Atomically precise Cu14 and Cu13 nanoclusters for the oxygen evolution reaction: one additional Cu atom matters

Ziyi Liu a, Pan Zhu b, Xianxing Zhou a, Lubing Qin a, Xunying Liu a, Qing Tang *b and Zhenghua Tang *ac
aNew Energy Research Institute, School of Environment and Energy, South China University of Technology, Guangzhou Higher Education Mega Center, Guangzhou, 510006, China. E-mail: zhht@scut.edu.cn
bChongqing Key Laboratory of Chemical Theory and Mechanism, School of Chemistry and Chemical Engineering, Chongqing University, Chongqing, 401331, China. E-mail: qingtang@cqu.edu.cn
cKey Laboratory of Functional Inorganic Materials Chemistry, Ministry of Education, Heilongjiang University, Harbin, 150001, China

Received 14th March 2025 , Accepted 5th April 2025

First published on 7th April 2025


Abstract

Atomically precise coinage metal nanoclusters have been widely explored as model catalysts to study structure–performance relationships, yet single atom addition to the metal core to tailor the catalytic properties remains challenging. Herein, we report a pair of atomically precise Cu nanoclusters, namely [Cu14(Fur)3(PPh3)8H10]+ and [Cu13(Nap)3(PPh3)7H10]0 (hereafter referred to as Cu14 and Cu13; Fur: 2-methyl-3-furanthiolate and Nap: 1-naphthalene thiolate), which exhibit an astonishingly high degree of structural similarity, differing merely by the addition of a single Cu atom, yet possessing drastically different catalytic performance toward the oxygen evolution reaction (OER). Specifically, in 1 M KOH solution, Cu14 has a much lower overpotential than Cu13 (306 mV vs. 382 mV) to afford a current density of 10 mA cm−2, a smaller Tafel slope and a lower charge transfer resistance. Density functional theory calculations were employed to further identify the catalytically active site, confirming that the additional Cu atom in Cu14 is the key catalytic site, which can significantly lower the energy barrier of the rate-determining step in the OER. This study highlights the crucial role of single-atom addition in modulating the properties and functionalities of atomically precise metal nanoclusters, shedding light on future catalyst design.


1. Introduction

Atomically precise coinage metal nanoclusters have been emerging as ideal models and effective catalysts for a variety of catalytic reactions, mainly thanks to their unique attributes.1–3 Firstly, these metal nanoclusters usually possess ultrasmall sizes with diameters less than 3 nm, and such a high surface-to-volume ratio concomitantly signifies a marked increase in the number of active sites available for catalytic reactions. Secondly, the identified formula, definitive composition, and atomically precise structure of the metal nanocluster provide great opportunities to establish the structure–performance correlation.4,5 Finally, these metal nanoclusters have molecule-like characteristics and rich chemical reactivity; that is, the metal core or the surface-capping ligands can be precisely modified to tune the composition/structure, thereby opening new avenues to optimize the catalytic performance.6 Generally speaking, when using atomically precise metal nanoclusters as catalysts, there are two strategies that are most widely employed to further boost the catalytic performance: ligand engineering and metal core tailoring. The surface ligand not only protects the metal core, but also plays a critical role in determining the catalytic properties.7–10 For instance, Liu et al. reported that, when using thiolate Au25 nanoclusters as paradigm electrocatalysts, the oxygen evolution reaction (OER) rate-determining step can be switched by rationally choosing ligands with different electron-withdrawing/donating capabilities.11 In electrocatalytic CO2 reduction, the precise incorporation of two 2-thiouracil-5-carboxylic acid ligands into the cavity of [Au25(p-MBA)18] clusters (p-MBA = para-mercaptobenzoic acid) caused an enhanced local CO2 concentration near the active sites, thereby substantially accelerating the reaction kinetics.12 Our group also discovered that alkynyl-protected Ag32 nanoclusters exhibited higher CO2 electroreduction activity than their thiolate- and phosphine-co-protected Ag32 counterparts,13 and when anchored on NiFe layered double hydroxide (NiFe-LDH), alkynyl-protected Au28 nanoclusters demonstrated superior OER performance compared to their thiolate Au28 counterparts, as the former exhibited more significant charge transfer between the clusters and the NiFe-LDH support.14 Recently, we discovered that alkynyl-protected Cu6 nanoclusters also outcompete phosphine-protected Cu nanocluster counterparts toward the OER.15 The three cases all highlight the ligand effect, where the alkynyl-protected metal nanoclusters were able to lower the energy barrier of the rate-determining step in the electrocatalytic process. Furthermore, different ligands can tune the chemical coordination environment, thereby modulating the catalytic performance. Wu et al. synthesized an atomically defined Cu6(MBD)6 (MBD = 2-mercaptobenzimidazole) nanocluster with symmetry-broken CuS2N1 active sites, and the Cu6(MBD)6 nanocluster showed a faradaic efficiency toward hydrocarbons as high as 65.5%, but the Cu6 nanocluster with symmetric CuS3 active sites can only yield the HCOOH product. DFT calculations revealed that the asymmetric binding mode is conducive to generating the key intermediate of *COOH instead of *OCHO, thereby favoring the *CO formation and hindering the HCOOH pathway.16

Compared with ligand engineering, the metal core tailoring approach is more straightforward, as the core surface metal atoms might be the catalytically active sites.17–19 For instance, the Zang and Wang team fabricated bulky carboranealkynyl-protected Cu nanoclusters (NCs) from the monomer Cu13 and the bridged dimer Cu26, where Cu26 demonstrated superior catalytic performance in nitrate electroreduction thanks to the metal core structure.20 In a recent study, Ma et al. reported the synthesis of a [AuCu24(dppp)6H22]+ cluster, which bears exposed Cu3H3 units in specific surface cavities. In electrochemical CO2 reduction, the lattice hydrogen (H) in the Cu3H3 active unit was recognized as indispensable for the formation of the C2H4 product.21 Alloying, especially single-atom doping/substitution, has shown great potential to improve catalytic properties,22,23 as the physicochemical properties and electronic structure of metal nanoclusters can be significantly tuned,24 and introducing new active sites has also been reported in several cases. In 2017, the Lee and Jiang team prepared PtAu24(SC6H13)18 nanoclusters by simply doping a Pt atom into Au25(SC6H13)18, which exhibited significantly higher hydrogen evolution reaction (HER) activity. The doped Pt atom can induce the H atoms to spontaneously move into the sub-surface, where they are directly adsorbed, which mainly contributes to the favorable HER energetics.25 A similar case has also been documented by Li et al., where the Au24Pd nanocluster exhibited improved CO2 electroreduction selectivity to CO formation compared to that of the Au25 nanocluster.26 Recently, Zhu and his colleagues reported a pair of Au8M1 nanoclusters (M = Ag/Cu), which differ only by the metal dopant at the same site. In electrochemical CO2 reduction, Au8Cu1 possessed a much higher faradaic efficiency for CO than Au8Ag1.27 It is worth noting that, in the above cases, the main chemical structure and configuration were preserved upon doping or substitution; however, there are rare cases where additional metal atoms, even one single metal atom, have been introduced into an existing nanocluster to achieve improved catalytic activity. The Wu group successfully added two Ag atoms onto an Au25 nanocluster to form an Au25Ag2 nanocluster, where the latter one demonstrated markedly superior catalytic performance in the hydrolysis of 1, 3-diphenylprop-2-ynyl acetate.28 Dong et al. developed a facile method that was able to transform a [Cu58H20PET36(PPh3)4]2+ (PET: phenylethane thiolate) nanocluster into a surface-defect analog of the [Cu57H20PET36(PPh3)4]+ nanocluster, where the one-Cu-atom-deficient Cu57 cluster showed improved catalytic activity in click chemistry, particularly for photoinduced [3 + 2] azide–alkyne cycloaddition.29

Inspired by the above findings, herein, we synthesized a pair of atomically precise Cu nanoclusters, [Cu14(Fur)3(PPh3)8H10]+ and [Cu13(Nap)3(PPh3)7H10]0 (hereafter referred to as Cu14 and Cu13; Fur: 2-methyl-3-furanthiolate and Nap: 1-naphthalene thiolate). Cu14 exhibits an astonishingly high degree of structural similarity to Cu13, and it can be regarded as Cu13 with one additional Cu atom added to its metal core. However, they exhibit drastically different catalytic performances toward the OER. To reach a current density of 10 mA cm−2, Cu14 requires an overpotential of 306 mV, much lower than that of Cu13 (382 mV). Density functional theory calculations further confirmed that the additional Cu atom is the key catalytic site that mainly leads to the superior OER performance of Cu14 compared to Cu13. As a note, Shingyouchi et al. reported a ligand-dependent surface effect in electrochemical CO2 reduction of two [Cu14(SR)3(PPh3)8H10]+ nanoclusters, where the [Cu14(PET)3(PPh3)8H10]+ cluster possessed a much higher faradaic efficiency of HCO2H than the [Cu14(CHT)3(PPh3)8H10]+ cluster (CHT = cyclohexane thiolate).30

2. Experimental section

The synthetic route to Cu14 is shown in Scheme S1 and all the experimental details are available in the ESI.

3. Results and discussion

The Cu14 NC was first synthesized, and yellow single crystals with well-defined shapes were obtained by a solvent diffusion crystal growth method (Fig. S1). The CCDC number of the Cu14 single crystal is 2417373. The crystal structure was then analyzed by single-crystal X-ray diffraction (SC-XRD). Cu14 crystallizes in the P[1 with combining macron] space group, and it is positively charged with BF4 as the counter-anion. The overall structure is shown in Fig. 1a, and it contains 14 Cu atoms, 3 Fur ligands, 8 PPh3 ligands, and 10 hydrides. Furthermore, as shown in Fig. S2, two Cu14 units are packed into one unit cell. The detailed structural parameters are summarized in Table S1. Subsequently, electrospray ionization mass spectrometry (ESI-MS) in positive-ion mode was employed to further determine the molecular composition of Cu14. As presented in Fig. 1b, the MS spectra exhibit a pronounced peak with m/z at 3336.8368 Da, in good agreement with the composition of [Cu14(Fur)3(PPh3)8H10]+ (cal. 3336.8369 Da, deviation: 0.0001 Da). The comparative cluster of Cu13 was also synthesized by following the previously reported approach.31 As a comparison, the single-crystal structure of Cu13 can be found in Fig. S3a.[thin space (1/6-em)]31 Its molecular composition was further confirmed by ESI-MS. The MS spectra in Fig. S3b show a prominent peak with an m/z of 3474.9011 Da (cal.: 3474.8997 Da, deviation: 0.0014 Da), and the simulated pattern agrees well with the experimental results. In addition, Fig. S4 shows the UV-Vis absorption spectra of Cu14 and Cu13 NCs in CH2Cl2. A featureless exponential decay pattern in absorbance is observed for both clusters, which has been known as a common phenomenon for Cu NCs.5 Both clusters can be well-dissolved in CH2Cl2 exhibiting a yellow solution (inset in Fig. S4). In addition, the chemical composition and valence state of Cu14 were investigated by X-ray photoelectron spectroscopy (XPS). Fig. S5a shows the survey scan spectra, which confirmed the presence of P, O, and C elements in the sample. As shown in the high-resolution Cu 2p XPS spectra (Fig. S5b), the binding energies of the Cu 2p3/2 and Cu 2p1/2 electrons were 932.7 and 952.6 eV, respectively, in good agreement with that of Cu(I) species.
image file: d5qi00735f-f1.tif
Fig. 1 (a) The total structure of Cu14 (blue, Cu; yellow, S; dark purple, P; red, O; grey, C; and white, H). (b) ESI-MS spectra of [Cu14(Fur)3(PPh3)8H10]+ in positive ion mode. Inset: isotopic distribution patterns of [Cu14(Fur)3(PPh3)8H10]+ – experimental (blue) and simulated (pink).

Subsequently, the metal core configurations of Cu14 and Cu13 are examined in detail. Both Cu14 and Cu13 are constructed by combining two nested trigonal antiprisms to form a Cu12 skeleton (Fig. 2a and e). Next, a Cu atom is attached along the C3 axis at the top of the skeleton in both clusters, and three thiolates are adhered at the bottom in a μ2 binding mode (Fig. 2b and f). Notably, Cu14 has an additional Cu atom bound to the three thiolates. This additional Cu atom does not induce a substantial structural change, but indeed results in a slight distortion of the overall structure. A detailed comparison reveals that the Cu6 trigonal antiprism core of Cu14 (marked in light purple in Fig. 2a) has a more regular and ordered spatial arrangement. The Cu–Cu bond lengths within the Cu6 core of Cu14 range from 2.461 Å to 3.100 Å, with an average value of 2.770 Å. Meanwhile, the bond lengths in the Cu6 core of Cu13 range from 2.494 Å to 3.192 Å, with an average value of 2.792 Å. The completely assembled core structures of both clusters are shown in Fig. 2c and g, and both are structurally symmetric along a C3 axis. Fig. 2d and h depict the coordination of both nanoclusters with PPh3 and the positions of ten hydrides. Cu13 is coordinated with seven PPh3 ligands, and the average Cu–P bond length is calculated to be 2.24 Å. Meanwhile, Cu14 is coordinated with eight PPh3 ligands, seven of which have an identical coordination mode to that of Cu13, and the 8th PPh3 is bound to an extra Cu atom, with an average Cu–P bond length of 2.23 Å. One may notice that the hydrogen atoms bonded to the Cu atoms associated with the three thiolates are pushed inwards due to the repulsive force exerted by the added Cu atom. Interestingly, despite Cu14 and Cu13 NCs having virtually similar metal core configurations, the introduction of one more Cu atom indeed causes some structural difference, where the most eye-catching part is that three thiolate ligands are more tightly compressed inward due to their coordination with the Cu atom. Consequently, the whole metal core of Cu14 is more compact compared with that of Cu13.


image file: d5qi00735f-f2.tif
Fig. 2 Analysis of the crystal structures of Cu14 and Cu13 NCs. The structural disassembly of (a)–(c) Cu14S3 and (e)–(g) Cu13S3. Two triangular antiprisms are combined to form a Cu12 skeleton to which two CuPPh3 or one CuPPh3 is added, respectively. The distribution of PPh3 ligands and ten hydrides in (d) Cu14 and (h) Cu13 (blue/light purple/pink, Cu; yellow, S; dark purple, P; and light grey, H).

Next, the OER was chosen as the model reaction to examine the catalytic effect of single-atom addition/introduction between the Cu14 and Cu13 clusters. The OER has been well recognized as the bottleneck reaction in electrochemical water splitting to produce H2, a high-energy-density and pollution-free green energy carrier.32,33 The OER is also the primary half-cell reaction occurring at the cathode of rechargeable metal–air batteries, and hence, it significantly determines the energy conversion efficiency.34 The OER performance of the Cu13 NC and the Cu1(PPh3)2(OH) compound was also investigated for comparison. As a note, Cu1(PPh3)2(OH) is the product obtained during the synthetic trials aimed at preparing a Cu nanocluster protected only by PPh3. The CCDC number of Cu1(PPh3)2(OH) is 2417860. Its structural details can be found in Fig. S6. Carbon black was used as a support to improve the electrical conductivity of the catalyst, and all the tests were conducted in 1 M KOH solution. Particular care was taken to ensure an equivalent Cu loading across all the samples.

Fig. 3a illustrates the OER polarization curves of the samples. At an applied voltage of 1.536 V, the Cu14 NC achieves a current density of 10 mA cm−2, outperforming the benchmark OER catalyst RuO2, as well as the Cu13 NC and Cu1(PPh3)2(OH). A direct comparison of the overpotentials of the four samples at a current density of 10 mA cm−2 is shown in Fig. 3b. The overpotential of the Cu14 NC at 10 mA cm−2 is 306 mV, which is significantly lower than that of the other samples. The improved catalytic performance of the Cu14 NC compared to the Cu13 NC is probably attributed to the incorporation of the 14th Cu atom, which modulates the geometric and electronic structures, hence facilitating the OER process. Conversely, the poor performance of Cu1(PPh3)2(OH) can be ascribed to the limited accessibility of its active sites, mainly due to the steric hindrance imposed by the bulky PPh3 ligands. The Tafel plots were recorded (Fig. S7) to reveal the reaction kinetics, where the Tafel slopes can be calculated. The Tafel slope of the Cu14 NC is lower than those of the other samples, indicating that the Cu14 NC has the most favourable reaction kinetics.


image file: d5qi00735f-f3.tif
Fig. 3 The OER performance of Cu14, Cu13, RuO2 and Cu1(PPh3)2(OH) in 1 M KOH. (a) The LSV polarization curves. (b) Overpotential comparison at 10 mA cm−2 current density. (c) The electrochemical impedance spectra. (d) Linear regression of cathodic charging currents versus scan rate. (e) ECSA comparison. (f) The it curve of Cu14 in 30 h.

To further evaluate the intrinsic catalytic activity of the samples, electrochemical impedance spectroscopy (EIS) measurements were performed and the experimental data were fitted using an equivalent circuit model. As shown in Fig. 3c, the Cu14 NC has the smallest semicircle in the Nyquist plot, corresponding to a small charge transfer resistance (Rct) value of 15.94 Ω. In contrast, the Rct values of the Cu13 NC and Cu1(PPh3)2(OH) were determined to be 41.54 Ω and 65.19 Ω, respectively. The significantly lower Rct value of the Cu14 NC indicates a much more facile electron transfer process during the OER, contributing to its enhanced catalytic activity. Furthermore, cyclic voltammetry (CV) measurements were performed in the potential range from 1.1 V to 1.2 V at scan rates varying from 10 mV s−1 to 60 mV s−1 (Fig. S8a–c). From these measurements, the double-layer capacitance (Cdl) values of the Cu14 NC, the Cu13 NC and Cu1(PPh3)2(OH) were calculated to be 20.07 mF cm−2, 13.03 mF cm−2 and 11.52 mF cm−2, respectively (Fig. 3d). A higher Cdl value indicates a larger electrochemically active surface area (ECSA), as illustrated in Fig. 3e. Finally, the long-term durability of the Cu14 NC catalyst for the OER was evaluated. As shown in Fig. 3f, after 30 hours of continuous chronoamperometry (it) testing at a constant potential of 1.536 V, the current density of the Cu14 NC catalyst exhibits only a negligible current decrease of ∼4%, demonstrating remarkably robust long-term stability for the OER. In contrast, the Cu13 NC exhibits a current decay of ∼8%, slightly inferior to that of Cu14 NC. The superior long-term stability of the Cu14 NC is probably attributed to its more symmetric and compact metal core structure.

The experimental results demonstrate that the Cu14 NC exhibits superior electrocatalytic performance for the OER compared to the Cu13 NC. In order to gain deeper insights, DFT calculations were employed to systematically compare the OER activity of Cu14 and Cu13 NCs, with a particular focus on evaluating the rate-determining step (RDS) and identifying the active sites. We primarily consider the OER activity of the NCs following the removal of a single ligand.35–39 Specifically, we first examined the removal of a single –PPh3 ligand in detail. Based on the distinct coordination environments of –PPh3 ligands in Cu14 and Cu13 NCs, we classified the –PPh3 ligands in Cu14 as P1Ph3, P2Ph3, P3Ph3, and P4Ph3 (as shown in Fig. S9a) and the –PPh3 ligands in Cu13 as P1Ph3, P2Ph3, and P3Ph3 (as shown in Fig. S9b) for differentiation. We found that, compared to the removal of other –PPh3 ligands, Cu14 with the –P4Ph3 ligand removed exhibits a more stable structure (Fig. S9c), while exposing a Cu active site coordinated with three thiolate ligands. Therefore, we selected Cu14 with the –P4Ph3 ligand removed as the catalyst for the electrocatalytic OER. The computational results indicate that at an equilibrium potential of U = 1.23 V, the conversion from *OH to *O is the RDS, with a corresponding free energy barrier of 0.96 eV. Furthermore, throughout the entire OER process, all key reaction intermediates preferentially adsorb onto the Cu active site at a top site (Fig. 4a). For the Cu13 NC, the removal of the –P1Ph3 ligand instead of –P2Ph3 or –P3Ph3 leads to superior stability of the catalyst. During the electrocatalytic OER process, the transformation of *O to *OOH is identified as the RDS, exhibiting a relatively high free energy barrier of 1.11 eV. Analysis of the adsorption configurations of the intermediates reveals that *OH and *O preferentially adsorb on the Cu atom via top-site and hollow-site adsorption, while *OOH tends to adopt a bridging adsorption mode.


image file: d5qi00735f-f4.tif
Fig. 4 The free energy diagrams (ΔG) for O2 evolution and corresponding absorption structures of (a) Cu14 and (b) Cu13 NCs after the removal of a –PPh3 ligand at U = 0 V and U = 1.23 V. Colour legend: Cu, dark orange; C, grey; S, blue; P, purple; H, white; and O, red.

Meanwhile, we also investigated the catalytic behaviour of Cu14 and Cu13 NCs after the removal of a single thiolate ligand. As shown in Fig. S10, for the Cu14 NC with the thiolate ligand removed, the key reaction intermediates preferentially adsorb at the Cu active sites in a hollow-site configuration during the catalytic process. In contrast, for the Cu13 NC after thiolate ligand removal, the reaction intermediates tend to adopt a bridge-site adsorption configuration. This difference in adsorption modes may originate from variations in the surface atomic coordination environment. Further reaction energy barrier calculations reveal that, for both Cu14 and Cu13 NCs, the conversion of *OOH to O2 is the RDS, with relatively high free energy barriers of 1.61 eV and 1.49 eV, respectively.

The aforementioned DFT calculations indicate that, in Cu14 and Cu13 NCs, the catalytically active sites are most likely the exposed Cu atoms resulting from the removal of a –PPh3 ligand. Notably, despite the high structural similarity of the metal cores in Cu14 and Cu13 NCs, the additional Cu atom in Cu14 not only induces local structural alterations but also significantly influences the catalytic properties of the NCs. Specifically, the additional Cu atom serves as the key catalytic site in the Cu14 NC. Compared to Cu13, Cu14 exhibits superior catalytic performance in the OER, highlighting the significance of atomic-level structural modifications in optimizing catalytic performance.

4. Conclusions

In summary, a pair of atomically precise Cu nanoclusters [Cu14(Fur)3(PPh3)8H10]+ and [Cu13(Nap)3(PPh3)7H10]0 were successfully synthesized and examined as model catalysts for the OER. They exhibit a high degree of structural similarity, but one additional Cu atom in Cu14 induces some local structural shrinking, and more importantly, when used as catalysts for the OER in 1 M KOH, Cu14 demonstrated markedly superior catalytic performance compared to Cu13, manifested by a lower overpotential at 10 mA cm−2, smaller charge transfer resistance, and a larger electrochemically active surface area. DFT calculations further validated that the additional Cu atom is the key catalytic site, which is able to significantly lower the energy barrier of the rate-determining step in the OER process. This study highlights the great advantages of employing atomically precise metal nanoclusters as model catalysts, and such atomic-level modification to tune catalytic performance can shed light on future catalyst design.

Data availability

All the data supporting the findings of this study are available within the paper and its ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Z. T. acknowledges funding from the Guangdong Natural Science Funds (No. 2023A0505050107). Q. T. acknowledges the National Natural Science Foundation of China (No. 22473017) and the Chongqing Science and Technology Commission (CSTB2024NSCQMSX0250).

References

  1. R. Jin, G. Li, S. Sharma, Y. Li and X. Du, Toward Active-Site Tailoring in Heterogeneous Catalysis by Atomically Precise Metal Nanoclusters with Crystallographic Structures, Chem. Rev., 2021, 121, 567–648 CrossRef CAS PubMed.
  2. Y. Du, H. Sheng, D. Astruc and M. Zhu, Atomically Precise Noble Metal Nanoclusters as Efficient Catalysts: A Bridge between Structure and Properties, Chem. Rev., 2020, 120, 526–622 CrossRef CAS PubMed.
  3. W. Jing, H. Shen, R. Qin, Q. Wu, K. Liu and N. Zheng, Surface and Interface Coordination Chemistry Learned from Model Heterogeneous Metal Nanocatalysts: From Atomically Dispersed Catalysts to Atomically Precise Clusters, Chem. Rev., 2023, 123, 5948–6002 CrossRef CAS PubMed.
  4. Z. Liu, X. Liu, L. Qin, H. Chen, R. Li and Z. Tang, Alkynyl ligand for preparing atomically precise metal nanoclusters: Structure enrichment, property regulation, and functionality enhancement, Chin. J. Struct. Chem., 2024, 43, 100405 CrossRef CAS.
  5. M. Chen, C. Guo, L. Qin, L. Wang, L. Qiao, K. Chi and Z. Tang, Atomically Precise Cu Nanoclusters: Recent Advances, Challenges, and Perspectives in Synthesis and Catalytic Applications, Nano-Micro Lett., 2025, 17, 83 CrossRef CAS PubMed.
  6. S. Li, N.-N. Li, X.-Y. Dong, S.-Q. Zang and T. C. W. Mak, Chemical Flexibility of Atomically Precise Metal Clusters, Chem. Rev., 2024, 124, 7262–7378 CrossRef CAS PubMed.
  7. Z. Guan and Q.-M. Wang, Ligand Engineering in Metal Nanoclusters: from Structural Control to Functional Modulation, Chin. J. Struct. Chem., 2020, 39, 1194–1200 CAS.
  8. Z.-J. Guan, R.-L. He, S.-F. Yuan, J.-J. Li, F. Hu, C.-Y. Liu and Q.-M. Wang, Ligand Engineering toward the Trade-Off between Stability and Activity in Cluster Catalysis, Angew. Chem., Int. Ed., 2022, 61, e202116965 CrossRef CAS PubMed.
  9. J. Wang, F. Xu, Z.-Y. Wang, S.-Q. Zang and T. C. W. Mak, Ligand-Shell Engineering of a Au28 Nanocluster Boosts Electrocatalytic CO2 Reduction, Angew. Chem., Int. Ed., 2022, 61, e202207492 CrossRef CAS PubMed.
  10. S. Yoo, S. Yoo, G. Deng, F. Sun, K. Lee, H. Jang, C. W. Lee, X. Liu, J. Jang, Q. Tang, Y. J. Hwang, T. Hyeon and M. S. Bootharaju, Nanocluster Surface Microenvironment Modulates Electrocatalytic CO2 Reduction, Adv. Mater., 2024, 36, 2313032 CrossRef CAS PubMed.
  11. Z. Liu, H. Tan, B. Li, Z. Hu, D.-E. Jiang, Q. Yao, L. Wang and J. Xie, Ligand effect on switching the rate-determining step of water oxidation in atomically precise metal nanoclusters, Nat. Commun., 2023, 14, 3374 CrossRef CAS PubMed.
  12. Z. Liu, J. Chen, B. Li, D.-E. Jiang, L. Wang, Q. Yao and J. Xie, Enzyme-Inspired Ligand Engineering of Gold Nanoclusters for Electrocatalytic Microenvironment Manipulation, J. Am. Chem. Soc., 2024, 146, 11773–11781 CrossRef CAS PubMed.
  13. L. Chen, F. Sun, Q. Shen, L. Qin, Y. Liu, L. Qiao, Q. Tang, L. Wang and Z. Tang, Homoleptic alkynyl-protected Ag32 nanocluster with atomic precision: Probing the ligand effect toward CO2 electroreduction and 4-nitrophenol reduction, Nano Res., 2022, 15, 8908–8913 CrossRef CAS.
  14. Q.-L. Shen, L.-Y. Shen, L.-Y. Chen, L.-B. Qin, Y.-G. Liu, N. M. Bedford, F. Ciucci and Z.-H. Tang, Heterointerface of all-alkynyl-protected Au28 nanoclusters anchored on NiFe-LDHs boosts oxygen evolution reaction: a case to unravel ligand effect, Rare Met., 2023, 42, 4029–4038 CrossRef CAS.
  15. M.-Y. Chen, L.-Y. Shen, L.-B. Qin, F. Ciucci and Z.-H. Tang, Atomically precise Cu6 nanoclusters for oxygen evolution catalysis: a combined experimental and theoretical study, Rare Met., 2025, 44, 2428–2437 CrossRef CAS.
  16. Q.-J. Wu, D.-H. Si, P.-P. Sun, Y.-L. Dong, S. Zheng, Q. Chen, S.-H. Ye, D. Sun, R. Cao and Y.-B. Huang, Atomically Precise Copper Nanoclusters for Highly Efficient Electroreduction of CO2 towards Hydrocarbons via Breaking the Coordination Symmetry of Cu Site, Angew. Chem., Int. Ed., 2023, 62, e202306822 CrossRef CAS PubMed.
  17. S. Wang, Q. Li, X. Kang and M. Zhu, Customizing the Structure, Composition, and Properties of Alloy Nanoclusters by Metal Exchange, Acc. Chem. Res., 2018, 51, 2784–2792 CrossRef CAS PubMed.
  18. A. Ghosh, O. F. Mohammed and O. M. Bakr, Atomic-Level Doping of Metal Clusters, Acc. Chem. Res., 2018, 51, 3094–3103 CrossRef CAS PubMed.
  19. X. Liu, X. Cai and Y. Zhu, Catalysis Synergism by Atomically Precise Bimetallic Nanoclusters Doped with Heteroatoms, Acc. Chem. Res., 2023, 56, 1528–1538 CrossRef CAS PubMed.
  20. J. Wang, J. Cai, K.-X. Ren, L. Liu, S.-J. Zheng, Z.-Y. Wang and S.-Q. Zang, Stepwise structural evolution toward robust carboranealkynyl-protected copper nanocluster catalysts for nitrate electroreduction, Sci. Adv., 2024, 10, eadn7556 CrossRef CAS PubMed.
  21. X. Ma, C. Fang, M. Ding, Y. Zuo, X. Sun and S. Wang, Atomic-Level Elucidation of Lattice-Hydrogens in Copper Catalysts for Selective CO2 Electrochemical Conversion toward C2 Products, Angew. Chem., Int. Ed., 2025, e202500191 Search PubMed.
  22. J. Yang, R. Pang, D. Song and M.-B. Li, Tailoring silver nanoclusters via doping: advances and opportunities, Nanoscale Adv., 2021, 3, 2411–2422 RSC.
  23. S. Masuda, K. Sakamoto and T. Tsukuda, Atomically precise Au and Ag nanoclusters doped with a single atom as model alloy catalysts, Nanoscale, 2024, 16, 4514–4528 RSC.
  24. H. Fakhouri, E. Salmon, X. Wei, S. Joly, C. Moulin, I. Russier-Antoine, P.-F. Brevet, X. Kang, M. Zhu and R. Antoine, Effects of Single Platinum Atom Doping on Stability and Nonlinear Optical Properties of Ag29 Nanoclusters, J. Phys. Chem. C, 2022, 126, 21094–21100 CrossRef CAS.
  25. K. Kwak, W. Choi, Q. Tang, M. Kim, Y. Lee, D.-e. Jiang and D. Lee, A Molecule-like PtAu24(SC6H13)18 Nanocluster as An Electrocatalyst For Hydrogen Production, Nat. Commun., 2017, 8, 14723 CrossRef PubMed.
  26. S. Li, D. Alfonso, A. V. Nagarajan, S. D. House, J. C. Yang, D. R. Kauffman, G. Mpourmpakis and R. Jin, Mono-Palladium Substitution in Gold Nanoclusters Enhances CO2 Electroreduction Activity and Selectivity, ACS Catal., 2020, 10, 12011–12016 CrossRef CAS.
  27. S. Su, Y. Zhou, L. Xiong, S. Jin, Y. Du and M. Zhu, Structure-Activity Relationships of the Structural Analogs Au8Cu1 and Au8Ag1 in the Electrocatalytic CO2 Reduction Reaction, Angew. Chem., Int. Ed., 2024, 63, e202404629 CrossRef CAS PubMed.
  28. C. Yao, J. Chen, M.-B. Li, L. Liu, J. Yang and Z. Wu, Adding Two Active Silver Atoms on Au25 Nanoparticle, Nano Lett., 2015, 15, 1281–1287 CrossRef CAS PubMed.
  29. C. Dong, R.-W. Huang, A. Sagadevan, P. Yuan, L. Gutiérrez-Arzaluz, A. Ghosh, S. Nematulloev, B. Alamer, O. F. Mohammed, I. Hussain, M. Rueping and O. M. Bakr, Isostructural Nanocluster Manipulation Reveals Pivotal Role of One Surface Atom in Click Chemistry, Angew. Chem., Int. Ed., 2023, 62, e202307140 CrossRef CAS PubMed.
  30. Y. Shingyouchi, M. Ogami, S. Biswas, T. Tanaka, M. Kamiyama, K. Ikeda, S. Hossain, Y. Yoshigoe, D. J. Osborn, G. F. Metha, T. Kawawaki and Y. Negishi, Ligand-Dependent Intracluster Interactions in Electrochemical CO2 Reduction Using Cu14 Nanoclusters, Small, 2024, 2409910 CrossRef PubMed.
  31. M. Bodiuzzaman, K. Murugesan, P. Yuan, B. Maity, A. Sagadevan, N. H. Malenahalli, S. Wang, P. Maity, M. F. Alotaibi, D.-E. Jiang, M. Abulikemu, O. F. Mohammed, L. Cavallo, M. Rueping and O. M. Bakr, Modulating Decarboxylative Oxidation Photocatalysis by Ligand Engineering of Atomically Precise Copper Nanoclusters, J. Am. Chem. Soc., 2024, 146, 26994–27005 CrossRef CAS PubMed.
  32. Y. Zhang, B. Wang, C. Hu, M. Humayun, Y. Huang, Y. Cao, M. Negem, Y. Ding and C. Wang, Fe–Ni–F electrocatalyst for enhancing reaction kinetics of water oxidation, Chin. J. Struct. Chem., 2024, 43, 100243 CrossRef CAS.
  33. L. Qi, Z. Chen, X. Luan, Z. Zheng, Y. Xue and Y. Li, Atomically dispersed Mn enhanced catalytic performance for overall water splitting on graphdiyne-coated copper hydroxide nanowire, Chin. J. Struct. Chem., 2024, 43, 100197 CrossRef CAS.
  34. S. Ding, H. Wang, X. Dai, Y. Lv, X. Niu, R. Yin, F. Wu, W. Shi, W. Liu and X. Cao, Mn-modulated Co–N–C oxygen electrocatalysts for robust and temperature-adaptative zinc-air batteries, Chin. J. Struct. Chem., 2024, 43, 100302 CrossRef CAS.
  35. L. Qin, F. Sun, X. Ma, G. Ma, Y. Tang, L. Wang, Q. Tang, R. Jin and Z. Tang, Homoleptic Alkynyl-Protected Ag15 Nanocluster with Atomic Precision: Structural Analysis and Electrocatalytic Performance toward CO2 Reduction, Angew. Chem., Int. Ed, 2021, 60, 26136–26141 CrossRef CAS PubMed.
  36. L. Qin, F. Sun, Z. Gong, G. Ma, Y. Chen, Q. Tang, L. Qiao, R. Wang, Z.-Q. Liu and Z. Tang, Electrochemical NO3 Reduction Catalyzed by Atomically Precise Ag30Pd4 Bimetallic Nanocluster: Synergistic Catalysis or Tandem Catalysis?, ACS Nano, 2023, 17, 12747–12758 CrossRef CAS PubMed.
  37. L. Qin, F. Sun, M. Li, H. Fan, L. Wang, Q. Tang, C. Wang and Z. Tang, Atomically precise (AgPd)27 nanoclusters for nitrate electroreduction to NH3: Modulating the metal core by a ligand induced strategy, Acta Phys.-Chim. Sin., 2025, 41, 100008 CrossRef.
  38. G. Deng, H. Yun, M. S. Bootharaju, F. Sun, K. Lee, X. Liu, S. Yoo, Q. Tang, Y. J. Hwang and T. Hyeon, Copper Doping Boosts Electrocatalytic CO2 Reduction of Atomically Precise Gold Nanoclusters, J. Am. Chem. Soc., 2023, 145, 27407–27414 CrossRef CAS PubMed.
  39. G. Deng, H. Yun, Y. Chen, S. Yoo, K. Lee, J. Jang, X. Liu, C. W. Lee, Q. Tang, M. S. Bootharaju, Y. J. Hwang and T. Hyeon, Ferrocene-Functionalized Atomically Precise Metal Clusters Exhibit Synergistically Enhanced Performance for CO2 Electroreduction, Angew. Chem., Int. Ed., 2025, 64, e202418264 CrossRef CAS PubMed.

Footnotes

Electronic supplementary information (ESI) available: experimental details, supporting figures and tables, and more calculation results. CCDC 2417373 and 2417860. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5qi00735f
These authors contributed equally.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.