SR)18S open shell.155 Unlike Ag25(DMT)18− (DMT = 2,4-dimethylbenzenethiol) and Au21(SAdm)16 with a closed shell, due to the open shell structure in AuAg26(SAdm)18S−, the partial Ag atoms of the AuAg12 kernel are not protected by thiols, resulting in the exposure of some facets of the kernel. Therefore, owing to the unique open shell structure and Au–Ag synergistic effect, AuAg26 exhibits lower *COOH formation free energy and displays increased CO2RR activity (98.4% FECO at −0.97 V), compared to Ag25 and Au21 NCs (Fig. 7a). Based on the superior CO2RR activity of AuAg alloy NCs, Li et al. further investigated the effect of the accessibility of metal sites on the CO2RR performance by systematically studying a series of alkynyl-protected AuAg NCs (AunAg46−n(CCR)24Cl4(PPh3)2, Au24Ag20(C
CR)24Cl2, and Au43(C
CR)20/Au42Ag1(C
CR)20) of similar size and structure but different surface ligand coverage.156 Among these NCs, Au43(C
CR)20 and Au42Ag1(C
CR)20 with the highest number of accessible metal sites exhibit the highest FECO (Fig. 7b). Xu et al. utilized two AuAg alloy NCs (Au24Ag20 and Au43Ag38) to explore how the hierarchical assembly influences CO2RR activity.157 There are some correlations and differences between the two NCs, such as Au43Ag38 maintaining the kernel framework from the parent Au24Ag20 and it can be regarded as a dimeric form of Au24Ag20 monomeric NCs; Au24Ag20 is racemic, while Au43Ag38 is mesomeric; Au24Ag20 exhibits superatomic electronic configurations, while Au43Ag38 has molecule-like characteristics. Possibly due to the different atomic packing structures (individual-core vs. dual-core) and surface motif arrangement (parallel vs. crossed), the Au24Ag20 monomers show better CO2RR activity than Au43Ag38 dimers.
![]() | ||
Fig. 7 (a) Schematic diagram of open shell structure of AuAg26(SAdm)18S− for enhanced CO2RR activity. Reproduced from ref. 155 with permission from American Chemical Society, copyright 2021. (b) The metal sites accessible to CO2 on the surface of different Au/Ag NCs. Reproduced from ref. 156 with permission from Royal Chemical Society, copyright 2023. (c) Schematic diagram of Au15Cu4-enhanced CO2RR activity compared to Au18 NCs. Reproduced from ref. 158 with permission from American Chemical Society, copyright 2023. (d) Comparison of CO2RR activity of structural analogs of Au8Cu1 and Au8Ag1 NCs. Reproduced from ref. 160 with permission from Wiley-VCH, copyright 2024. |
Cu is considered as a special element for CO2RR due to the suitable adsorption energy for *CO and *CHXO intermediates; therefore Cu-doped Au NCs have received a lot of attention. Deng et al. synthesized [Au15Cu4(DPPM)6Cl4(CCR)1]2+ NCs (DPPM = bis(diphenylphosphino)methane; HC
CR = 3,5-bis-(trifluoromethyl)phenylacetylene) via a one-pot method.158 Compared to the monometallic structural analogue, [Au18(DPPM)6Br4]2+, Au15Cu4 shows a dramatic enhancement in CO2RR activity with a high FECO of >90% and an up to industrial CO partial current density of −413 mA cm−2 at −3.75 V in a membrane electrode assembly (MEA) cell (Fig. 7c). The exposed Au–Cu dual sites synergistically modulate the d-state shift, contributing to the enhanced catalytic activity. Ding et al. also explored the Au–Cu synergistic effect on CO2RR activity.159 They found that Au1Cu24 NCs are preferable for the CO2RR, while Cu25 NCs tends to progress HER, because AuCu synergy can effectively suppress HER due to the contracted electronic distribution. Based on the single Cl-terminated coordination strategy, Su et al. obtained a pair of structural analogs, [Au8Ag1(SAdm)4(Dppm)3Cl]2+ and [Au8Cu1(SAdm)4(Dppm)3Cl]2+.160 Due to the same overall structure and metal doping number and position, the metal doping type effect can be decoupled from various entangled influencing factors to elucidate metal synergies solely in terms of atomic differences. Au8Cu1 shows higher FECO than Au8Ag1 due to the generation of thermodynamically more stable *COOH and lower *CO formation energy (Fig. 7d). The rare structural analogs help us clearly reveal the impact of single-metal change on the catalytic performance. Doping a foreign metal atom to form Au-based alloy NCs can not only regulate the electronic and geometric structure to enhance catalytic activity but also minimize Au usage and reduce the catalyst cost when doping with non-precious metal atoms. Seong et al. developed a highly-efficient and economic CO2RR catalyst with high mass activity, Au4Ni2 NCs, by the transplantation of Au active sites into non-precise Ni4 NCs (Fig. 8a).161 In Au4Ni2 NCs, the Au atoms are the active sites, the Ni atoms mainly play the role of a cost reducer, and thereby the utilization efficiency of the Au atoms is improved. The TOF and mass activity of Au4Ni2 NCs for CO2-to-CO are 206 molCO molNC s−1 and 25228 A/gAu, respectively.
![]() | ||
Fig. 8 (a) Schematic diagram of active site transplantation from Ni4 to Au4Ni2 NCs for enhanced CO2RR. Reproduced from ref. 161 with permission from American Chemical Society, copyright 2023. (b) Schematic diagram of the different degrees of symmetry breaking in Au13Cux (x = 0–4) NCs for different CO2RR performances. Reproduced from ref. 166 with permission from Wiley-VCH, copyright 2024. (c) Electronic structure energy level and (d) comparison of CO2RR activity of Au38, Pt1Au37, and Pt2Au36. Reproduced from ref. 167 with permission from Wiley-VCH, copyright 2022. |
For CO2RR, the activation of CO2 is critical because it is the first step in CO2 conversion. CO2 is a linear molecule with high stability, resulting in difficulty in its activation.162 Constructing asymmetric charge distribution sites is an effective way to activate CO2,163,164 and our previous review introduced several strategies of symmetry-breaking sites in metal NCs for enhancing CO2RR activity, including a heteroatom doping induced strategy.165 For example, Tan et al. synthesized a series of Au13Cux alloy NCs (x = 0–4) with different degrees of symmetry-breaking on the crystal structure by manipulating Cu atom doping (Fig. 8b).166 Au13Cu1, Au13Cu2, and Au13Cu4 have C3V symmetry, C3h symmetry, and Td symmetry, respectively, while Au13Cu3 exhibits an asymmetric surface structure with 2 three-coordinate Cu1S3 and 1 two-coordinate Cu1S2 motifs. With the highest asymmetry degree of Au13Cu3, the unsaturated Cu1S2 site on the surface more readily coordinates with CO2, promoting CO2 adsorption and CO2RR activity. As a result, Au13Cu3 shows the highest current density of 70 mA cm−2 within the Au13Cux family of NCs (Au13Cux (x = 1, 2, 4): 40 mA cm−2, and Au13:
15 mA cm−2). In addition, Liu et al. obtained identical geometric structural Pt1Au37(SR)24 and Pt2Au36(SR)24 (SR = 4-tert-butylbenzyl mercaptan) by the controllable asymmetrical and symmetrical doping of Pt atoms into parent Au38(SR)24 NCs.167 The structure of Pt1Au37 is that one Pt atom asymmetrically dopes into one of the two cores of Au38, and relative to Au38, Pt1Au37 displays higher HOMO energy with the loss of one valence electron. While the two Pt atoms in Pt2Au36 are symmetrically located in the two cores of Au38, and Pt2Au36 exhibits a narrowed HOMO–LUMO gap and loses two valence electrons, compared to Au38 (Fig. 8c). The difference in doping mode causes the modulation of electronic structure; thereby, Pt1Au37 shows high electron-spin-induced CO2RR activity, followed by Au38 and Pt2Au36 (Fig. 8d). Besides, Wang et al. utilized an Fe single atom (SA) site to modify Au8 NCs and integrate SA and NC into one system.168 Due to the introduction of Fe SA, asymmetric charge distributed active sites are constructed and the key intermediate *COOH adsorption energy is optimized to benefit an improvement in CO2RR activity, leading to ∼18.07-fold amplification in the FE of CO2-to-CO compared to isolated Au8 NCs.
Yao et al. firstly synthesized Pt/Pd-doped Au-based NCs (Au4M2(SR)8, M = Pd/Pt) with a precise and controllable size distribution and metal doping and then removed partial ligands by thermal treatment to anchor it onto defective graphene to create a robust supported NC-based catalyst (Fig. 9a).176 DFT calculations indicate that heteroatom doping plays an essential role in N2 activation via enhanced electron back donation to N2 antibonding π*-orbitals. By precise metal doping, Au4Pd2 shows higher FENH3 and NH3 yield at −0.2 V than Au4Pt2 in the electrocatalytic N2 reduction reaction (NRR) (Fig. 9b and c). From the perspective of tuning intrinsic catalytic activity of NCs, Han et al. introduced non-precise metal Ni with the ability to suppress HER (the competitive reaction to NRR) to prepare M4Ni2 (M = Au/Ag) NCs.177 Compared to monometallic NCs (Au6, Ag6, and Ni6), bimetallic M4Ni2 (M = Au/Ag) shows better NRR performance. On the one hand, the superior NRR activity is attributed to the effective inhibition of HER by the introduction of Ni. On the other hand, partial ligand detachment during the electrochemical process provides exposed active sites to access more N2 reactant and induces reconstruction of the electronic structure, benefiting the NRR (Fig. 9d).
![]() | ||
Fig. 9 (a) Schematic diagram of Au4M2/G (M = Pt, Pd) for NRR. (b) NH3 yield rate and (c) FE of NH3 comparison of Au4Pt2 NC and Au4M2/G (M = Pt, Pd). Reproduced from ref. 176 with permission from Springer Nature, copyright 2020. (d) Schematic diagram of regulation of intrinsic activity of M4Ni2 (M = Au, Ag) NCs for NRR. Reproduced from ref. 177 with permission from Wiley-VCH, copyright 2022. (e) Structure of Au28Cu12, (f) Au28, and (g) Cu28 NCs. (h) NH3 yield rate comparison of Au28Cu12, Au28, and Cu28 NCs. Reproduced from ref. 181 with permission from Royal Chemical Society, copyright 2024. (i) NH3 yield rate of Au4Ru2/TiO2–Ov, TiO2–Ov, Au4Ru2/TiO2, and TiO2 under different sources of light irradiation. (j) Proposed mechanism for photocatalytic NRR on Au4Ru2/TiO2–Ov. Reproduced from ref. 182 with permission from Royal Chemical Society, copyright 2020. |
It has been reported that Cu is considered to be a promising candidate for electrocatalytic NO3− reduction (NO3RR) to NH3.178 However, Cu is prone to surface poisoning due to spontaneous oxidative dissolution as well as easy adsorption of other species in the electrolyte.179,180 Tang et al. doped Cu atoms on the surface of Au28 NCs to form stable AuCu alloy NCs (Au28Cu12(SR)24(PPh4)4 (SR = 2,4-dichlorothiophenol).181 Due to the synergistic effect between Au and Cu, the geometric and electronic structure of Au28Cu12 are tailored to be suitable for NO3RR. Compared to monometallic Au28(TBBT)20 (TBBT = 4-tert-butylbenzenethiol) and Cu28(CHT)18(PPh3)3 (CHT = cyclohexanethiol), bimetallic Au28Cu12 shows higher NH3 production activity and stability (Fig. 9e–h).
Sun et al. first screened the feasibility of light-driven NRR by a series of Aun NCs (n = 8, 9, 11, 18, 23, 24, 25, 28, 36, 44).182 Due to the physical adsorption of N2 on the surface of Aun (the distance between N2 and Aun is longer than 3.2 Å), it is impossible for N2 to coordinate to Aun for activation. Considering Ru is a recognized NRR candidate, Ru-doped Au NCs (Au4Ru2(PPh3)2(SC2H4Ph)8) are synthesized and then loaded onto the oxygen vacancies of TiO2 to form an NC-based heterogeneous photocatalyst (Au4Ru2/TiO2–Ov) (Fig. 9i). Under light irradiation, both Au4Ru2 NCs and TiO2 can directly generate e–h pairs and photoinduced electron transfer from TiO2 to Au4Ru2 NCs. The introduction of Ru atoms can effectively promote electron injection to adsorbed N2 and then realize N2 activation (Fig. 9j). The supported bimetallic Au4Ru2 NCs displays better photocatalytic NRR activity than homogold Aun NCs (Fig. 9i). This work not only indicates the ability for light harvesting and photogenerating e–h pairs over NCs with non-metallic or excitonic behavior, but also demonstrates a cooperative effect based on heteroatom doping.
In the metal NC family, Au-based NCs not only exhibit the advantages of simple synthesis, high yield, and easy functionalization but also possess the characteristics of high stability and solubility in various solvents, therefore attracting lots of attention from catalysis researchers. Unlike previous reviews on alloy NCs and NC catalysis (either focusing on the controllable synthesis of NCs or involving general catalytic reactions), this perspective aims at the unsatisfactory activity and dilemma about the unclear reaction mechanism facing the energy-related small-molecule catalytic reaction at the present stage, mainly summarizing the synergistic effect on an enhancement in catalytic performance in atomically precise Au-based bimetallic NCs. From the above summary and discussion, great efforts have been made in this field, and gratifying developments and advances have been achieved. Finally, in this section, some of our own perspectives and insights are put forward, to promote further development, innovation and breakthroughs in this field in the future.
(1) The synthesis range of doped metal NCs needs to be further expanded. It is not limited to bimetallic NCs. At present, there have been reports on trimetallic alloy NCs. We can learn from ideas about the preparation of high-entropy alloy NPs,183,184 to try to synthesize polymetallic alloy NCs, and perhaps some novel and unexpected phenomena or properties will appear in the subnanometer region. In addition, as far as synthesis is concerned, the rapid development of society no longer allows us to continue to blindly use the “trial and error” approach, which is too inefficient. We can use the advanced scientific and technological achievements of computer science, allowing chemistry, catalysis, computer technology and other interdisciplinary subjects to cross-fertilise and integrate, use artificial intelligence to predict the kind of doping, the amount of doping, and the position of doping for the best catalytic performance, which can greatly improve preparation efficiency.185,186 Additionally, they can be combined with a high-throughput NC synthesis technique to quickly prepare multiple types of NCs at one time in order to quickly screen their catalytic properties.187
(2) Considering future practical application scenarios, the cost-effectiveness of catalysts is an important consideration. It is necessary to develop non-noble metal clusters, such as transition metal clusters (iron-based, cobalt-based, nickel-based, copper-based clusters, etc.). However, due to their inherent instability (easily reacting with oxygen and water), the synthesis conditions are relatively harsh, and the purity and yield of the obtained product are unsatisfactory. Therefore, it is necessary to choose more suitable protective ligands to stabilize the transition metal core. In addition, synthesis can also be considered in an anhydrous and oxygen-free environment in a glove box. The large-scale preparation of cluster-based catalysts also needs to be given attention. Currently, chemical synthesis methods are relatively mature, and there are many types of clusters that can be prepared by this method, and high-purity and high-yield products can be obtained through precise control over chemical reaction conditions. However, the synthesis of cluster-based catalysts is still in the laboratory stage, and there are only a few reports in the literature where the yield of one-time synthesis reaches the gram level.188–191 There is still a long way to go to achieve industrial-scale preparation. Furthermore, for the recycling and recovery of clusters after catalytic reaction, one strategy with the most potential and promise is the construction of cluster-based composites. Due to the small size of the clusters, a support is required to load them, while avoiding aggregation or mass loss after the reaction. Porous materials are good candidates due to their abundant and adjustable pores and channels. If the surface of the material could be modified with various functional groups, it would be even better because it could solidly connect with clusters, and various types of surface groups may also bring about unexpected improvements in catalytic performance.
(3) For bimetallic NCs, and even for polymetallic NCs that may be used in the future, the enhancement in their catalytic performance is mainly attributed to the synergistic effect between metal atoms, but this statement is very general. Specifically, it is not particularly clear how the components affect each other or regulate each other, and how they play a synergistic role. Do the different metal atoms regulate each other's electronic structure, jointly affect the adsorption, activation and desorption behaviour of reaction molecules/intermediates/product molecules, and then realize synergistic catalysis? or is it tandem/cascade catalysis? for example, in CO2RR, in an Au–Cu alloy, the Au site usually plays the role of converting CO2 into CO, and then CO reduction and coupling continue to occur at the Cu site, achieving efficient tandem/cascade catalysis.192 This is a scientific question worthy of detailed discussion and careful investigation.
(4) An NC is a complex system involving an inner core of metal atoms and outer organic ligands, each of which plays a role in catalysis. In some catalytic cases, it has been found that it is not only the surface atoms that participate in the catalytic reaction; even the metal atoms at the center of the metal core also contribute to catalysis.153,167 Additionally, in cases of the Au–Cd NCs catalysis of CO2RR, the real catalytically active center (the exposed Au site with –SR detachment, or the exposed S site with –R removal) has been investigated in detail.149,150 All these inspire us to make more careful explorations in the future to determine the real active sites. In addition, for an expansion of types of energy catalysis, especially those involving multiple reactants and steps, it will be meaningful and useful to clarify the real role each component plays in catalysis to guide catalyst design in the future.
(5) Although most current energy catalytic reactions are still laboratory-level, the future will certainly tend toward industrialization, and there are already some reports about HER, ORR, CO2RR, etc. at industrial current density at this stage.193–195 As mentioned above, using NCs alone still faces the problem of easy aggregation and declining activity. Therefore, in industrial catalysis, functional supports are needed to load NCs to strengthen their stability and increase service lifetime. Therefore, it is necessary to clarify the interface and interaction between NCs and supports, so as to purposefully select appropriate supports and form supported NC-based catalysts with high activity for target reactions.
Overall, atomically precise bimetallic NCs have unique advantages in energy-related catalysis not only reflecting their extraordinary activity but also providing the opportunity to build structure–activity correlation; therefore, it deserves much more effort for further in-depth exploration and wide application, and the insights and general rules obtained from it that can be extended to other alloy materials and more types of catalytic reaction in the future.
Series NCS name | NCS formula | Synthesis method | Structural characteristic | Reference | ||
---|---|---|---|---|---|---|
Doping type | Doping number | Doping position | ||||
M1Au3 series NCs | [(Audppy)3AgO](BF4)2 | Metal-exchange strategy | Ag | 1 | Kernel surface doping | 196 |
M2Au4 series NCs | Au4Ni2(PPh2)2S2(PCP)2 | Direct co-reduction strategy | Ni | 2 | Kernel surface doping | 197 |
Au4Pd2(SR)8, SR = SC2H4Ph | Direct co-reduction strategy | Pd | 2 | Kernel surface doping | 105 | |
Pt2Au4L8, L = C21H28O2 | Direct co-reduction strategy | Pt | 2 | Kernel surface doping | 198 | |
M4Au4 series NCs | [Au4Cu4(π-CH![]() |
Direct co-reduction strategy | Cu | 4 | Kernel surface doping | 199 |
Au4Cd4(2, 4-DMBT)12 | Cation-assisted strategy | Cd | 4 | Kernel surface doping | 200 | |
[Au4Cu4(DPPM)2(S-Adm)5]+Br− | Direct co-reduction strategy | Cu | 4 | Kernel surface doping | 201 | |
M2Au6 series NCs | [Au6Ag2(C)(L2)6](BF4)4, L2 = 2-(diphenylphosphino)-5-pyridinecarboxaldehyde | Direct co-reduction strategy | Ag | 2 | Kernel surface doping | 202 |
[Au6Ag2C(dppy)6](BF4)4 | Direct co-reduction strategy | Ag | 2 | Kernel surface doping | 203 | |
M6Au2 series NCs | Au2Cu6(PPh2Py)2(S-Adm)6, Au2Cu6(PPh2Py)2(TBM)6 | Direct co-reduction strategy | Cu | 6 | Kernel surface doping | 204 |
Au2Cu6(S-Adm)6(P(Ph-OMe)3)2, Au2Cu6(S-Adm)6(PPh3)2, Au2Cu6(S-Adm)6(P(Ph-F)3)2 | Direct co-reduction strategy | Cu | 6 | Kernel surface doping | 205 | |
(M–Au)9 series NCs | [PdAu8(PPh3)8]2+ | Direct co-reduction strategy | Pd | 1 | Center doping | 206 |
Au4Ag5(dppm)2(S-Adm)6-(BPh4) | Direct co-reduction strategy | Ag | 5 | Kernel surface doping | 207 | |
(M–Au)10 series NCs | Au4Pd6(TBBT)12 | Direct co-reduction strategy | Pd | 6 | Kernel surface doping | 152 |
[PdAu9H(PPh3)8Cl]+ | Direct co-reduction strategy | Pd | 1 | Center doping | 208 | |
(M–Au)12 series NCs | [Au10Ag2(2-py-C![]() |
Direct co-reduction strategy | Ag | 2 | Kernel surface doping | 209 |
[Au11Cu1(PPh3)7(SPy)3]+ | Direct co-reduction strategy | Cu | 1 | Kernel surface doping | 210 | |
Au6Cu6MBT12, Au6Cu6PET12 | Metal-exchange strategy | Cu | 6 | Kernel surface doping | 211 | |
AuCu11[S2P(OiPr)2]6(C![]() |
Intercluster reaction strategy | Cu | 11 | Kernel surface doping | 212 | |
(M–Au)13 series NCs | [AuCu12(SR)6(C![]() |
Direct co-reduction strategy | Cu | 12 | Kernel doping | 213 |
(M–Au)15 series NCs | [Au7Ag8(tBuC![]() |
Direct co-reduction strategy | Ag | 8 | Kernel surface doping | 214 |
(M–Au)16 series NCs | [Au7Ag9(dppf)3(CF3CO2)7BF4]n | Direct co-reduction strategy | Ag | 9 | Kernel surface doping, Motif doping | 215 |
Au13Cux series NCs | [Au13Cu4(PPh2Py)4(SC6H4-tert-C4H9)8]+ | Intercluster reaction strategy | Cu | 4 | Motif doping | 216 |
Au13Cu4(PPh2Py)3(SePh)9 | Direct co-reduction strategy | Cu | 4 | Motif doping | 217 | |
[Au13Cu2(DPPP)3(SPy)6]+ | Direct co-reduction strategy | Cu | 2 | Motif doping | 218 | |
[Au13Cu2(PPh3)6(SPy)6]+ | Intercluster reaction strategy | Cu | 2 | Motif doping | 216 | |
(M–Au)17 series NCs | Au16Ag1(S-Adm)13 | Direct co-reduction strategy | Ag | 1 | Center doping | 219 |
Au4Ag13(DPPM)3(SR)9, SR = C8H10S | Direct co-reduction strategy | Ag | 13 | Kernel surface doping, Motif doping | 220 | |
AuxAg17−xN3(TBBT)12, N = a counter cation, x = 0, 1 | Direct co-reduction strategy | Ag | 16, 17 | Kernel doping | 221 | |
(M–Au)18 series NCs | Au15Ag3(SR)14 SR![]() |
Intercluster reaction strategy | Ag | 3 | Kernel surface doping | 222 |
[AgxAu18−x(Dppm)6Br4](BPh4)2 (x = 1, 2) | Metal-exchange strategy | Ag | 1, 2 | Kernel surface doping | 223 | |
(M–Au)19 series NCs | [Au7Cu12(dppy)6(TBBT)6Br4]3+ | Direct co-reduction strategy | Cu | 12 | Kernel surface doping | 224 |
(M–Au)21 series NCs | [Au19Cd2(SR)16]–, SR = C6H11S | Direct co-reduction strategy | Cd | 2 | Motif doping | 225 |
[Au9Ag12(SR)4(dppm)6X6]3+, SR = S-Adm/S-tBu, X = Cl/Br | Direct co-reduction strategy | Ag | 12 | Kernel surface doping, Motif doping | 226 | |
Au20Ag1(S-Adm)15 | Direct co-reduction strategy | Ag | 1 | All | 227 | |
Au20Ag1(SR)15, Au21–xAgx(SR)15 (x = 4–8), Au21–xCux(SR)15 (x = 0–5), SR = SC4H10O | Direct co-reduction strategy | Ag, Cu | 0–8 | Kernel surface doping/Motif doping | 228 | |
(M–Au)23 series NCs | Au23-xCux(SR)16, SR = SC6H12O | Direct co-reduction strategy | Cd | 2 | Motif doping | 225 |
M1Au24 series NCs | [Au24Pd(PPh3)10(SR)5Cl2], SR = SC2H4Ph | Direct co-reduction strategy | Pd | 1 | Center doping | 229 |
Au24M1(SR)18, SR = SC2Ph/PET | Direct co-reduction strategy | Hg, Cd, Pt, Pd | 1 | Kernel surface doping/Center doping | 230–232 | |
M24Au1 series NCs | [Ag24Au(SR)18]– PPh4+, SR = SPhMe2 | Direct co-reduction strategy | Ag | 24 | Kernel surface doping, Motif doping | 233 |
[Au1Ag24(Dppm)3(SR)17]2+, SR = C6H12S | Direct co-reduction strategy | Ag | 24 | Kernel surface doping, Motif doping | 234 | |
AuCu24H22(PPh3)12, AuCu24H22((p-FPh)3P)12 | Direct co-reduction strategy | Cu | 24 | Kernel surface doping | 189 | |
(M–Au)25 series NCs | [Cu13Au12(PPh3)10(SR)5Cl2]2+, SR = PhC2H4S | Direct co-reduction strategy | Cu | 13 | Kernel surface doping | 235 |
Au22Ir3(PET)18 | Intercluster reaction strategy | Ir | 3 | Kernel doping | 236 | |
Pd2Au23(PPh3)10Br7 | Direct co-reduction strategy | Pd | 2 | Center doping | 237 | |
(M–Au)27 series NCs | [Au4Ag23(C![]() |
Direct co-reduction strategy | Ag | 23 | Kernel surface doping, Motif doping | 238 |
Au25Ag2(SR)18, SR = SC2H4Ph | Direct co-reduction strategy | Ag | 2 | Motif doping | 239 | |
(M–Au)29 series NCs | Ag29−xAux(SR)12(pph)4, SR = S2C6H6 (x = 1–5) | Direct co-reduction strategy | Ag | 24–28 | Kernel surface doping, Motif doping | 240 |
[Au5Ag24(C![]() |
Direct co-reduction strategy | Ag | 24 | Kernel surface doping, Motif doping | 238 | |
(M–Au)30 series NCs | Au24Cu6(SPhtBu)22 | Intercluster reaction strategy | Cu | 6 | Motif doping | 241 |
(M–Au)34 series NCs | [Ag33AuS2(SR)18(CF3COO)9(DMF)6] | Direct co-reduction strategy | Ag | 33 | Kernel surface doping, Motif doping | 242 |
[AuAg33(BTCA)3(C![]() |
Direct co-reduction strategy | Ag | 33 | Kernel surface doping, Motif doping | 243 | |
(M–Au)38 series NCs | Au38−xCux(2,4-(CH3)2C6H3S)24 (x = 0–6) | Intercluster reaction strategy | Cu | 0–6 | Motif doping | 244 |
M2Au36 series NCs | M2Au36(SR)24, SR = SC2H4Ph/SC6H13 | Direct co-reduction strategy | Pd, Pt | 2 | Center doping | 245–248 |
(M–Au)40 series NCs | AuAg39(TBBM)21(CH3COO)11 | Intercluster reaction strategy | Ag | 39 | Kernel surface doping, Motif doping | 249 |
(M–Au)41 series NCs | [Au3Ag38(SR)24X5]2− (X = Cl or Br), SR = SCH2Ph | Direct co-reduction strategy | Ag | 38 | Kernel surface doping, Motif doping | 250 |
(M–Au)42 series NCs | Au38Cd4(DMBT)30 | Metal-exchange strategy | Cd | 4 | Kernel surface doping | 251 |
(M–Au)44 series NCs | Au12@S8@Ag32(PS)24]2+ | Direct co-reduction strategy | Ag | 32 | Kernel surface doping | 252 |
[Au12+xCu32(SR)30+x]4–, SR = SPhCF3 | Direct co-reduction strategy | Cu | 32 | Motif doping | 253 | |
Au24Ag20(2-SPy)4(PhC![]() |
Direct co-reduction strategy | Ag | 6 | Kernel surface doping | 254 | |
Au2Ag42(S-Adm)27(BPh4) | Direct co-reduction strategy | Ag | 42 | Kernel surface doping, Motif doping | 255 | |
M32Au12 series NCs | [Au12Ag32(FTP)30]4– | Metal-exchange strategy | Ag | 32 | Kernel surface doping, Motif doping | 256 |
Au12Ag32(SR)30 SR = SPhF/SPhF2/SPhCF3 | Direct co-reduction strategy | Ag | 32 | Kernel surface doping, Motif doping | 190 | |
(M–Au)45 series NCs | Au9Ag36(SR)27(PPh3)6, SR = SPhCl2 | Direct co-reduction strategy | Ag | 36 | Kernel surface doping, Motif doping | 257 |
(M–Au)48 series NCs | Au26Ag22(TBBT)30 | Direct co-reduction strategy | Ag | 22 | Kernel surface doping, Motif doping | 258 |
(M–Au)49 series NCs | [Au19Cu30(C![]() ![]() ![]() ![]() ![]() |
Direct co-reduction strategy | Cu | 30 | Kernel surface doping | 259 |
(M–Au)50 series NCs | [Au2Ag48(S-tBu)20-(Dppm)6Br11]Br(BPh4)2 | Direct co-reduction strategy | Ag | 48 | Kernel surface doping, Motif doping | 255 |
(M–Au)62 series NCs | Au34Ag28(PhC![]() |
Direct co-reduction strategy | Ag | 28 | Kernel surface doping | 260 |
(M–Au)70 series NCs | [Ag46Au24(StBu)32](BPh4)2 | Direct co-reduction strategy | Ag | 46 | Kernel doping | 261 |
(M–Au)110 series NCs | [Au80Ag30(C![]() |
Direct co-reduction strategy | Ag | 30 | Kernel surface doping | 262 |
Au57Ag53(C![]() |
Direct co-reduction strategy | Ag | 53 | All | 263 | |
(M–Au)124 series NCs | [Au52Cu72(p-MBT)55]+Cl−, p-MBT = SPh-p-CH3 | Direct co-reduction strategy | Cu | 72 | Kernel doping | 264 |
(M–Au)130 series NCs | Au130−xAgx(TBBT)55 | Direct co-reduction strategy | Ag | 98 | All | 265 |
Ligand abbreviations | Formula |
---|---|
dppy | C17H14NP |
pph | C18H15P |
dppm | C25H22P2 |
S-Adm | C10H16S |
2,4-DMBT | C8H10S |
TBM | C22H18 |
py | C5H5N |
TBBT | C10H14S |
S-tBu | C4H10S |
dppf | C34H28FeP2 |
PET | C8H10S |
CHM | C6H12S |
SSR | C6H6S2 |
BTCA | C8H10O8 |
This journal is © The Royal Society of Chemistry 2025 |