Highlights
Summarizing recent advances on synthesis and catalytic applications of Cu nanoclusters.
The structure–property–functionality relationship is clearly elucidated.
Critical analysis of the current challenges and future perspectives.
Keywords: Atomically precise Cu nanoclusters, Controllable synthesis, Catalytic applications, Structure–performance relationship, Challenges and perspectives
Abstract
Atomically precise metal nanoclusters are an emerging type of nanomaterial which has diverse interfacial metal–ligand coordination motifs that can significantly affect their physicochemical properties and functionalities. Among that, Cu nanoclusters have been gaining continuous increasing research attentions, thanks to the low cost, diversified structures, and superior catalytic performance for various reactions. In this review, we first summarize the recent progress regarding the synthetic methods of atomically precise Cu nanoclusters and the coordination modes between Cu and several typical ligands and then discuss the catalytic applications of these Cu nanoclusters with some explicit examples to explain the atomical-level structure–performance relationship. Finally, the current challenges and future research perspectives with some critical thoughts are elaborated. We hope this review can not only provide a whole picture of the current advances regarding the synthesis and catalytic applications of atomically precise Cu nanoclusters, but also points out some future research visions in this rapidly booming field.
Introduction
Uniqueness of Atomically Precise Cu Nanoclusters
The last decade has witnessed the great success of nanoscience and nanotechnology, yet ideal research model with well-defined composition and structure is still lacking. Specifically, most of the studied nanomaterials are quite polydisperse, that means, in most studied systems, the nano-scientists are extremely difficult to find two same nanoparticles with identical size, morphology, composition, and structure. To advance nanoscience and nanotechnology, one of the ultimate goals for the nano-scientists is to find a truly uniform system, in another word, atomically precise nanoparticles as research models [1, 2].
The emergence of atomically precise coinage metal (Au, Ag, Cu, and its alloy, etc.) nanoclusters can realize such goal [1]. Atomically precise metal nanocluster is a novel type of nanomaterial with the size in the sub-nanometer regime, normally 1–3 nm in diameter. It usually comprises 10–300 metal atoms with surface ligand as the protecting agents capping on the metal core. Note that as the nanoparticle size decreases downward to the sub-nanometer range, due to the strong quantum confinement effect, distinctly different physicochemical properties of metal nanoclusters from relatively large metal nanoparticles are observed [3]. For instance, discrete optical absorbance features can be readily identified in molecular Au nanoclusters, but such feature is absent in larger Au nanoparticle counterparts [4]. More importantly, the sub-nanometer size of the metal nanocluster is still within the resolving limitation of single crystal X-ray diffraction (SC-XRD), that allows nano-chemists to resolve their structure with atomic precision [5]. Such precise structure cannot be available for many other nanomaterials, which render metal nanoclusters unique advantages to comprehensively study the structure–property relationship in various fields, such as sensing [6, 7], assembly [8, 9], catalysis [10–12], optoelectronic [13, 14], and cancer therapy [15–17]. Furthermore, the chemical stability of these metal nanoclusters in terms of electronic structure can be explained by the “superatom” theory, where the electrons are confined in the spherical metal core of a jellium model [18]. It is believed that if the free electron number of a cluster is in good agreement of inert gas atoms (e.g., 2, 8, 18, 34, 52 electrons in the outmost orbital for He, Ne, Ar, Kr, Xe), it can be considered as a superatom having robust stability [18]. Meanwhile, the thermodynamic stability of thiolate metal nanoclusters is associated with the energy balance between the adsorption strength of the ligand shell to the metal core and the cohesive energy of the metal core [19]. In addition, for the non-magic number metal nanoclusters, the Wang group developed a superatomic orbital splitting (SOS) theory to understand the electronic configuration, where the shape of the metal core is considered in determining the order of the group orbital levels [20].
It is worth noting that, compared with noble metal nanoclusters such as Au, Ag, Pd, and Pt, Cu nanoclusters possess some unique characteristics. First, Cu is more earth abundant hence can be more cost effective for large-scale production to prepare functional nanomaterials. Secondly, Cu possesses different valence states including 0, +1, +2, which render Cu atom with different spatial arrangement in the metal core. So far, superatomic Cu nanoclusters with Cu(0) core, Cu(I) clusters, and Cu(II) complexes have been extensively reported [21, 22]. It is worth noting that different charge states impart Cu nanomaterial with drastically different physicochemical properties and reactivities, and the variation of the valence state during some catalytic reaction has also been observed [23–25]. Finally, Cu nanoclusters hold some peculiar functionalities that are not available from other noble metal nanoclusters [26]. For instance, Cu has good adsorption for CO2 than hydrogen, and it has strong capability to construct C–C bond, so in electrochemical CO2 reduction, using Cu nanoclusters can suppress the hydrogen evolution reaction and obtain HCO2H, CH3OH, CH, and even more valuable C2+ products such as C2H4 and C2H5OH [27], while only CO can be acquired when using Au or Ag nanoclusters as catalysts in most cases [28–34]. Table 1 summarizes all the Cu nanoclusters which have been discussed in this paper.
Table 1.
Formulas | Cluster configuration | Catalytic reaction | Refs. |
---|---|---|---|
[Cu6(MBD)6] | Octahedron | CO2RR | [27] |
[Cu29Cl4H22(Ph2phen)12]Cl | Cu13@Cu16 | – | [35] |
[Cu20(C≡CPh)12(OAc)6)] | Cu4@Cu16 | [3 + 2] Cycloaddition | [36] |
[Cu53(C≡CPhPh)9(dppp)6Cl3(NO3)9] | Layer-structure | – | [37] |
[Cu26(DPPE)3(CF3CO2)8(CH3O)2(tBuC≡C)4H11]+ | Cu4@Cu6@Cu16 | CO2RR | [38] |
[Cu32(PET)24H8Cl2](PPh4)2 | Cu18@Cu4PET2@ Cu14(PET)22Cl2 | Carbonylation of aniline | [39] |
[Cu25H22(PPh3)12]Cl | Cu13@Cu12 | – | [40] |
[Cu18H17(PPh3)10]Cl | Cu8@Cu10 | – | [41] |
[Cu18H3(S-Adm)12(PPh3)4Cl2] | Cu10@Cu8S12P4 | – | [42] |
[Cu31(4-MeO-PhC≡C)21(dppe)3](ClO4)2 | Cu13@Cu18 | – | [43] |
[Cu6(4-MeO-PhC≡C)5(dppe)3](ClO4) | Trigonal-prismatic configuration | – | [44] |
[Cu50(CF3COO)12(3, 5-diMe-PhS)18(PPh3)4H2] | Cu44@Cu6 | – | [45] |
[Cu41Cl2(2-F-C6H4S)12(CF3COO)6(PPh3)6H19]2− | Cu29@Cu12 | p-Nitrophenol reduction | [46] |
[Cu41(2, 5-di-Methyl-C6H3S)12(BO3)3Cl3(PPh3)6H19] | Cu29@Cu12 | p-Nitrophenol reduction | [46] |
[Cu23(tBuC≡C)13(CF3COO)6] | Cu4@Cu19 | – | [47] |
[Cu23(tBuC≡C)13(CF3COO)6]·CHCl3 | Cu4@Cu19 | – | [47] |
[Cu53(RCOO)10(C≡CtBu)20Cl2H18]+ | Cu3@Cu10Cl2@Cu20@Cu20 | – | [48] |
[Cu13Na2(CZ-PrA)6(TC4A)2Cl(CH3OH)2] | Cu9@Cu4 | Sulfide oxidation | [49] |
[Cu13Na(CZ-PrA)6(TC4A)2(CH3OH)]·CH3OH·CH2Cl2·CH3COCH3 | Cu9@Cu4Cl | Sulfide oxidation | [49] |
[Cu13(S2CNnBu2)6(C≡CR)4](PF6) | Cu@Cu12 cuboctahedron | – | [50] |
[Cu32H20(S2P(OiPr)2)12] | Cu14@Cu9@Cu9 | CO2RR | [51] |
[Cu8(H)(L1)6PF6] (L1 = 9H-carbazole-9-carbodithioate) |
Cu4 double tetrahedron | CO2RR | [52] |
[Cu8(tBuS)4(L1)4] (L1 = 9H-carbazole-9-carbodithioate |
Cu4 double tetrahedron | CO2RR | [52] |
[Cu8(tBuS)4(L2)4] (L2 = O-ethyl carbonodithiolate) | Cu4 double tetrahedron | CO2RR | [52] |
[Cu13(C4B10H11)10(PPh3)2(CH3CN)2](PF6)3 | – | Nitrate reduction | [53] |
[Cu26(C4B10H11)16(C2B9H10C2)2(PPh3)2(CH3CN)4](PF6)4 | – | Nitrate reduction | [53] |
[Cu6(HL1)2(L1)4(PF6)2] | Cu6 octahedron | CO2RR | [54] |
[Cu61(StBu)26S6Cl6H14] | Cu19@Cu42(StBu)26S6Cl6 | – | [55] |
[Cu57H20(PET)36(TPP)4]+ | Cu14@Cu43S36 | – | [56] |
[Cu58H20PET36(PPh3)4]2+ |
Cu8@Cu6@Cu24@Cu12 @Cu8 |
[3 + 2] Cycloaddition | [57] |
[Cu57H20PET36(PPh3)4]+ |
Cu8@Cu6@Cu24@Cu12 @Cu7 |
[3 + 2] Cycloaddition | [57] |
[Cu28H10(C7H7S)18(TPP)3] | Cu13@Cu15S18P3 | C−C coupling reaction | [58] |
[Cu7(SC5H9)7(PPh3)3] | – | C−C coupling reaction | [59] |
[Cu3(NHC)3(PF6)3] | – | A3 coupling and Redox-A3 coupling reaction | [60] |
[Cu8(Tf-dpf)4(NO3)2](NO3)2 | Two thin slices of supramolecular layers | “Aldehyde–acetylene–amine” A3 coupling reaction | [61] |
[Cu8(L2)2(L3)2](L2, L3 = deprotonated pyrazole) | – | Synthesis of indolizine | [62] |
Cu18H(PET)14(PPh3)6(isothiocyanate)3 | Cu15@Cu3 | Reduction of ferricyanide to ferrocyanide | [63] |
[Cu66Cl8(PPh3)8(SC2H5)32H24](SbF6)2 | Layered arrangement | Hydrogenation of cyclohexanone | [64] |
[Cu20H9(Tf-dpf)10]·BF4 | Cu12@2Cu4 | Conjugate reduction of cinnamaldehyde | [65] |
[Cu20H8(Tf-dpf)10]·(BF4)2 | Cu12@2Cu4 | Conjugate reduction of cinnamaldehyde | [65] |
Se@Cu20(PhSe)12(PPh3)2(C6H5COO)6 | Se@Cu6@Cu3@Cu3@Cu3@Cu3@Cu2 | 4-Nitrophenol reduction | [66] |
Se@Cu20(PhSe)12(PPh3)2(CF3COO)6 | Se@Cu6@Cu3@Cu3@Cu3@Cu3@Cu2 | 4-Nitrophenol reduction | [66] |
– denotes none or not available
Cu–Ligand Coordination Modes in Cu Nanoclusters
To prevent the Cu atoms from agglomeration or aggregation, surface capping ligand is critical as the stabilizing agent to allow Cu atoms to form a certain small core but not into a bulky core or an agglomerate. Furthermore, the ligand can significantly affect the physical and chemical properties of Cu nanoclusters through binding with the surface Cu atoms [67–69]. These interfacial coordination moieties form the ligament between the metal core and the surface ligands, and such ligament governs the electron coupling behaviors. So far, various types of molecules have been employed as capping ligand to stabilize the Cu core, and four typical widely employed organic molecules are thiolate [1, 3], phosphine containing molecule [70], nitrogen containing molecule [71], and alkynyl ligand [72, 73], yet they have markedly different coordination modes to bind the Cu atoms (Scheme 1). As illustrated in Scheme 1a, one S atom can coordinate with one, two, three, and four Cu atoms [39], and one Cu atom can coordinate with one, two, three, and four S atoms [74, 75]. One Cu atom can coordinate with one or two P atoms [40, 41], and one P atom can coordinate with three Cu atoms (Scheme 1b). Interestingly, three Cu atoms and three P atoms can form a cyclic triangle coordination mode [76]. The coordination mode between N and Cu is somewhat similar to that between P and Cu, as one Cu atom can coordinate with one or two P atoms, and both N and P atoms can form the dimer mode of R–N(P)–Cu (Scheme 1c) [21, 35, 77, 78]. However, unlike P, one N atom cannot coordinate with two or more Cu atoms, but it can form the trimer mode of R–N–Cu, and a cyclic quadrangle mode of Cu4N4 is also available [77]. It is worth noting that, for S, N, P atoms, they coordinate with the Cu atom mainly through σ bonding, nevertheless, for alkynyl ligand, it can coordinate with Cu atom with either σ bonding or π bonding or both. As shown in Scheme 1d, one alkynyl ligand can not only coordinate with three or four Cu atoms with only σ bonding, but also can coordinate with two, three, or four Cu atoms with both σ bonding and π bonding [36, 37, 79, 80]. There are seven essential types of coordination modes: μ2-ƞ1, ƞ2; μ3-ƞ1, ƞ1, ƞ1; μ3-ƞ1, ƞ1, ƞ2; μ3-ƞ1, ƞ2, ƞ2; μ4-ƞ1, ƞ1, ƞ1, ƞ1; μ4-ƞ1, ƞ1, ƞ1, ƞ2; μ4-ƞ1, ƞ1, ƞ2, ƞ2. Such unique binding modes can impart alkynyl-protected Cu nanoclusters with some drastically different physicochemical properties and functionalities from thiolate, nitrogen, and phosphine ligand stabilized Cu nanoclusters [72, 73].
Synthesis of Atomically Precise Cu Nanoclusters
There are various methods to synthesize bulky metal nanoparticles. However, preparing metal nanoclusters with atomic precision is still quite challenging. This is because metal nanoclusters with molecular purity typically form under specific thermodynamic or kinetic conditions. Since the pioneering Brust method to prepare thiolate-protected Au nanoclusters reported at 1990s [81, 82], a number of methods have been developed to prepare Au nanoclusters or Ag nanoclusters protected by thiolate, alkynyl molecule, and other ligands. The typical approaches include, but are not limited to, the direct reduction of the precursor, one-pot strategy, bi-phase method, ligand exchange or etching and so on [4, 83–87]. It is worth noting that, to prepare monodisperse metal nanoclusters, there are a lot of factors influencing the output, e.g., the reaction temperature, the solvent, the reducing agent, the nature and chemical property of the ligand molecule, the stoichiometric ratio of ligand–reductant–reactant, and so on. For instance, our group developed a synchronous nucleation and passivation strategy to synthesize alkynyl-protected coinage metal nanoclusters [88], and so far, several Au [88, 89], Ag [29, 32], AuAg [90–92], AgPd [93], AgRh [94, 95] nanoclusters have been successfully fabricated by this method [73, 96], but this approach is not applicable to thiolate-protected metal nanoclusters. Meanwhile, no successful case has been achieved on Cu nanoclusters yet. It also implies that, the approaches developed for synthesizing molecular thiolate-protected metal nanoclusters or alkynyl-protected metal (and alloy but not pure Cu) nanoclusters are not applicable for preparing atomically precise Cu nanoclusters, probably due to the different metal–ligand interactions can alter the nuclei growth and surface passivation behaviors [97].
Nevertheless, several generic methods have been developed for synthesizing monodisperses or molecular Cu nanoclusters. It includes the precursor reduction method, gradient reduction strategy (GRS), one-pot synthesis (OPS), ligand-exchange-induced growth, and other ingenious methods (Fig. 1). Each method has its own advantages or disadvantages. For example, the precursor reduction method is quite straightforward, but polydisperse cluster product may be acquired; the gradient reduction strategy features two step reduction with manipulation accessibility at each step, but the overall yield might be quite low; the one-pot synthesis holds facile operation but complex product is highly possible; the other methods are more applicable to special ligand to target specific cluster molecule. The advantages and disadvantages of these methods with specific cases will be elaborated next.
Precursor Reduction Method
The simplest and most straightforward method to fabricate Cu nanoclusters is to form the precursor first and then reduce it. However, under most circumstances, polydispersed Cu nanoclusters with a wide size distribution are acquired. To improve the monodispersity, kinetic control is inevitable. One typical example is the synthesis of [Cu18H3(S-Adm)12(PPh3)4Cl2] molecule reported by Mandal group [42]. As shown in Scheme 2, Cu(CH3CN)4BF4 was treated with auxiliary PPh3 ligand first, and then the adamantanethiol (Adm-SH) ligand was added to form the Cu(I) complex. Upon the addition of the NaBH4 methanol solution, the mixture changed from colorless into red, indicating the Cu18 nanocluster was formed. Note that, the employment of the bulky Adm-SH ligand is critical for yielding the Cu18 nanocluster here, as it acts as the main surface protecting ligand to stabilize the framework of the Cu18 nanocluster [42]. The Cu(I) precursor, also called as the Cu(I) complex, can be prepared step by step. In 2020, Li et al. reported a copper hydride cluster of [Cu32(PET)24H8Cl2](PPh4)2 (PET = phenylethyl thiolate) [39]. For its synthesis, Cu(TMEDA)Cl (TMEDA: tetramethylethylenediamine) was first obtained by reacting CuCl with TMEDA. Upon addition of PPh4Br and the thiol ligand (PETH), the final stage precursor of CuPET(TMEDA)Cl was obtained. The reduction of CuPET(TMEDA)Cl by sodium borohydride can yield Cu32 nanoclusters [39]. In 2023, Jia et al. utilized this method to fabricate an eight-electron superatom of [Cu31(4-MeO-PhC≡C)21(dppe)3](ClO4)2 (Cu31, dppe = 1, 2-bis(diphenylphosphino)ethane) cluster [43]. Specifically, Cu(ClO4)2·6H2O reacts with 4-methoxyphenylacetylene and dppe in the presence of triethylamine to form the precursor first, then Cu31 nanoclusters were obtained upon the reduction of borohydride [43]. Interestingly, side product of [Cu6(4-MeO-PhC≡C)5(dppe)3](ClO4) (it is a Cu(I) cluster not a superatom) was also isolated from the obtained single crystals of the Cu31 nanoclusters [44]. Recently, Fang, Wei and Shen et al. documented the fabrication of four Cu50 clusters with nearly identical metal frameworks [45]. Using Cu50(CF3COO)12(3, 5-diMe-PhS)18(PPh3)4H2 (Cu50-1) as an example, Cu(CF3COO)2 was first prepared as the Cu source, and after it reacted with 3, 5-dimethylbenzenethiol and PPh3, NaBH4 was added as a reducing agent to obtain the raw cluster product [45].
It is worth noting that, after forming the raw product by the precursor reduction method, adding some auxiliary ligand to etch the clusters may help to improve the yield and stability of the final product. For instance, Shen and coworkers recently reported a couple of “isostructural” Cu clusters of [Cu41Cl2(2-F-C6H4S)12(CF3COO)6(PPh3)6H19]2− and [Cu41(2, 5-di-Methyl-C6H3S)12(BO3)3Cl3(PPh3)6H19] [46]. The raw product was formed by reacting Cu(CF3COO)2 with 2-fluorothiophenol and PPh3 ligand, but after that, additional PPh3Cl ligand was added to introduce Cl onto the two Cu41 clusters for enhanced stability [46].
Gradient Reduction Strategy (GRS)
The gradient reduction strategy was first employed by the Zheng group to synthesize Cu53 nanoclusters, but the method was not named then in 2019 [48]. Specifically, Cu(CF3COO)2 was first reduced by Cu powder, after the tert-butylacetylene (HC≡CtBu) ligand was added, PhSiH2 as the second reducing agent was introduced into the solution under vigorous stirring. It is believed that the cooperation between Cu powder and PhSiH2 contributed together to achieve the high-nuclearity Cu(I)/Cu(0) cluster of [Cu53(RCOO)10(C≡CtBu)20Cl2H18]+ [48]. In another study, the synthesis and overall structure of Cu53(C≡CPhPh)9(dppp)6Cl3(NO3)9 nanocluster was recorded by Li and Zhang group. In a typical trial, 4-ethynylbiphenyl (BP) and 1, 3-bis(diphenylphosphino)propane (dppp) were sequentially added into an ethanol solution of Cu(NO3)2, then NaBH4 was introduced, followed by aging the reaction to obtain the target cluster [37].
In 2020, the name of gradient reduction strategy was formally proposed by Sun group [47]. In this method, the valence of the Cu atom evolved from +2 to +1, then to 0 by using different reducing agents at different stages. As illustrated in Scheme 3, the comproportionation of Cu(II) and Cu powder can yield the Cu(I) complex first, then the Cu(I) complex was further reduced to generate [Cu23(tBuC≡C)13(CF3COO)6] (SD/Cu23a) or [Cu23(tBuC≡C)13(CF3COO)6]·CHCl3 (SD/Cu23b). Such method showed the solvent-dependent polymorphism. It is worth noting that the Cu23 nanoclusters are superatoms, not Cu(I) complex, and both of them have four valent electrons. They also contain a very rare [Cu4]0 kernel surrounded by an outer Cu19 shell. In addition, depending on the solvent, the Cu23 nanocluster can crystallize into two polymorphs [47]. In 2022, the same group conducted the synthesis of two quasi-structurally isomeric 13-nuclei Cu nanoclusters (Cu13a and Cu13b) by the GRS method in a similar manner [49]. In this study, after the reaction of Cu(II) and CZ-PrAH (9-(prop-2-yn-1-yl)-9H-carbazole) with Cu powder to yield a yellow Cu(I) intermediate, the ligand was added into the mixture, following that a freshly prepared ethanol solution of NaBH4 was added under vigorous stirring. The reducing agent of NaBH4 with strong reducing capability was employed to avoid the re-oxidation of Cu(I) species [49]. Recently, Li et al. reported the comprehensive characterization and electrocatalytic CO2 reduction of [Cu26(DPPE)3(CF3CO2)8(CH3O)2(tBuC≡C)4H11]+, which was also synthesized by this GRS approach [38].
One-Pot Synthesis (OPS)
Compared to the above two methods, one-pot synthesis is the most facile method to operate, as it does not require to form the Cu(I) precursor. Basically, all the reactants including the Cu salt, the ligand molecules, the reducing agent and the base are mixed to generate the Cu nanoclusters. Of course, the Cu(I) precursor is formed in the reaction system, but no isolation is necessary. It is known that such method has been employed for synthesizing quite a number of coinage metal nanoclusters, e.g., thiolate Au25 nanoclusters can be easily prepared by one-pot approach [98, 99], the Tsukuda group fabricated a series of alkynyl-protected Au22 clusters by this method [100], and the Wang group reported the preparation of Au23(C≡CR)15 through one-pot synthesis as well [101].
It is worth noting that making Cu superatom cluster is rather different from making Au and Ag superatoms, as Cu(I) is more resistant to reduction. Therefore, in the presence of hydride source, a reduction can easily yield Au or Ag superatoms [2, 102], but Cu(I)-hydride complex is usually formed. It is more difficult to observe Cu(0) clusters, mainly due to the higher stability of Cu(I)-hydride complex than Au(I)-hydride or Ag(I)-hydride complexes. However, in a ligand-deficient environment, the reduction of Cu(I)-hydride complex might generate some unstable (CuH)x species that are amenable for cluster growth. In 2015, the Scott and Hayton group recorded the synthesis of [Cu25H22(PPh3)12]Cl and [Cu18H17(PPh3)10]Cl by this method [40]. As summarized in Scheme 4, adding 13 equiv. of Ph2SiH2 to a slurry containing 24 equiv. of Cu(OAc), 12 equiv. of PPh3, and 1 equiv. of CuCl in C6H6 can result in a rapid color change from pale green to dark red then into deep green, concomitant with the precipitation of a dark brown solid after keeping stirring in 24 h. Two clusters of [Cu25H22(PPh3)12]Cl and [Cu18H17(PPh3)10]Cl were isolated with a yield of 23% and 14%, respectively [40]. Following this study, the same group accomplished the synthesis of [Cu20(C≡CPh)12(OAc)6)] by the same approach [36].
Other Synthetic Methods
Besides the above approaches, researchers have developed several other ingenious methods to prepare atomically precise Cu nanoclusters. In 2016, the Hayton group reported a ligand-exchange-induced growth from a smaller cluster to synthesize Cu29 nanoclusters [35]. As shown in Scheme 5, [Cu25H22(PPh3)12]Cl was first synthesized, and then the addition of 16 equiv of 4, 7-diphenyl-1, 10-phenanthroline (Ph2phen) can cause an immediate color change from dark green to dark blue, and workup of this mixture for 15 min resulted in the isolation of [Cu29Cl4H22(Ph2phen)12]Cl, a deep blue black crystalline material in a yield of 84%. Meanwhile, the reaction of [Cu25H22(PPh3)12]Cl with 1, 10-phenanthroline also yielded a deep blue molecule, which is probably the isostructural molecule of [Cu29Cl4H22(Ph2phen)12]Cl, but no single crystal XRD data were available to confirm that [35].
Note that the introduction of additional ligand does not necessarily lead to cluster growth or size expansion; instead, cluster size decrease is also possible. Chakrahari et al. reported the [Cu13(S2CNnBu2)6(C≡CR)4](PF6) (R=C(O)OMe, C6H4F) cluster which was synthesized through size transformation (Scheme 6) [103]. As the hydride in [Cu28H15(S2CNnBu2)12]+ is substantially hydridic, the terminal alkyne is sufficiently acidic enough to react with it to yield [Cu13(S2CNnBu2)6(C≡CR)4]+, which features a centered cuboctahedral [Cu13]11+ core with two free electrons, and [Cu8H(S2CNnBu2)6(C≡CR)4]+ was also isolated as the side product [103].
Atomically Precise Cu Nanoclusters for Catalytic Applications
Thanks to the precise composition and structure of Cu nanoclusters, they can serve as ideal model catalysts for a variety of reactions [50]. Also, theoretical simulations can build precise models to reveal the reaction pathway and help to elucidate the reaction mechanism [104, 105]. In line with this, a great deal of research effort has been devoted to exploring the catalytic applications of atomically precise Cu nanoclusters, including but not limited to electrocatalytic reactions, photocatalytic reactions, organic reactions, as well as other catalytic processes (Fig. 2).
Cu NCs for Electrocatalysis
The electrocatalysis technology has drawn widespread attention and found immerse values in a wide spectrum of fields especially in green energy storage and conversion, as electrocatalysis is not only highly efficient, mild condition operational, multifunctional, but also can couple with renewable intermittent energy sources such as solar, wind, and geothermal energy [106]. Therefore, electrocatalysis represents a sustainable strategy to address global environmental pollution and heavy reliance on fossil fuels [107].
In the last decade, the fast industrialization and huge demand for fossil fuels have caused a dramatic worldwide increase of CO2 emission, which is the main culprit of global warming. However, CO2 is an abundant source that can be used as substrate to fabricate various valuable products such formic acid, formamide, urea, and other hydrocarbon compounds [108, 109]. Electrochemical CO2 reduction reaction (eCO2RR) can realize such conversion with high efficiency and in an environmentally friendly manner; however, due to the chemical inertness and thermodynamic stability of linear CO2 molecules, a prerequisite for achieving CO2 reduction lies in finding a catalyst that can lower the chemical energy required to break the C-O bond [110, 111]. Moreover, eCO2RR involves multiple electron transfer, and various products including C1 compounds (e.g., CO, HCO2H, CH3OH) and C2 compounds (e.g., CH3CO2H, CH3CHO, C2H4, C2H6, etc.) can be generated. The thermodynamic potentials, various reactions, and corresponding products are summarized in Table 2 [112]. It can be noted that different reactions occur at very similar thermodynamic equilibrium potentials, plus the competing hydrogen evolution reaction (HER), it is extremely challenging to achieve highly selective CO2 reduction product. Therefore, designing efficient, durable, and high-selectivity catalysts for eCO2RR to acquire target product is highly desired [113].
Table 2.
Reactions | E0 (V vs. RHE) | Product |
---|---|---|
CO2 + 2H+ + 2e− → CO (g) + H2O | − 0.11 | Carbon monoxide |
CO2 + 2H+ + 2e− → HCOOH (aq) | − 0.12 | Formic acid |
CO2 + 6H+ + 6e− → CH3OH (aq) + H2O | 0.03 | Methanol |
CO2 + 8H+ + 8e− → CH4 (g) + 2H2O | 0.17 | Methane |
2CO2 + 10H+ + 10e− → CH3CHO (aq) + 3H2O | 0.06 | Acetaldehyde |
2CO2 + 12H+ + 12e− → C2H4 (aq) + 4H2O | 0.08 | Ethylene |
2CO2 + 12H+ + 12e− → C2H5OH (aq) + 3H2O | 0.09 | Ethanol |
2CO2 + 14H+ + 14e− → C2H6 (aq) + 4H2O | 0.14 | Ethane |
2CO2 + 16H+ + 16e− → C2H5CHO (aq) + 5H2O | 0.09 | Propionaldehyde |
3CO2 + 18H+ + 18e− → C3H7OH (aq) + 5H2O | 0.1 | Propanol |
2H+ + 2e− → H2 | 0 | Hydrogen evolution reaction |
2H2O → O2 + 4H+ + 4e− | 1.23 | Oxygen evolution reaction |
So far, all kinds of nanostructured materials including metal oxides, metal alloys, carbon substrates, two-dimensional nanomaterials have been investigated as catalysts for eCO2RR [114–118]. Among a series of metal-based catalysts, Cu nanomaterials are one of the most promising catalysts that can deep electrochemically reduce CO2 into valuable chemicals, especially hydrocarbon products [119–122]. However, for most Cu-based catalysts, the chemical nature is not uniform, e.g., no homogeneous size, shape, architecture, and structure hence it is extremely challenging to ascertain the chemical nature of the catalyst and eventually establish the structure–activity relationship [123–125]. Recently, molecular Cu nanoclusters have been drawing attention from the heterogeneous catalysis field, thanks to the high surface-to-volume ratio, strong binding capability to the key reaction intermediates, and more importantly, the atomically precise crystallographic structure and well-defined architecture [50, 90, 126].
In 2017, Tang et al. reported a structurally precise Cu–hydride nanocluster of Cu32H20L12 (L is a dithiophosphate ligand), which offered unique selectivity for eCO2RR at low overpotentials [51]. By density functional theory calculations, the authors first predicted that the presence of negatively charged hydride in Cu nanocluster plays a crucial role in determining the selectivity of the products, yielding HCO2H over CO at lower overpotentials. The optimized structure of the Cu32H20L12 cluster is shown in Fig. 3a, and the Cu32 cluster has a distorted 14-Cu-atom formed hexacapped rhombohedral core sandwiched between two 9-Cu-atom formed triangular cupola fragments. Meanwhile, the 20 hydrides are divided into 12 triply coordinated H (µ3-H), 6 tetra-coordinated H (µ4-H), and 2 penta-coordinated H (µ5-H). For CO2 reduction into CO on the Cu32H20L12 cluster, it adopts either the proton-reduction channel or lattice-hydride mechanism (Fig. 3b). It can be clearly noted that the lattice hydride pathway is more favorable, where the rate-determining step is the μ3-H1 hydride being transferred to form the Cu32H19L12-HCOO intermediate with a free energy change of 0.32 eV (Fig. 3b right). In contrast, the formation of Cu32H19L12-HCOO in the proton reduction channel requires an energy gap of 1.08 eV (Fig. 3b left). Finally, the authors conducted the eCO2RR test of the Cu32H20L12 cluster to verify the theoretical predictions. As shown in Fig. 3c, the average current density becomes significant at the overpotential of 0.3 V and increases with the increasing of the overpotential. H2, CO, and HCOOH are the main products, and the cumulative Faradaic efficiency (FE) is over 90%. The product selectivity is illustrated in Fig. 3d. HCO2H is the predominant major product in the overpotential window from 0.3 to 0.4 V, and HER dominates at higher overpotentials (from 0.5 to 0.6 V) [51]. These experimental results confirmed the theoretical predictions, and this study showcases that the hydride-containing Cu nanoclusters may offer a unique product selectivity over conventional transition metal nanocatalysts for eCO2RR.
Furthermore, the core configuration of Cu nanoclusters can be manipulated to further mediate the activity and selectivity of eCO2RR. In 2022, the Zang and Wang group reported three novel isomeric Cu8-cluster with cores composed of two types of metal kernels (ditetrahedral vs. cubic) and studied the morphological kernel influence on the electrochemical eCO2RR at the atomic level [52]. Specifically, three Cu clusters of Cu8(H)-(L1)6PF6, Cu8(tBuS)4(L1)4, and Cu8(tBuS)4(L2)4 (Cu8-1, Cu8-2, and Cu8-3, respectively, where L1 = 9H-carbazole-9-carbodithioate and L2 = O-ethyl carbonodithiolate) were prepared, and the total structures are illustrated in Fig. 4a–c. Cu8-1 contains a slightly twisted cubic Cu88+ core, while Cu8-2 and Cu8-3 show the identical di-tetrahedral configurations. In eCO2RR test, H2, CO, and HCO2H were the main products, and the total FE values were over 90% (Fig. 4d). Moreover, the di-tetrahedron-shaped Cu8 clusters (Cu8-2 and Cu8-3) exhibited a higher FEHCOH2 and greater selectivity than the cube-shaped Cu8 cluster (Cu8-1), where Cu8-2 demonstrated remarkable HCO2H selectivity, manifested by the FEHCO2H values of 90% and 92% at −0.9 and −1.0 V, respectively (Fig. 4d). Meanwhile, in the long-term stability test of 8 h, only the current density from Cu8-2 remained almost unchanged, indicating robust durability for prolonged operation (Fig. 4e). At last, the free energies of each step for Cu8-1 and Cu8-2 were calculated. Noteworthily, for the adsorbed *CO2 being converted into HCOO*, the di-tetrahedron-shaped Cu8-2 cluster (−0.91 eV) has a much lower free energy than the cube-shaped Cu8-1 cluster (−0.23 eV). Since the rate determining step is the formation of COOH* to generate HCO2H, the Cu8-2 cluster is more favorable for HCO2H production [52]. This investigation showcases the modulation of activity and selectivity toward CO2 electroreduction by tailoring the Cu core, which can probably trigger more endeavors to design tailored Cu based catalysts for eCO2RR.
It is worth noting that for most reported atomically precise Cu nanoclusters in the eCO2RR, C1 products of CO and HCO2H with high selectivity are normally obtained. This is probably due to the Cu active sites are coordination symmetric (typically CuS3). Compared to the C1 products of CO/HCO2H, the high energy density hydrocarbon products such as CH4 and C2H4 are highly sought as potential fuels [127–130]. To switch the C1 products to more valuable hydrocarbons, Wu et al. recently reported a catalyst based Cu6 nanocluster with symmetry-broken CuS2N1 active sites [27]. The total structure of the as-prepared Cu6(MBD)6 nanocluster (MBD = 2-mercaptobenzimidazole) is shown in Fig. 5a. Each MBD ligand has the tridentate sites to coordinate three Cu atoms, where the thiolate S atom binds with two Cu atoms and the N atom binds to the other Cu atom (Fig. 5b). Meanwhile, each Cu atom is coordinated by two S atoms and one N atom (Fig. 5c). It is symmetry-broken distorted CuS2N1 geometry, as the bonding length values are all different (Cu1−S1 = 2.2118 A, Cu1−S1a = 2.2924 A, Cu1−N1 = 1.9757 A). In addition, six Cu atoms formed a distorted octahedron with the Cu–Cu bonding ranging from 3.083 to 3.254 Å, which might be favorable for C–C coupling (Fig. 5d). There is sufficient space above the distorted CuS2N1 geometry, which can be accessible for CO2 adsorption (Fig. 5e). When using Cu6(MBD)6 nanocluster as the catalyst, the product distribution is shown in Fig. 5f. When the potential goes from −0.7 to −1.4 V, the FECO and FEH2 gradually decreased; meanwhile, the FEC2H4 and FEC2H4 gradually increased (Fig. 5f). The highest FECH4 reached 42.5% at −1.4 V with a large partial current density of −119.0 mA cm−2, with the FEC2H4 of 23.0% and the partial current density of −64.4 mA cm−2 (Fig. 5g). That is, in a wide potential window, CH4 and C2H4 are the dominant products rather than CO. To gain more insights into the advantages of symmetry-breaking at the catalytic sites, in situ Cu K-edge XANES measurement of Cu6(MBD)6 was conducted, revealing that there is clear electron transfer from Cu6(MBD)6 to the CO2 molecule. Thus, the Cu6(MBD)6 cluster with asymmetric CuS2N1 sites has a better CO2 activation capacity than the Cu8(tBuS)4(L2)4 cluster with the symmetric CuS3 sites. The integrated projected density of states (IPDOS) of Cu sites in Cu6(MBD)6 and Cu8(tBuS)4(L2)4 were then calculated. For Cu6(MBD)6, the dx2-y2 orbital of Cu-S2N1 site has the highest number of electrons close to the Fermi energy level (Fig. 5h), whereas the highest d orbital is Cu dxz in CuS3 sites of Cu8(tBuS)4(L2)4 (Fig. 5i). It suggests that the coordination-symmetry breaking significantly affects the highest occupied d-orbital, which plays a critical role in regulating the coordination mode of the reactant and intermediate. When the C atom of CO2 is adsorbed onto the Cu6(MBD)6, the highest occupied dx2-y2 orbital can energy-match with the lowest occupied π* orbital of CO2 to form a π-complex rather than the conventional σ-complex, where the π-complex can lower the energy barrier of CO2 activation. If the O atom of CO2 is adsorbed onto the Cu6(MBD)6, a σ-complex is formed, and for Cu8(tBuS)4(L2)4, a π-complex is formed only when the O atom in CO2 is the adsorbed atom. Consequently, due to superiority of the C atom adsorbing on Cu6(MBD)6, the hydrogenation on the less hindered O atom to form the key *COOH intermediate rather than the *OCHO intermediate is more favorable, making the reaction proceed to yield the CH4/C2H4 products [27]. This study highlights the importance of regulating the coordination mode of Cu nanoclusters, which might be a promising direction for designing metal nanoclusters as highly efficient CO2 reduction electrocatalysts toward valuable hydrocarbon production.
The large surface area, good catalytic activity and strong capability to suppress HER also enable Cu nanoclusters as potent catalysts for electrochemical nitrate reduction reaction (eNO3RR) [131, 132], a process that can turn nitrate contaminant into valuable NH3 product. Specifically, NH3 is an important feedstock for preparing nitrogen-containing fertilizer, pharmaceuticals, chemicals, and agricultural products, and it is also a valuable green energy carrier [133, 134]. However, the current global production of NH3 is largely relying on the energy-intensive and large CO2-emissive Haber–Bosch process. The electrochemical N2 reduction powered by renewable energy offers a promising approach but it is significantly restricted by the low solubility of N2 in water and the extremely high energy required to break the N≡N bond (941 kJ mol−1). In stark contrast, nitrate is one of the major contaminants in wastewater and also has much weaker N–O bonding energy (250 kJ mol−1). Therefore, developing efficient and stable electrocatalysts for NO3RR is imperative for offering a sustainable strategy to produce NH3 [135–138]. Recently, atomically precise metal nanoclusters have demonstrated great potential for catalyzing eNO3RR, mainly thanks to the strong catalytic capability, and more importantly, the atomic precise structure can provide an ideal model to elucidate the complicated reaction pathway and establish the structure–activity correlation [131, 132, 139]. For instance, our group found that Ag30Pd4(C6H9)26](BPh4)2 nanocluster can achieve the Faradaic efficiency of NH3 over 90% in eNO3RR, where the Ag site is responsible for converting NO3− into NO2− and the Pd site makes the major contribution to catalyze NO2− into NH3 hence the whole reaction adopted a tandem catalytic mechanism [93].
Recently, Wang and Zang et al. utilized bulky carborane-alkynyl ligand to prepare atom-precise monomer Cu13·3PF6 and bridged dimer Cu26·4PF6 clusters, and both clusters exhibited remarkable catalytic activity and selectivity in eNO3RR [53]. As illustrated in Fig. 6a, Cu13·3PF6 has a metal skeleton, which could be viewed as two pentagonal bipyramids merged by sharing one equatorial edge but with one vertex of the pentagonal bipyramid being lost. Cu26·4PF6 is a dimer of Cu13·3PF6. To be specific, in Cu26·4PF6, each Cu13 monomer loses a carbocycloalkynyl ligand and the adjacent PPh3, and the vacated space is occupied by the nido-carboranealkynyl ligands unit of the other monomer via cyclopentadienyl anionic coordination (Fig. 6b). Interestingly, both clusters have fair stability and accessible open metal sites for eNO3RR, where eNO3RR was carried out in an H-type cell containing 0.1 M KNO3 in 0.5 M K2SO4 medium (Fig. 6c). When the applied potential goes more negatively, the FENH3 exhibited a volcano-shape change, and both clusters reached the maximal value at -0.85 V (vs. RHE). The maximal FENH3 of Cu26·4PF6 is 85.1%, much higher than that of Cu13·3PF6. Meanwhile, at each measured potential, the FE and yield rate of NH3 for Cu26·4PF6 surpass that of Cu13·3PF6 (Fig. 6d, e), despite the NH3 yield rate increasing with the increasing of the applied potential (Fig. 6e). Finally, the authors elucidated the reaction pathway and mechanism with the aid of theoretical simulations and in-situ FTIR spectroscopic study. The free energy diagram in Fig. 6f suggests that Cu13·3PF6 and Cu26·4PF6 share the identical reaction pathway (*NO3 → *NO2 → *NO → *NHO → *NH2O → *NH2OH → *NH2 → *NH3). For both clusters, the rate determining step (RDS) is identified as *NHO-to-*NH2O, and the energy consumption for Cu26·4PF6 is much lower than that of Cu13·3PF6 (0.27 vs. 0.58 eV), indicating Cu26·4PF6 is more energetically favorable for eNO3RR. Moreover, the subsequent *NH2O-to-*NH3 reduction step is also a downhill process, demonstrating a strong interaction between the key intermediates and the active sites, which can promote the electron transfer and the following hydrogenation steps to accelerate the eNO3RR process. This study not only establishes a platform to disclose the structure–activity relationship of Cu nanoclusters for eNO3RR, but also provides a feasible strategy to access desirable Cu nanoclusters as efficient eNO3RR catalysts [53].
Cu NCs for Photocatalysis
Solar energy is an inexhaustible natural energy source, and photocatalytic technology can utilize solar energy to realize environmental decontamination and energy conversion, particularly, photocatalytic technology can reduce CO2 into valuable hydrocarbon fuels [140, 141]. The principle of photocatalytic reduction of CO2 is to use photoexcited semiconductor photocatalysts to produce photogenerated electrons and holes, and the carriers take redox reactions on the surface of the catalyst, reducing CO2 into useful chemical raw materials [142, 143]. Compared with other methods, the photocatalytic reduction approach has some unique merits. On one hand, photocatalytic method directly uses solar energy for CO2 reduction and transformation, which is safe and pollution-free. On the other hand, the device for photocatalytic reduction is quite simple, and it can operate at very mild conditions, so the whole process is economically feasible [144–147].
Among a variety of photocatalysts, the artificial photosynthesis system using inexhaustible solar energy to simultaneously reduce CO2 and oxidize H2O to produce valuable chemicals has been attracting more and more research attention [148, 149]. The current widely reported photocatalytic system usually comprises noble-metal-containing photosensitizers and/or organic dyes as electron donors, which are sophisticated and synthetically challenging to be available [150]. Therefore, it is highly desired to prepare a single component photocatalyst with suitable optical band and catalytic active center, that is similar to natural photocatalytic system, which can directly convert CO2 and H2O into value-added chemicals.
Metal nanoclusters, an emerging type of organic–inorganic hybrid material, composed of an organic layer and metal core, are promising candidates as a single component catalyst to realize the above goal [151, 152]. The ultrasmall size, well-defined configuration, and most importantly, the atomically precise structure can provide a perfect platform to probe the atomical level structure–performance understanding regarding the photocatalysis mechanism [153, 154].
Recently, the Zang group reported a stable Cu–S–N cluster photocatalyst with local protonated N–H groups, and such cluster can achieve ~ 100% selectivity for CO evolution under visible light. Specifically, two clusters of Cu6-NH and Cu6-N were prepared [54]. As shown in Fig. 7a, b, Cu6-NH cluster possesses a distorted Cu octahedron, and one Cu atom coordinates with one N atom and two S atoms from three ligands. Such configuration is quite identical with the previously reported Cu6-N cluster (Fig. 7c, d), except that the ligands in Cu6-N are fully deprotonated, despite the two clusters having the same metal kernel (Fig. 7e). Both clusters have considerable absorbance capacities in the wavelength region of 400–550 nm, and comparable band gap energies (2.36 and 2.39 eV for Cu6-N and Cu6-NH, respectively). However, their visible-light-driven photoreduction capability of CO2 showed some significant difference. Both clusters exhibited exceptionally high selectivity of CO with over 99%, but the CO evolution rate is much higher for Cu6-NH than that of Cu6-N (Fig. 7f). Furthermore, the Cu6-NH photocatalyst demonstrated superior catalytic stability, as no detectable change is observed in the XRD patterns before and after the photocatalytic test (Fig. 7g). Finally, they conducted DFT calculations to elucidate the underlying physical origin, and found that, in the rate-determining step of forming *COOH, the Cu6-NH cluster has a much lower energy barrier. This is due to the structural difference, that is, the presence of deprotonated pyrimidine N in Cu6-N and protonated pyrimidine N in Cu6-NH (Fig. 7h). The protonated pyrimidine N atoms in Cu6-NH acted as a proton relay station to provide a local proton, hence facilitating the proton coupling process, leading to enhanced photocatalytic efficiency [54]. This study highlights the great promise of using Cu nanoclusters as photosensitive semiconductors for photoreduction of CO2.
The precise structure of Cu nanoclusters allows the in-depth study of structure–property relationships in photocatalysis at the atomical level. Moreover, isomeric Cu nanoclusters can make the comparative structure–property relationship study feasible, as they have minimal structural difference. Isomeric Cu nanoclusters usually originate from the chiral isomerism, the structural isomerism, but the quasi-structural isomerism has been ignored for a long time. The quasi-structural isomeric Cu nanocluster usually displays comparative structural features, e.g., core geometric configuration, surface spatial arrangement.
In 2022, the Sun group reported two quasi-structurally isomeric Cu13 nanoclusters with highly similar kernel and different spatial arrangements of peripheral ligands (Fig. 8a) [49]. Both Cu13a and Cu13b clusters have a highly similar Cu13 kernel but different degrees of distortion (Fig. 8b, c). The formation of these two Cu cluster isomers is governed by the exotic chlorine ion, and also due to the charge transfer from Cl− to Cu core, Cu13a presented lower superoxide radical (O2•−) yield and higher singlet oxygen (1O2) compared to that of Cu13b. The conduction band minimum (CBM) is −1.20 and −1.26 V for Cu13a and Cu13b, respectively, so through the band gap value, the valence band maximum (VBM) is calculated as 0.74 and 0.66 V for Cu13a and Cu13b, respectively (Fig. 8d). More negative CBM value drives Cu13b more energetically feasible for O2•−, while for Cu13a, the larger band gap favors the exchange of an electron with 3O2, also more efficient for 1O2 generation (Fig. 8e). Consequently, the two clusters demonstrated different performance particularly in selectivity in sulfides oxidation into sulfoxides. As summarized in Fig. 8f, despite methyl sulfide can be transformed into dimethyl sulfoxide by two clusters with nearly identical high conversion rates (near unity), Cu13a showed a certainly higher degree of enhancement than that of Cu13b [49]. This work provides a new pathway for boosting the photocatalytic selectivity of Cu nanoclusters.
In modern synthesis, bond-forming reactions are critical for generating new products, particularly, bond-forming reactions at mild conditions can find versatile applications in fine chemical synthesis, new drug discovery, and sustainable production of the industrial precursor. To that end, visible light photocatalysis is more advantageous as compared to thermal catalysis, however, developing high-efficiency, stable catalyst is the key [155]. Cu-based photocatalysis has shown great promise as an inexpensive and attractive method, compared to the expensive metal-complex and less stable organic dyes [156, 157]. Nevertheless, the cross-coupling reactions of aryl bromides and iodides are relatively easy to realize, but the cross-coupling reactions of aryl chloride remained extremely challenging.
In 2022, the Bakr and Rueping team reported a [Cu61(StBu)26S6Cl6H14] nanocluster (Cu61NC)-based catalyst that can enable C-N bond-forming reaction of aryl chlorides under visible-light irradiation at ambient conditions [55]. As shown in Fig. 9a, the pronounced molecular ion is observed, and the experimental pattern agrees well with the simulated pattern. Figure 9b illustrates that it has two characteristic bands, and the crystal structure can be found inset. In the presence of base, at room temperature, the Cu61 NC cluster can catalyze the Ullman reaction for het(aryl) very well (Fig. 9c). The authors also proposed a plausible mechanism (Fig. 9d). Specifically, Cu NC reacts with the base forming CuNC-amine complex A first, then blue LED irradiation leads to the formation of complex B. After a SET process to give an electron to aryl chloride, the oxidized CuNC-Nu complex, a halide anion, and an arene radical were generated, evidenced by the fluorescence and quenching lifetime tests. After that, the radical attacks the CuNC-Nu complex, delivering the C–N bond formation and recovering the Cu61 clusters [55]. This study introduces atomically precise Cu clusters as a new family of photocatalysts for bond-forming reactions under mild conditions.
Cu NCs for Organic Reactions
Click Reaction
The perfect monodisperse and atomically precise nature of molecular metal nanoclusters can facilitate the development of detailed structure–activity relationships [158–160]. Using these atomically precise metal nanoclusters as catalysts can uncover new insights, but some of them may not be suited for this purpose. Specifically, a vast majority of these metal nanoclusters are protected by a shell of thiolate ligands [1, 152]. These ligands can impart the clusters significant thermal and chemical stability, but they can also block some active sites and hence must be partially removed before catalysis can occur [161]. Meanwhile, the strong bonding strength of the metal-sulfur bond makes it hard to remove completely [161]. However, the harsh reaction conditions often lead to the significant structure or nuclearity change of the metal nanoclusters, making the structure–activity relationship unreliable. To address this issue, the researchers switched to other ligands, which do not require harsh pretreatment for activation. To that end, alkynyl molecules, phosphine ligand, halogen, hydride, and their mixed combination have been widely utilized to prepare atomically precise metal nanoclusters including Cu nanoclusters [56, 72, 73, 162].
In 2018, the Hayton group reported the synthesis of [Cu20(C≡CPh)12(OAc)6] cluster and its catalytic application in Click reaction [36]. Figure 10a shows the crystal structure of the Cu20 cluster, which has 4 THF molecules incorporated into the crystal lattice as solvates. It has a tetrahedral [Cu4]2+ core (Fig. 10b), which is encapsulated by a [Cu16(C≡CPh)12(OAc)6]2− shell (Fig. 10c). There are 12 acetylide ligands in the [Cu16(C≡CPh)12(OAc)6]2− shell distributed by 4 [cyclo-Cu(C≡CPh)]3 units, which are located at the vertices of a tetrahedron, meanwhile, there are six acetate ligands situated at the edges of the tetrahedron (Fig. 10c). Then the Cu20 cluster was immobilized on dry, partially dehydroxylated silica, and the supported cluster catalyst displayed excellent performance toward the Click reaction of cycloadditions between benzyl azide and terminal alkynes. As demonstrated in Fig. 10d, the catalyst is effective for various terminal alkynes without harsh pre-treatment and even with a rather low cluster loading of 0.05 mol%. More importantly, X-ray absorption near edge structure (XANES) test after the catalytic test shows that the cluster undergoes no significant nuclearity changes under click reaction conditions [36]. It also indicates that the Cu species mainly kept as Cu(0) and Cu(I) in the cluster during the reaction process, where the Cu(I) species is well known to be the active site for click reaction. Great promises of using atomically precise Cu nanoclusters to examine the structure/activity relationship have been demonstrated in this study, especially for the non-thiolate protected Cu nanoclusters.
In nanocatalytic regime, elucidating single-atom effect of the nanocatalyst is interesting yet extremely challenging, as one single-atom modification can cause appreciable change to the overall particle’s structure, not to mention that, the size polydispersity and ambiguous surface structure of the metal nanoparticle [163]. Nevertheless, atomically precise metal nanoclusters have created great opportunities to tailor the chemical structure at the atomic level.
Recently, the Bakr and Rueping team reported the synthesis of [Cu58H20PET36(PPh3)4]2+ (Cu58) and its analogue of [Cu57H20PET36(PPh3)4]+ (Cu57) with one surface Cu atom removed, and Cu57 showed much higher catalytic activity toward [3 + 2] azide-alkyne cycloaddition (AAC, click reaction) than that of Cu58 [57]. As illustrated in Fig. 11a, Cu58 has a multishell architecture, and the metal skeleton possesses five concentric shells of Cu8@Cu6@Cu24@Cu12@Cu8. Meanwhile, the ligand layer in Cu58 consists of a P4 tetrahedron, an S12 icosahedron, and an S24 truncated cube. Cu57 has nearly the identical metal kernel structure except one Cu atom was removed on shell-4. Subsequently, the two nanoclusters were employed as catalyst for the click reaction. With the 0.025 mol% of Cu57 or Cu58 stoichiometric ratio, and in the presence of 1.5 h light irradiation, the product yields of 97% and 77% were achieved for Cu57 and Cu58, respectively (Fig. 11b). In addition, both clusters showed good catalytic stability, as no obvious product yield decline was observed after three catalytic cycles (Fig. 11c). Overall, Cu57 performed more efficiently than Cu58 in the click reaction, and one single atom removal made great difference in the nanoclusters’ catalytic activity [57]. This work may open a new avenue for the nanoparticles’ catalytic design by targeted isostructural single-atom manipulation.
C–C Coupling Reaction
The C–C bond formation is a widely employed transformation in natural product synthesis, pharmaceutical synthesis, medicinal chemistry, and preparation of functional materials. So far, various catalytic methods have been developed for this important cross-coupling reaction. Among that, Sonogashira cross-coupling reaction is one of the most important and widespread used sp2–sp2 carbon–carbon bond formation reactions in organic synthesis, which is usually catalyzed by Pd and other transition metals [157, 164]. It can realize the coupling of aryl or vinyl halides with terminal acetylenes, but most of the catalysts suffer from that the undesired alkyne homocoupling side reaction is extremely difficult to suppress.
In 2023, the Rueping and Bakr group reported the [Cu28H10(C7H7S)18(TPP)3] cluster (Cu28; C7H7S: o-thiocresol; TPP: triphenylphosphine) with a defined defect for Sonogashira C–C coupling reaction. The skeleton structure of Cu28 is shown in Fig. 12a, and it features a centered Cu13 anti-cuboctahedron core, like the Cu25, Ag19, and Ag25 nanoclusters, and Cu28 also has a unique cage-like Cu15S18P3 shell. Interestingly, Cu28 is a defective nanocluster which contains a surface vacancy, as one Cu atom is missing in one tetrahedron’s vertex, and this vertex is expected to connect with a phosphine ligand. Subsequently, Cu28 was examined as a catalyst for Sonogashira C–C coupling reaction. Upon UV-LEDs irradiation, the desired C–C coupling product with a high yield of 82% was acquired without the byproduct of homocoupling reaction (Fig. 12b) [58]. The authors then conducted the mechanistic control studies and proposed the tentative reaction mechanism (Fig. 12c). Upon blue-light irradiation, Cu28 was photoexcited (A), and it can induce a single electron transfer (SET) process with aryl iodides to yield the Ar-radical (C) and oxidized Cu28+ (B). The Ar-radical then attacks the alkyne C–C triple bond to generate vinylic type C-radical intermediate (D). Another single electron transfer process to B can regenerate the Cu28 photocatalyst. Note that the C–C triple bond can be activated by Cu28 through forming a Cu NC-π-alkynyl type complex hence easily being attacked by the Ar-radical. The authors detected a little stilbene-type byproduct in GC–MS measurement generated by a hydrogen atom transfer (HAT) process. Finally, the base-induced deprotonation of vinyl cation E leads to the desired Sonogashira C–C coupled product [58]. Upon the photoexcitation of blue-LEDs, the oxidation state of the Cu28 cluster undergoes a transient transition from 0 to I, form [Cu28]+ NC, while after the second single electron transfer (SET) process, it returns to the ground state of Cu28.The results demonstrate that Cu28 is an active photocatalyst enables the SET process with aryl iodides to produce Ar-radical which attacks the C–C triple bond, totally preventing the formation of homocoupling product. This investigation paves a way to acquire Cu NCs with defined surface defects as active sites, and more importantly, these defective Cu NCs can serve as model systems to provide profound understanding on defect effect for heterogeneous catalysis.
Recently, Biswas et al. reported a hydride-free [Cu7(SC5H9)7(PPh3)3] (Cu7) nanocluster (NC), which demonstrated remarkable specificity in a photoinduced C–C coupling reaction [59]. The Cu7 NC has a metal core with 4 Cu atoms forming a tetrahedron (Fig. 13a), and three Cu atoms are connected with the bottom face of the tetrahedron with another three Cu atoms (Fig. 13b). The metal–ligand binding configurations are shown in Fig. 13c, d, where the thiolate ligands mainly adopt the bridging modes, and three P atoms connected with three Cu atoms directly. Interestingly, the Cu7 NC exhibits intriguing photoluminescence (PL). Upon exciting the cluster at 395 nm, a robust PL emission is observed at room temperature with an emission peak occurring at 436 nm. Such outstanding photophysical properties inspired the authors to explore the Cu7 NC as catalyst for photoinduced C–C bond formation reaction. Under purple light irradiation from LEDs, the Cu7 NC (5 mol% Cu) was able to catalyze the Csp2-Csp coupling reaction between phenyl iodide and phenyl acetylene to generate the desired product with high yield (Fig. 13e). With the aid of control experiments and DFT calculations, the authors also proposed tentative reaction mechanism (Fig. 13f). Initially, the Cu active site in Cu7 NC binds to alkyne to form an intermediate of Int-I in the presence of MeONa. Subsequently, Int-I is photoexcited to generate Int-I*, which engages the oxidative addition step to yield the intermediate of Int-II. The following reductive elimination of Int-II produces the final product, along with the release of the Cu7 NC. In this mechanistic rationale cycle, the oxidative addition step is identified as the rate-determining step [59]. This work provides some mechanistic insights of using atomically precise Cu nanoclusters for C–C coupling reactions, and underscore the great potential of hydride-free Cu nanoclusters for catalysis.
A3 Coupling Reaction
To forge new C–C and C–N bonds, three-component dehydrogenative coupling reaction represents a new practical methodology. Yet, to realize all-in-one three-component dehydrogenative coupling in a single catalytic system remains very challenging [165]. For instance, to fulfill the efficient synthesis of propargylamines including C1-propargylamines, Cu compounds are the most widely recognized catalysts to implement the A3 coupling reactions or redox-A3 coupling reactions, but the reaction conditions are rather harsh [166]. Therefore, a Cu-based catalytic system with high efficiency and regio-selectivity with broad substrate scope working under mild conditions is highly desirable.
Recently, the Zang group reported a tridentate N-heterocyclic carbene ligand stabilized [Cu3(NHC)3(PF6)3] (Cu3NC) cluster for both A3 coupling reactions and redox-A3 coupling reactions [60]. The crystal structure of the NHC ligand and Cu3NC are shown in Fig. 14a, b, respectively, where the NHC ligand endows the Cu3NC with dual attributes of flexibility and rigidity. On one hand, the stable Cu–C and Cu–N bonds between the NHC ligand shell and the metal core favors the stability of the Cu3NC. On the other hand, the pyridine of the N‑heterocyclic carbene has somewhat dynamic balance between the aliphatic amines and the pyridine to protect the catalytic centers and prevent the Cu3NC deactivation. Such dynamic balance endows Cu3NC with flexible features. Consequently, high activity and high regioselectivity can be achieved in all-in-one A3 coupling reactions with inert substrates at room temperature (Fig. 14c), as manifested by 71 examples of A3 coupling reactions and 14 examples of redox A3 coupling reaction, both up to 99% yield [60]. The following mechanistic and control experimental tests demonstrated that the remarkable catalytic performance originates from the flexible and rigid dual attributes of the Cu3NC.
It is worth noting that most ligand protected Cu nanoclusters feature a core–shell framework hence the active sites can be blocked by metal–ligand shell or surface passivation units. Therefore, a strategy that can largely expose the metal active sites is long pursued for catalytic coupling reaction. In 2023, Xu et al. reported a novel approach to construct array-based ([Cu8(Tf-dpf)4(NO3)2](NO3)2 clusters [61]. Compared with the ([Cu8(RS)6(PPh3)4(MeCN)4H]+ with the cubic core, the above array based Cu8 cluster displays largely uncoordinated metal sites, hence exhibited significantly enhanced catalytic activity in the “aldehyde-acetylene-amine” A3-coupling reaction for synthesizing propargylamines. It shows the great potential to fabricate array-based Cu nanoclusters for A3 coupling reaction and beyond [61].
Synthesis for Indolizines
To realize the practical catalytic application of Cu nanoclusters, the long-term stability remains a long-standing challenge. In the regime of coinage metal nanocluster, compared with Au, Ag, and Au/Ag alloy clusters, the study of Cu nanoclusters lags far behind, mainly due to the high reactivity of Cu and low standard reduction potential (E0Cu2+/Cu+ = 0.34 V, E0Cu+/Cu = 0.52 V) [22]. The surface organic ligand such as thiolate molecules, alkynes, halides, hydride can improve the stability due to the formation of strong chemical bonds, but the long-term stability for some specific organic reactions are still unsatisfactory. These ligands are considered as soft base, and Cu2+ or Cu+ ions belong to soft acid; therefore, the ligand–metal interaction fits the Pearson’s hard/soft acid/base theory [167].
In a recent study conducted by the Li and Ni team, pyrazolate (denote as Pz) ligands were employed to synthesize Cu8 nanoclusters [62]. Pz is a soft base, and thanks to the high pKa value of 19.8 for deprotonated pyrazole [168], Pz protected Cu clusters are expected to exhibit exceptional alkali-resistant capability. As determined from SCXRD (Fig. 15a), the flexible bis-pyrazole ligands (H2L1) can transform into two novel deprotonated bis-pyrazole ligands (L2 and L3), and all these L2 and L3 ligands can assemble with Cu+ to yield Cu8 nanoclusters. The crystal of Cu8 clusters can keep intact in various organic solvents, highly concentrated acid, saturated alkali, oxidant, and boiling water, and still are fine for SCXRD test. More impressively, the introduction of organic acid or base (100 eq. HOAc, or 400 eq. dibutylamine) cannot impose structural decomposition, neither, confirming the remarkable stability never documented for Cu nanoclusters (Fig. 15b). Subsequently, the synthesis for indolizines from 2-pyridinecarboxaldehyde derivatives, terminal alkynes and secondary amines was tested using Cu8 nanocluster as a catalyst. Interestingly, the reaction cannot proceed without Cu8 cluster as the catalyst (Fig. 15c). The optimized condition is 0.005% mol of substrates. Other Cu nanoclusters such as Cu10 or Cu18 had a much lower yield, as they are not stable in basic conditions (Fig. 15d). Consequently, the turnover frequency (TOF) of Cu8 nanoclusters can reach 3880 h−1, about 3 orders of magnitude larger than the reported catalysts toward the synthesis of indolizines [62]. This study provides a new strategy to synthesize pyrazolate-protected Cu nanoclusters with ultrahigh chemical stability for practical applications.
Reduction of Ferricyanide to Ferrocyanide
Ferricyanide ([Fe(CN)6]3−) is an environmentally unfriendly material that readily reacts with acids to produce toxic hydrogen cyanide gas. The catalytic reduction of [Fe(CN)6]3− to ferrocyanide ([Fe(CN)6]4−) is one effective strategy to limit the toxicity of [Fe(CN)6]3− [169].
In a recent study, the Luo and Sun team reported the Cu18H(PET)14(PPh3)6(isothiocyanate)3 (Cu18H in short) nanocluster [63], which exhibits excellent catalytic performance in the reduction of ferricyanide. Cu18H comprises a pseudo D3-symmetrical triple-stranded helical Cu15 kernel, and can be structurally described as layer-by-layer combination of multiple chiral Cu nanoclusters linked through copper-thiolate bonds. Remarkably, the aggregate state of the Cu18H nanoclusters can catalyze electron transfer reactions efficiently. To assess its catalytic capability, ultraviolet/visible absorption spectroscopy was used to monitor the reduction process. When Cu18H and NaBH4 were added simultaneously, [Fe(CN)6]3− can be completely converted into [Fe(CN)6]4− very rapidly, and the solution turned into colorless within 14 s. In a control test, the reduction reaction of [Fe(CN)6]3− without Cu18H required 16 min. It indicates that Cu18H has good catalytic reactivity for the reduction of [Fe(CN)6]3− [63]. This study makes an interesting model for investigating and elucidating the aggregated state of copper nanoclusters.
Hydrogenation of Cyclohexanone
The chemical hydrogenation of cyclohexanone is among the most important and prevalent transformations in industrial organic synthesis, thanks to its effectiveness and economic viability [170, 171]. A previous report suggested that copper hydride nanoclusters are very promising for catalyzing the hydrogenation of carbonyl compounds, despite that there are challenges to control the product selectivity [172].
In a recent work done by Sun et al., a cluster of [Cu66Cl8(PPh3)8(SC2H5)32H24](SbF6)2 (Cu66 in short) was synthesized and its catalytic performance toward the hydrogenation of cyclohexanone was examined [64]. Cu66 contains an orderly of 16 Cu4 squares, and the ligands coordinate to the surface of the cluster in a regiospecific manner, displaying square pattern as well. To enhance the monodispersity and stability of the cluster catalyst, the Cu66 nanoclusters were immobilized on carbon black (XC-72) to afford the Cu66/XC-72 catalyst, and it demonstrated an unprecedented performance toward the hydrogenation of cyclohexanone. The selectivity of cyclohexanol is nearly 100%. Under specific conditions, complete conversion of cyclohexanone to the desired product was achieved within 40 h. The authors also tested all the other Cu catalysts, including (Cu(CF3CO2)2, CuCl, Cu54Cl12(NO3)12(SC4H9)20S (Cu54), Cu50(PhCOO)10(4-F-PhS)20(PPh3)H2 (Cu50), SeCu20(PhSe)12(PPh3)2(C6H5COO)6 (Cu20), and [Cu25(SPhCl2)18H10]3− (Cu25), but they showed significantly lower yields for the corresponding product. Moreover, the Cu66/XC-72 catalyst exhibited no significant decay in cyclohexanol product selectivity, highlighting the robustness of the Cu66/XC-72 catalyst [64]. This study showcases Cu nanoclusters have high stability and exceptional catalytic activity in the hydrogenation of carbonyl compounds, envisioning a bright future in both fundamental research and practical applications.
Cu NCs for Other Reactions
Atomically precise Cu nanoclusters can also find versatile catalytic applications toward other chemical reactions. One typical reaction is the catalytic hydrogenation of nitroarene. It is worth noting that the catalytic hydrogenation of toxic nitroaromatics into non-toxic amino molecules is a quite efficient and economical method, as amino molecules are valuable intermediates for pharmaceuticals and dyes [169, 173, 174]. Coinage metal nanoclusters have been emerging as a new type of important hydrogenation catalysts, and the catalytic hydrogenation of 4-nitrophenol into 4-aminophenol in the presence of BH4− reducing agent has been employed as a model reaction in several previous studies [32, 92, 95, 175].
In 2023, Luo et al. reported an intrinsically chiral Cu hydride-rich nanocluster of [Cu57H20(PET)36(TPP)4]+ (Cu57H20) and its catalytic activity toward reduction of 4-nitrophenol [56]. Single crystal X-ray diffraction measurement shows that the unit cell of the Cu57H20 cluster contains a pair of enantiomers, C-Cu57H20 and A-Cu57H20 (C, clockwise; A, anti-clockwise) (Fig. 16a). The hydride positions were further validated by DFT optimization of the model cluster of (Cu57H20)opt. Interestingly, the DFT analysis also suggests that the interfacial μ3-H in (Cu57H20)opt can be transformed into an interstitial μ5-H species in the perfect cluster of (Cu57H20)opt as long as the vacant position is recovered by the 58th Cu atom (Fig. 16b). The Cu57H20 cluster was then employed as the catalyst to catalyze the reduction of 4-nitrophenol. The reaction process was monitored by UV–vis absorption spectra (Fig. 16c), where strong absorbance peak at 400 nm from 4-nitrophenol decreased rapidly in 20 min and the absorbance peak at 300 nm from 4-aminophenol increased. The bright yellow solution also turned colorless. Using the absorbance change at 400 nm as metric, the reduction adopts a pseudo-first-order reaction kinetics, and the reaction rate was calculated as 18 min−1 (Fig. 16d). A control experiment in the absence of Cu57H20 was also conducted, and no reduction occurs on 4-nitrophenol when adding BH4−, probably due to the high activation barrier between the repelling 4-nitrophenol anion and BH4− [176]. Finally, the authors performed a 4-nitrophenol reduction test by using BD4− to check the accessibility of the hydride in Cu57H20. After the catalytic test, the cluster was examined by the ESI–MS measurement (Fig. 16e). It shows well defined peak feature of [Cu57H19D(PET)36(TPP)4 + Cu]2+, indicating one H− is indeed replaced by one D− of BD4−, that means, the exposed interfacial μ3-H might be involved in the catalytic reaction [56]. By using defective Cu57H20 cluster as model catalyst, this work offers atom-precise insights into the vacant defect role in catalysis.
In addition, a very small change in composition can make a big difference in structure and catalytic performance for some specific catalytic reactions. In a recent report, the Wang and Jiang team disclosed the dramatic difference between Cu20H8 and Cu20H9 clusters in catalysis [65]. Specifically, two copper hydride clusters, [Cu20H9(Tf-dpf)10]·BF4 (Cu20H9) and [Cu20H8(Tf-dpf)10]·(BF4)2 (Cu20H8) (Tf-dpf = N, N′-di(5-trifluoromethyl-2-pyridyl)formamidinate) were fabricated, and both have twenty Cu atoms and ten amidinate ligands but with one hydride difference in the ligand shell. Such differences led to drastically different geometric and electronic structures, resulting in different catalytic properties. Cu20H8 showed 25 times higher catalytic activity than Cu20H9 in the conjugate reduction of cinnamaldehyde, mainly due to the easier dissociation process of a Tf-dpf− ligand in Cu20H8 [65]. This work highlights the sensitivity of structure and composition toward catalytic reactions by using atomically precise copper hydride nanoclusters as catalysts.
Recently, the Shen group found that the open metal site plays a critical role in the catalytic reduction of 4-nitrophenol by using two Cu41 nanoclusters as model catalysts [46]. Specifically, two isostructural Cu nanoclusters of [Cu41Cl2(2-FC6H4S)12(CF3COO)6(PPh3)6H19]2− (1H) and [Cu41(2,5-di-Methyl-C6H3S)12(BO3)3Cl3(PPh3)6H19] (2H) were prepared and fully characterized, and the main structural difference of them lies in the absence or presence of two additional chlorides on the surface, which in turn heavily governs the exposure of metal sites. The 2H cluster with more open active sites exhibited ~ sixfold increase of rate constant in the catalytic reduction of 4-nitrophenol than 1H cluster [46]. This work manifests that atomically precise Cu nanoclusters can serve as catalyst models to directly visualize the active sites that drive the chemical transformation. Meanwhile, the same group also discovered that the small distinctions in two carboxylate-protected Cu20 clusters can cause distinct catalytic performance toward 4-nitrophenol reduction [66]. Two clusters Se@Cu20(PhSe)12(PPh3)2(C6H5COO)6 (Cu20-1) and Se@Cu20(PhSe)12(PPh3)2(CF3COO)6 (Cu20-2) share with the identical metal skeleton and similar ligand distributions, and the main difference is the carboxylate ligand: C6H5COO− for Cu20-1 while CF3COO− for Cu20-2. Consequently, Cu20-1 has a catalytic activity enhancement of 16-fold than Cu20-2. Such catalytic performance distinction is attributed to the carboxylate effect, that is, the functional group (-C6H5CO2 or -CF3CO2) altered the electronic structure of the Cu nanocluster [66].
Challenges and Perspectives
The last decade, especially the recent five years, have witnessed great success in atomically precise Cu nanoclusters, including synthesis, structural analysis, property exploration, and various catalytic applications. Despite the significant progress, there are still some important obstacles for nourishing the atomically precise Cu nanocluster field, especially in both synthetic regime and catalytic field. Some possible challenges and future perspectives are discussed below, aiming to motivate or advance this fast-growing field:
High-efficiency synthetic method to obtain atomically precise Cu nanoclusters is still highly desired. Compared with the fruitful achievements in Au and Ag nanoclusters, the synthesis and total structure determination of Cu nanoclusters have been lagging behind, probably due to the variability of Cu valence states. It is known that the subtle reaction conditions may affect the output of the final product, e.g., the solvent, the reaction temperature, the ligand (discuss below), the reducing agent, the pH value in the reaction system, the stoichiometric ratio of all the reactants, etc. For example, by using CH2Cl2 as the solvent, the Cl atoms can coordinate with Ag and become the capping ligand in alkynyl-protected Ag112 nanoclusters [177], and it might also be applicable for preparing Cu nanoclusters. More importantly, the current synthetic methods of atomically precise Cu nanoclusters are heavily dependent on the “the-trial-and-error” efforts, which are quite tedious, expensive, and low efficient. High throughput synthesis with the aid of artificial intelligence or machine learning technique represents an embryonic avenue yet still have a long way to go [178, 179].
The role of ligand should be carefully considered. The choosing of ligand molecule not only affects the synthesis, but also plays a critical role in the catalytic process. Thiolate, alkynyl molecule, hydride, halogen, phosphine, N-heterocyclic carbene, inorganic anions (e.g., BF4-, CH3COO-) are most widely employed ligand molecules for stabilizing Cu nanoclusters, but they possess different binding strength to the Cu core. It is commonly believed, the surface ligand can block some active sites for catalysis [161], as they can inhibit the reactants for accessing the metal sites. However, more and more cases have shown that, the ligand can facilitate the catalytic process, e.g., accelerate the transfer of the reactants or intermediates via weak interactions. Weak surface ligand is more easily stripped off or exchanged by the reactant to create catalytic sites [10]. Moreover, the type of ligand molecule affects the catalytic activity drastically, e.g., Tsukuda group reported that alkynyl-protected Au25 clusters exhibited markedly higher HER activity than thiolate-protected Au25 molecules [180], while our group discovered that, the Faradaic efficiency of CO for alkynyl-protected Ag32 clusters is much higher than the thiolate and phosphine ligand co-protected Ag32 clusters in eCO2RR [32]. Such catalytic property difference is attributed to the electronic structure change owing to the electronic perturbation of the π-conjugated units [72]. Even with the same type of molecule in ligand, the steric hindrance and electronic structure can affect the catalytic properties, e.g., using bulky ligand may create some low coordinated or undercoordinated metal sites for catalysis [159, 181]. The Xie group found that, in the oxygen evolution reaction catalyzed by three thiolates protected Au25 nanoclusters, p-mercaptobenzoic acid stabilized Au25 nanoclusters exhibited markedly superior catalytic performance, simply because the ligand’s stronger electron-withdrawing ability can create more partial positive charges on Au(I) as active site for facilitating feasible adsorption of OH- in alkaline media [182]. Yoo et al. found that the locally induced hydrophobicity by bulky alkyl functionality near the surface of the Ag25(SR)18 cluster dramatically enhanced the eCO2RR activity, where the hydrophobic Ag25 cluster exhibited remarkable selectivity for CO (FECO > 90%) and achieved a high current density of up to -240 mA cm-2 with excellent durability lasting for over 120 h [183]. Nevertheless, the ligand role of atomically precise Cu nanoclusters during the synthetic process and catalytic duration remains largely to be explored.
The metal core tailoring. The metal core configuration and atomic spatial arrangement can be critical for exposing available catalytic active sites for Cu nanoclusters. So far, there are core-shell, rod-like, array-like core structure and other core configurations for Cu nanoclusters, but a trade-off between activity, selectivity and stability must be taken into account for specific catalytic reactions [69]. More profoundly, more dedicated core structure engineering might be necessary for Cu nanoclusters. Vacancy engineering has demonstrated great potential, which is expected to catch more future research attention [58]. Another important direction is alloying other metals to form atomically precise Cu–alloy nanoclusters. For catalytic reactions, the size, composition, configuration, electronic structure are the main factors that influence the catalytic performance, and introducing another metal to form Cu–alloy can modulate all the above factors. Cu-based bimetallic nanoclusters have demonstrated unusual properties and extraordinary catalytic capabilities [90, 184], yet controlling the number and the exact position of a foreign metal atom remains challenging [163, 185]. Particularly, single-metal-atom doping have been gaining tremendous efforts in catalysis study but increasing the production yield for generating single-atom doping during the synthetic process, understanding the geometric and electronic structure change in the on-working catalytic process is essential for future high-performance catalyst design [186–189].
In situ reaction mechanism study. Understanding the reaction mechanism in on-working status is quite valuable but still challenging. The in situ and operando spectroscopic and synchronous techniques such as surface enhanced Raman scattering, infrared spectroscopy, high-energy X-ray diffraction (HE-XRD), extended X-ray absorption fine structure (EXAFS), X-ray absorption near edge structure (XANES), and small-angle neutron scattering (SANS) can provide the possibility of acquiring the structure of the catalytic intermediates, which can enhance the evidence for proposed reaction mechanism [190–193]. For instance, the in situ infrared spectroscopy was employed to capture the reaction intermediates during the electrocatalytic nitrate reduction catalyzed by Ag30Pd4 nanoclusters; hence, the tandem catalytic mechanism was successfully revealed by our group [93]. Nevertheless, the consecutive picture of the whole reaction process including adsorption, transfer, desorption of the reaction intermediates on Cu nanoclusters are quite limited in current studies, and more advanced in situ/operando study and the necessary instrumental modification for some specific in situ/operando investigations are expected.
The more precise theoretical simulations. Theoretical simulations can calculate the energy barrier of each step, acquire the optimal configuration between the cluster and the reactant, the intermediate, or the product, hence plays a critical role in speculating the active sites and deciding the rate-determining step [104]. However, the accuracy of these calculations depends on the authenticity and the extent of the structure proximity between the real crystal structure and the theoretical model. The fully ligand capped Cu nanoclusters are too complicated to build a real model, so the “ligand simplification” strategy is widely used. For instance, Au25(SCH3)18 has been used to replace Au25(SC2H4Ph)18, Au25(SC6H13)18 and Au25(SC12H25)18 in several cases for building theoretical models [194–196]. Authentic models can help to understand the true reaction process and gain more realistic reaction mechanism but given the complexity of the cluster structures and the reaction conditions, more precise and realistic theoretical simulations still have a long way to go.
Acknowledgements
The study is supported by the open funds of Key Laboratory of Functional Inorganic Material Chemistry (Heilongjiang University), Ministry of Education, China. Z. T. acknowledges the funding from Guangdong Natural Science Funds (No. 2023A0505050107).
Author Contributions
MC contributed to the investigation and original draft writing. CG, LQ, LW, LQ, and KC were involved in the literature search and formal analysis. ZT performed the funding acquisition and draft finalization.
Declarations
Conflict of interest
The authors declare no conflict of interest. They have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Footnotes
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.R. Jin, C. Zeng, M. Zhou, Y. Chen, Atomically precise colloidal metal nanoclusters and nanoparticles: fundamentals and opportunities. Chem. Rev. 116, 10346–10413 (2016). 10.1021/acs.chemrev.5b00703 [DOI] [PubMed] [Google Scholar]
- 2.Y. Lv, T. Jiang, Q. Zhang, H. Yu, M. Zhu, Recent progress in atomically precise Ag/Cu-based hydride clusters. Polyoxometalates 3, 9140050 (2024). 10.26599/pom.2023.9140050
- 3.X.-M. Luo, Y.-K. Li, X.-Y. Dong, S.-Q. Zang, Platonic and Archimedean solids in discrete metal-containing clusters. Chem. Soc. Rev. 52, 383–444 (2023). 10.1039/d2cs00582d [DOI] [PubMed] [Google Scholar]
- 4.R.W. Murray, Nanoelectrochemistry: metal nanoparticles, nanoelectrodes, and nanopores. Chem. Rev. 108, 2688–2720 (2008). 10.1021/cr068077e [DOI] [PubMed] [Google Scholar]
- 5.P.D. Jadzinsky, G. Calero, C.J. Ackerson, D.A. Bushnell, R.D. Kornberg, Structure of a thiol monolayer-protected gold nanoparticle at 1.1 A resolution. Science 318, 430–433 (2007). 10.1126/science.1148624 [DOI] [PubMed]
- 6.L. Shang, J. Xu, G. Nienhaus, Recent advances in synthesizing metal nanocluster-based nanocomposites for application in sensing, imaging and catalysis. Nano Today 28, 100767 (2019). 10.1016/j.nantod.2019.100767 [Google Scholar]
- 7.S. Qian, Z. Wang, Z. Zuo, X. Wang, Q. Wang et al., Engineering luminescent metal nanoclusters for sensing applications. Coord. Chem. Rev. 451, 214268 (2022). 10.1016/j.ccr.2021.214268 [Google Scholar]
- 8.Y. Jin, C. Zhang, X.-Y. Dong, S.-Q. Zang, T.C.W. Mak, Shell engineering to achieve modification and assembly of atomically-precise silver clusters. Chem. Soc. Rev. 50, 2297–2319 (2021). 10.1039/d0cs01393e [DOI] [PubMed] [Google Scholar]
- 9.Q. Yao, L. Liu, S. Malola, M. Ge, H. Xu et al., Supercrystal engineering of atomically precise gold nanoparticles promoted by surface dynamics. Nat. Chem. 15, 230–239 (2023). 10.1038/s41557-022-01079-9 [DOI] [PubMed] [Google Scholar]
- 10.W. Jing, H. Shen, R. Qin, Q. Wu, K. Liu et al., Surface and interface coordination chemistry learned from model heterogeneous metal nanocatalysts: from atomically dispersed catalysts to atomically precise clusters. Chem. Rev. 123, 5948–6002 (2023). 10.1021/acs.chemrev.2c00569 [DOI] [PubMed] [Google Scholar]
- 11.M.S. Bootharaju, C.W. Lee, G. Deng, H. Kim, K. Lee et al., Atom-precise heteroatom core-tailoring of nanoclusters for enhanced solar hydrogen generation. Adv. Mater. 35, e2207765 (2023). 10.1002/adma.202207765 [DOI] [PubMed] [Google Scholar]
- 12.H. Shan, J. Shi, T. Chen, Y. Cao, Q. Yao et al., Modulating catalytic activity and stability of atomically precise gold nanoclusters as peroxidase mimics via ligand engineering. ACS Nano 17, 2368–2377 (2023). 10.1021/acsnano.2c09238 [DOI] [PubMed] [Google Scholar]
- 13.Y.-S.Chen, H. Choi, P.V. Kamat, Metal-cluster-sensitized solar cells. A new class of thiolated gold sensitizers delivering efficiency greater than 2%. J. Am. Chem. Soc. 135, 8822–8825 (2013). 10.1021/ja403807f [DOI] [PubMed]
- 14.Y.-S. Chen, P.V. Kamat, Glutathione-capped gold nanoclusters as photosensitizers. Visible light-induced hydrogen generation in neutral water. J. Am. Chem. Soc. 136, 6075–6082 (2014). 10.1021/ja5017365 [DOI] [PubMed]
- 15.B. Du, X. Jiang, A. Das, Q. Zhou, M. Yu et al., Glomerular barrier behaves as an atomically precise bandpass filter in a sub-nanometre regime. Nat. Nanotechnol. 12, 1096–1102 (2017). 10.1038/nnano.2017.170 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Y. Hua, J.-H. Huang, Z.-H. Shao, X.-M. Luo, Z.-Y. Wang et al., Composition-dependent enzyme mimicking activity and radiosensitizing effect of bimetallic clusters to modulate tumor hypoxia for enhanced cancer therapy. Adv. Mater. 34, 2203734 (2022). 10.1002/adma.202203734 [DOI] [PubMed] [Google Scholar]
- 17.T.-T. Jia, G. Yang, S.-J. Mo, Z.-Y. Wang, B.-J. Li et al., Atomically precise gold–levonorgestrel nanocluster as a radiosensitizer for enhanced cancer therapy. ACS Nano 13, 8320–8328 (2019). 10.1021/acsnano.9b03767 [DOI] [PubMed] [Google Scholar]
- 18.M. Walter, J. Akola, O. Lopez-Acevedo, P.D. Jadzinsky, G. Calero et al., A unified view of ligand-protected gold clusters as superatom complexes. Proc. Natl. Acad. Sci. U.S.A. 105, 9157–9162 (2008). 10.1073/pnas.0801001105 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.M.G. Taylor, G. Mpourmpakis, Thermodynamic stability of ligand-protected metal nanoclusters. Nat. Commun. 8, 15988 (2017). 10.1038/ncomms15988 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.S.-Y. Kang, Z.-A. Nan, Q.-M. Wang, Superatomic orbital splitting in coinage metal nanoclusters. J. Phys. Chem. Lett. 13, 291–295 (2022). 10.1021/acs.jpclett.1c03563 [DOI] [PubMed] [Google Scholar]
- 21.J.D. Schneider, B.A. Smith, G.A. Williams, D.R. Powell, F. Perez et al., Synthesis and characterization of Cu(II) and mixed-valence Cu(I)Cu(II) clusters supported by pyridylamide ligands. Inorg. Chem. 59, 5433–5446 (2020). 10.1021/acs.inorgchem.0c00008 [DOI] [PubMed] [Google Scholar]
- 22.L.L.-M. Zhang, W.-Y. Wong, Atomically precise copper nanoclusters as ultrasmall molecular aggregates: appealing compositions, structures, properties, and applications. Aggregate 4, e266 (2023). 10.1002/agt2.266 [Google Scholar]
- 23.J. Sang, P. Wei, T. Liu, H. Lv, X. Ni et al., A reconstructed Cu2P2O7 catalyst for selective CO2 electroreduction to multicarbon products. Angew. Chem. Int. Ed. 61, e202114238 (2022). 10.1002/anie.202114238 [DOI] [PubMed] [Google Scholar]
- 24.C. Sun, K. Liu, J. Zhang, Q. Liu, X.J. Liu et al., In situ transmission electron microscopy and three-dimensional electron tomography for catalyst studies. Chin. J. Struct. Chem. 41, 2210056–2210076 (2022). 10.14102/j.cnki.0254-5861.2022-0187
- 25.H. Sun, L. Lin, W. Hua, X. Xie, Q. Mu et al., Atomically dispersed Co−Cu alloy reconstructed from metal-organic framework to promote electrochemical CO2 methanation. Nano Res. 16, 3680–3686 (2023). 10.1007/s12274-022-4728-1 [Google Scholar]
- 26.X. Gong, Z.-H. Liu, Q. Xu, L. Wang, Q. Guo et al., Single-molecule manipulation of copper nanoclusters for modulating nonlinear optics. Polyoxometalates 4, 9140072 (2025). 10.26599/pom.2024.9140072
- 27.Q.-J. Wu, D.-H. Si, P.-P. Sun, Y.-L. Dong, S. Zheng et al., Atomically precise copper nanoclusters for highly efficient electroreduction of CO2 towards hydrocarbons via breaking the coordination symmetry of Cu site. Angew. Chem. Int. Ed. 62, e202306822 (2023). 10.1002/anie.202306822 [DOI] [PubMed] [Google Scholar]
- 28.S. Zhao, N. Austin, M. Li, Y. Song, S.D. House et al., Influence of atomic-level morphology on catalysis: the case of sphere and rod-like gold nanoclusters for CO2 electroreduction. ACS Catal. 8, 4996–5001 (2018). 10.1021/acscatal.8b00365 [Google Scholar]
- 29.L. Qin, F. Sun, X. Ma, G. Ma, Y. Tang et al., Homoleptic alkynyl-protected Ag15 nanocluster with atomic precision: structural analysis and electrocatalytic performance toward CO2 reduction. Angew. Chem. Int. Ed. 60, 26136–26141 (2021). 10.1002/anie.202110330 [DOI] [PubMed] [Google Scholar]
- 30.X.-K. Wan, J.-Q. Wang, Q.-M. Wang, Ligand-protected au55 with a novel structure and remarkable CO2 electroreduction performance. Angew. Chem. Int. Ed. 60, 20748–20753 (2021). 10.1002/anie.202108207 [DOI] [PubMed] [Google Scholar]
- 31.S.-F. Yuan, R.-L. He, X.-S. Han, J.-Q. Wang, Z.-J. Guan et al., Robust gold nanocluster protected with amidinates for electrocatalytic CO2 reduction. Angew. Chem. Int. Ed. 60, 14345–14349 (2021). 10.1002/anie.202103060 [DOI] [PubMed] [Google Scholar]
- 32.L. Chen, F. Sun, Q. Shen, L. Qin, Y. Liu et al., Homoleptic alkynyl-protected Ag32 nanocluster with atomic precision: probing the ligand effect toward CO2 electroreduction and 4-nitrophenol reduction. Nano Res. 15, 8908–8913 (2022). 10.1007/s12274-022-4812-6 [Google Scholar]
- 33.J. Wang, F. Xu, Z.-Y. Wang, S.-Q. Zang, T.C.W. Mak, Ligand-shell engineering of a Au28 nanocluster boosts electrocatalytic CO2 reduction. Angew. Chem. Int. Ed. 61, e202207492 (2022). 10.1002/anie.202207492 [DOI] [PubMed] [Google Scholar]
- 34.S. Li, A.V. Nagarajan, X. Du, Y. Li, Z. Liu et al., Dissecting critical factors for electrochemical CO2 reduction on atomically precise Au nanoclusters. Angew. Chem. Int. Ed. 61, e202211771 (2022). 10.1002/anie.202211771 [DOI] [PubMed] [Google Scholar]
- 35.T.-A D. Nguyen, Z.R. Jones, D.F. Leto, G. Wu, S.L. Scott et al., Ligand-exchange-induced growth of an atomically precise Cu29 nanocluster from a smaller cluster. Chem. Mater. 28, 8385–8390 (2016). 10.1021/acs.chemmater.6b03879
- 36.A.W. Cook, Z.R. Jones, G. Wu, S.L. Scott, T.W. Hayton, An organometallic Cu20 nanocluster: synthesis, characterization, immobilization on silica, and click chemistry. J. Am. Chem. Soc. 140, 394–400 (2018). 10.1021/jacs.7b10960 [DOI] [PubMed] [Google Scholar]
- 37.M. Qu, F.-Q. Zhang, D.-H. Wang, H. Li, J.-J. Hou et al., Observation of non-FCC copper in alkynyl-protected Cu53 nanoclusters. Angew. Chem. Int. Ed. 59, 6507–6512 (2020). 10.1002/anie.202001185 [DOI] [PubMed] [Google Scholar]
- 38.S. Li, X. Yan, J. Tang, D. Cao, X. Sun et al., Cu26 nanoclusters with quintuple ligand shells for CO2 electrocatalytic reduction. Chem. Mater. 35, 6123–6132 (2023). 10.1021/acs.chemmater.3c01247 [Google Scholar]
- 39.S. Lee, M.S. Bootharaju, G. Deng, S. Malola, W. Baek et al., Cu32(PET)24H8Cl2](PPh4)2: a copper hydride nanocluster with a bisquare antiprismatic core. J. Am. Chem. Soc. 142, 13974–13981 (2020). 10.1021/jacs.0c06577 [DOI] [PubMed] [Google Scholar]
- 40.T.-A D. Nguyen, Z.R. Jones, B.R. Goldsmith, W.R. Buratto, G. Wu et al., A Cu25 nanocluster with partial Cu(0) character. J. Am. Chem. Soc. 137, 13319–13324 (2015). 10.1021/jacs.5b07574 [DOI] [PubMed]
- 41.N. Zhang, Y. Li, S. Han, Y. Wei, H. Hu et al., Cluster light-emitting diodes containing copper iodine cube with 100% exciton utilization using host-cluster synergy. Angew. Chem. Int. Ed. 62, e202305018 (2023). 10.1002/anie.202305018 [DOI] [PubMed] [Google Scholar]
- 42.A.K. Das, S. Biswas, V.S. Wani, A.S. Nair, B. Pathak et al., Cu18H3(S-adm)12(PPh3)4Cl2]: fusion of Platonic and Johnson solids through a Cu(0) center and its photophysical properties. Chem. Sci. 13, 7616–7625 (2022). 10.1039/d2sc02544b [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.T. Jia, Z.-J. Guan, C. Zhang, X.-Z. Zhu, Y.-X. Chen et al., Eight-electron superatomic Cu31 nanocluster with chiral kernel and NIR-II emission. J. Am. Chem. Soc. 145, 10355–10363 (2023). 10.1021/jacs.3c02215 [DOI] [PubMed] [Google Scholar]
- 44.X.-Z. Zhu, T. Jia, Z.-J. Guan, Q. Zhang, Y. Yang, Elongation of a trigonal-prismatic copper cluster by diphosphine ligands with longer spacers. Inorg. Chem. 61, 15144–15151 (2022). 10.1021/acs.inorgchem.2c02306 [DOI] [PubMed] [Google Scholar]
- 45.J. Sun, X. Tang, Z.-H. Liu, Z. Xie, B. Yan et al., Labile ligands protected Cu50 nanoclusters with tailorable optical limiting effect. ACS Mater. Lett. 6, 281–289 (2024). 10.1021/acsmaterialslett.3c01305 [Google Scholar]
- 46.J. Sun, X. Tang, X. Yan, W. Jing, Z. Xie et al., Atomically precise Cu41 clusters as model catalysts: open metal sites matter. Next Mater. 3, 100091 (2024). 10.1016/j.nxmate.2023.100091 [Google Scholar]
- 47.B.-L. Han, Z. Liu, L. Feng, Z. Wang, R.K. Gupta et al., Polymorphism in atomically precise Cu23 nanocluster incorporating tetrahedral[Cu4]0 kernel. J. Am. Chem. Soc. 142, 5834–5841 (2020). 10.1021/jacs.0c01053 [DOI] [PubMed] [Google Scholar]
- 48.P. Yuan, R. Chen, X. Zhang, F. Chen, J. Yan et al., Ether-soluble Cu53 nanoclusters as an effective precursor of high-quality CuI films for optoelectronic applications. Angew. Chem. Int. Ed. 58, 835–839 (2019). 10.1002/anie.201812236 [DOI] [PubMed] [Google Scholar]
- 49.C. Zhang, Z. Wang, W.-D. Si, L. Wang, J.-M. Dou et al., Solvent-induced isomeric Cu13 nanoclusters: chlorine to copper charge transfer boosting molecular oxygen activation in sulfide selective oxidation. ACS Nano 16, 9598–9607 (2022). 10.1021/acsnano.2c02885 [DOI] [PubMed] [Google Scholar]
- 50.L. Qin, G. Ma, L. Wang, Z. Tang, Atomically precise metal nanoclusters for (photo)electroreduction of CO2: recent advances, challenges and opportunities. J. Energy Chem. 57, 359–370 (2021). 10.1016/j.jechem.2020.09.003 [Google Scholar]
- 51.Q. Tang, Y. Lee, D.-Y. Li, W. Choi, C.W. Liu et al., Lattice-hydride mechanism in electrocatalytic CO2 reduction by structurally precise copper-hydride nanoclusters. J. Am. Chem. Soc. 139, 9728–9736 (2017). 10.1021/jacs.7b05591 [DOI] [PubMed] [Google Scholar]
- 52.L.-J. Liu, Z.-Y. Wang, Z.-Y. Wang, R. Wang, S.-Q. Zang et al., Mediating CO2 electroreduction activity and selectivity over atomically precise copper clusters. Angew. Chem. Int. Ed. 61, e202205626 (2022). 10.1002/anie.202205626 [DOI] [PubMed] [Google Scholar]
- 53.J. Wang, J. Cai, K.-X. Ren, L. Liu, S.-J. Zheng et al., Stepwise structural evolution toward robust carboranealkynyl-protected copper nanocluster catalysts for nitrate electroreduction. Sci. Adv. 10, eadn7556 (2024). 10.1126/sciadv.adn7556 [DOI] [PMC free article] [PubMed]
- 54.J.-P. Dong, Y. Xu, X.-G. Zhang, H. Zhang, L. Yao et al., Copper-sulfur-nitrogen cluster providing a local proton for efficient carbon dioxide photoreduction. Angew. Chem. Int. Ed. 62, e202313648 (2023). 10.1002/anie.202313648 [DOI] [PubMed] [Google Scholar]
- 55.A. Sagadevan, A. Ghosh, P. Maity, O.F. Mohammed, O.M. Bakr et al., Visible-light copper nanocluster catalysis for the C-N coupling of aryl chlorides at room temperature. J. Am. Chem. Soc. 144, 12052–12061 (2022). 10.1021/jacs.2c02218 [DOI] [PubMed] [Google Scholar]
- 56.G.-G. Luo, Z.-H. Pan, B.-L. Han, G.-L. Dong, C.-L. Deng et al., Total structure, electronic structure and catalytic hydrogenation activity of metal-deficient chiral polyhydride Cu57 nanoclusters. Angew. Chem. Int. Ed. 62, e202306849 (2023). 10.1002/anie.202306849 [DOI] [PubMed] [Google Scholar]
- 57.C. Dong, R.-W. Huang, A. Sagadevan, P. Yuan, L. Gutiérrez-Arzaluz et al., Isostructural nanocluster manipulation reveals pivotal role of one surface atom in click chemistry. Angew. Chem. Int. Ed. 62, e202307140 (2023). 10.1002/anie.202307140 [DOI] [PubMed] [Google Scholar]
- 58.S. Nematulloev, A. Sagadevan, B. Alamer, A. Shkurenko, R. Huang et al., Atomically precise defective copper nanocluster catalysts for highly selective C-C cross-coupling reactions. Angew. Chem. Int. Ed. 62, e202303572 (2023). 10.1002/anie.202303572 [DOI] [PubMed] [Google Scholar]
- 59.S. Biswas, A. Pal, M.K. Jena, S. Hossain, S. Jin et al., Luminescent hydride-free[Cu7(SC5H9)7(PPh3)3]nanocluster: facilitating highly selective C-C bond formation. J. Am. Chem. Soc. 146, 20937–20944 (2024). 10.1021/jacs.4c05678 [DOI] [PubMed] [Google Scholar]
- 60.T. Jia, Y.-X. Li, X.-H. Ma, M.-M. Zhang, X.-Y. Dong et al., Atomically precise ultrasmall copper cluster for room-temperature highly regioselective dehydrogenative coupling. Nat. Commun. 14, 6877 (2023). 10.1038/s41467-023-42688-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.X. Xu, Y. Liu, F. Sun, Y. Jia, Q. Xu et al., Array-based clusters of copper with largely exposed metal sites for promoting catalysis. Chem. Mater. 35, 7588–7596 (2023). 10.1021/acs.chemmater.3c01277 [Google Scholar]
- 62.Y.-M. Wang, X.-C. Lin, K.-M. Mo, M. Xie, Y.-L. Huang et al., An atomically precise pyrazolate-protected copper nanocluster exhibiting exceptional stability and catalytic activity. Angew. Chem. Int. Ed. 62, e202218369 (2023). 10.1002/anie.202218369 [DOI] [PubMed] [Google Scholar]
- 63.G. Dong, Z. Pan, B. Han, Y. Tao, X. Chen et al., Multi-layer 3D chirality and double-helical assembly in a copper nanocluster with a triple-helical Cu15 core. Angew. Chem. Int. Ed. 62, e202302595 (2023). 10.1002/anie.202302595 [DOI] [PubMed] [Google Scholar]
- 64.X. Sun, Y. Wang, Q. Wu, Y.-Z. Han, X. Gong et al., Cu66 nanoclusters from hierarchical square motifs: synthesis, assembly, and catalysis. Aggregate, e651 (2024). 10.1002/agt2.651
- 65.C.-Y. Liu, T.-Y. Liu, Z.-J. Guan, S. Wang, Y.-Y. Dong et al., Dramatic difference between Cu20H8 and Cu20H9 Clusters in catalysis. CCS Chem. 6, 1581–1590 (2024). 10.31635/ccschem.023.202303448
- 66.B. Yan, X. You, X. Tang, J. Sun, Q. Xu et al., Carboxylate-protected “isostructural” Cu20 nanoclusters as a model system: carboxylate effect on controlling catalysis. Chem. Mater. 36, 1004–1012 (2024). 10.1021/acs.chemmater.3c03131 [Google Scholar]
- 67.Y. Li, R. Jin, Seeing ligands on nanoclusters and in their assemblies by X-ray crystallography: atomically precise nanochemistry and beyond. J. Am. Chem. Soc. 142, 13627–13644 (2020). 10.1021/jacs.0c05866 [DOI] [PubMed] [Google Scholar]
- 68.B. Zhang, J. Chen, Y. Cao, O.J.H. Chai, J. Xie, Ligand design in ligand-protected gold nanoclusters. Small 17, e2004381 (2021). 10.1002/smll.202004381 [DOI] [PubMed] [Google Scholar]
- 69.Z.-J. Guan, R.-L. He, S.-F. Yuan, J.-J. Li, F. Hu et al., Ligand engineering toward the trade-off between stability and activity in cluster catalysis. Angew. Chem. Int. Ed. 61, e202116965 (2022). 10.1002/anie.202116965 [DOI] [PubMed] [Google Scholar]
- 70.W. Fei, Y. Tao, Y. Qiao, S.-Y. Tang, M.-B. Li, Structural modification and performance regulation of atomically precise metal nanoclusters by phosphine. Polyoxometalates 2, 9140043 (2023). 10.26599/pom.2023.9140043
- 71.S.-F. Yuan, W.-D. Liu, C.-Y. Liu, Z.-J. Guan, Q.-M. Wang, Nitrogen donor protection for atomically precise metal nanoclusters. Chemistry 28, e202104445 (2022). 10.1002/chem.202104445 [DOI] [PubMed] [Google Scholar]
- 72.Z. Lei, X.-K. Wan, S.-F. Yuan, Z.-J. Guan, Q.-M. Wang, Alkynyl approach toward the protection of metal nanoclusters. Acc. Chem. Res. 51, 2465–2474 (2018). 10.1021/acs.accounts.8b00359 [DOI] [PubMed] [Google Scholar]
- 73.L. Chen, L. Wang, Q. Shen, Y. Liu, Z. Tang, All-alkynyl-protected coinage metal nanoclusters: from synthesis to electrocatalytic CO2 reduction applications. Mater. Chem. Front. 7, 1482–1495 (2023). 10.1039/d2qm01282k [Google Scholar]
- 74.R.W. Huang, J. Yin, C. Dong, A. Ghosh, M.J. Alhilaly et al., Cu81(PhS)46(tBuNH2)10(H)32]3+ reveals the coexistence of large planar cores and hemispherical shells in high-nuclearity copper nanoclusters. J. Am. Chem. Soc. 142, 8696–8705 (2020). 10.1021/jacs.0c00541 [DOI] [PubMed] [Google Scholar]
- 75.P.-P. Sun, B.-L. Han, H.-G. Li, C.-K. Zhang, X. Xin et al., Real-time fluorescent monitoring of kinetically controlled supramolecular self-assembly of atom-precise Cu8 nanocluster. Angew. Chem. Int. Ed. 61, e202200180 (2022). 10.1002/anie.202200180 [DOI] [PubMed] [Google Scholar]
- 76.Q.-Q. Huang, M.-Y. Hu, Y.-L. Li, N.-N. Chen, Y. Li et al., Novel ultrabright luminescent copper nanoclusters and application in light-emitting devices. Chem. Commun. 57, 9890–9893 (2021). 10.1039/d1cc03799d [DOI] [PubMed] [Google Scholar]
- 77.M.K. Osei, S. Mirzaei, X. Bogetti, E. Castro, M.A. Rahman et al., Synthesis of square planar Cu4 clusters. Angew. Chem. Int. Ed. 61, e202209529 (2022). 10.1002/anie.202209529 [DOI] [PubMed] [Google Scholar]
- 78.C.-Y. Liu, S.-F. Yuan, S. Wang, Z.-J. Guan, D.-E. Jiang et al., Structural transformation and catalytic hydrogenation activity of amidinate-protected copper hydride clusters. Nat. Commun. 13, 2082 (2022). 10.1038/s41467-022-29819-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.H.-Y. Zhuo, H.-F. Su, Z.-Z. Cao, W. Liu, S.-A. Wang et al., High-nuclear organometallic copper(I)-alkynide clusters: thermochromic near-infrared luminescence and solution stability. Chemistry 22, 17619–17626 (2016). 10.1002/chem.201603797 [DOI] [PubMed] [Google Scholar]
- 80.M.-M. Zhang, X.-Y. Dong, Z.-Y. Wang, H.-Y. Li, S.-J. Li et al., AIE triggers the circularly polarized luminescence of atomically precise enantiomeric copper(I) alkynyl clusters. Angew. Chem. Int. Ed. 59, 10052–10058 (2020). 10.1002/anie.201908909 [DOI] [PubMed] [Google Scholar]
- 81.M. Brust, M. Walker, D. Bethell, D.J. Schiffrin, R. Whyman, Synthesis of thiol-derivatised gold nanoparticles in a two-phase Liquid–Liquid system. J. Chem. Soc., Chem. Commun., 801–802 (1994). 10.1039/c39940000801
- 82.M. Brust, J. Fink, D. Bethell, D.J. Schiffrin, C. Kiely, Synthesis and reactions of functionalised gold nanoparticles. J. Chem. Soc., Chem. Commun., 1655 (1995). 10.1039/c39950001655
- 83.R. Jin, H. Qian, Z. Wu, Y. Zhu, M. Zhu et al., Size focusing: a methodology for synthesizing atomically precise gold nanoclusters. J. Phys. Chem. Lett. 1, 2903–2910 (2010). 10.1021/jz100944k [Google Scholar]
- 84.Q. Yao, T. Chen, X. Yuan, J. Xie, Toward total synthesis of thiolate-protected metal nanoclusters. Acc. Chem. Res. 51, 1338–1348 (2018). 10.1021/acs.accounts.8b00065 [DOI] [PubMed] [Google Scholar]
- 85.X. Yuan, L.L. Chng, J. Yang, J.Y. Ying, Miscible-solvent-assisted two-phase synthesis of monolayer-ligand-protected metal nanoclusters with various sizes. Adv. Mater. 32, e1906063 (2020). 10.1002/adma.201906063 [DOI] [PubMed] [Google Scholar]
- 86.Z. Wang, X. Pan, S. Qian, G. Yang, F. Du et al., The beauty of binary phases: a facile strategy for synthesis, processing, functionalization, and application of ultrasmall metal nanoclusters. Coord. Chem. Rev. 438, 213900 (2021). 10.1016/j.ccr.2021.213900 [Google Scholar]
- 87.Y. Li, M. Zhou, R. Jin, Programmable metal nanoclusters with atomic precision. Adv. Mater. 33, e2006591 (2021). 10.1002/adma.202006591 [DOI] [PubMed] [Google Scholar]
- 88.X. Ma, G. Ma, L. Qin, G. Chen, S. Chen et al., A synchronous nucleation and passivation strategy for controllable synthesis of Au36(PA)24: unveiling the formation process and the role of Au22(PA)18 intermediate. Sci. China Chem. 63, 1777–1784 (2020). 10.1007/s11426-020-9819-4 [Google Scholar]
- 89.X. Ma, Z. Tang, L. Qin, J. Peng, L. Li et al., Unravelling the formation mechanism of alkynyl protected gold clusters: a case study of phenylacetylene stabilized Au144 molecules. Nanoscale 12, 2980–2986 (2020). 10.1039/c9nr10930g [DOI] [PubMed] [Google Scholar]
- 90.X. Ma, F. Sun, L. Qin, Y. Liu, X. Kang et al., Electrochemical CO2 reduction catalyzed by atomically precise alkynyl-protected Au7Ag8, Ag9Cu6, and Au2Ag8Cu5 nanoclusters: probing the effect of multi-metal core on selectivity. Chem. Sci. 13, 10149–10158 (2022). 10.1039/d2sc02886g [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Y. Tang, F. Sun, X. Ma, L. Qin, G. Ma et al., Alkynyl and halogen co-protected (AuAg)44 nanoclusters: a comparative study on their optical absorbance, structure, and hydrogen evolution performance. Dalton Trans. 51, 7845–7850 (2022). 10.1039/d2dt00634k [DOI] [PubMed] [Google Scholar]
- 92.G. Ma, Y. Tang, L. Chen, L. Qin, Q. Shen et al., A homoleptic alkynyl-protected Au(I)9-Ag(I)9 cluster: structure analysis, optical properties, and catalytic implications. Eur. J. Inorg. Chem. 2022, e202200176 (2022). 10.1002/ejic.202200176 [Google Scholar]
- 93.L. Qin, F. Sun, Z. Gong, G. Ma, Y. Chen et al., Electrochemical NO3- reduction catalyzed by atomically precise Ag30Pd4 bimetallic nanocluster: synergistic catalysis or tandem catalysis? ACS Nano 17, 12747–12758 (2023). 10.1021/acsnano.3c03692 [DOI] [PubMed] [Google Scholar]
- 94.L. Chen, F. Sun, Q. Shen, L. Wang, Y. Liu et al., Structure, optical properties, and catalytic applications of alkynyl-protected M4Rh2 (M = Ag/Au) nanoclusters with atomic precision: a comparative study. Dalton Trans. 52, 9441–9447 (2023). 10.1039/d3dt01326j [DOI] [PubMed] [Google Scholar]
- 95.L. Wang, L. Chen, L. Qin, Y. Liu, Z. Tang, Alkynyl-protected Ag20 Rh2 nanocluster with atomic precision: structure analysis and tri-functionality catalytic application. Chem. Asian J. 18, e202300685 (2023). 10.1002/asia.202300685 [DOI] [PubMed] [Google Scholar]
- 96.X. Ma, Y. Tang, G. Ma, L. Qin, Z. Tang, Controllable synthesis and formation mechanism study of homoleptic alkynyl-protected Au nanoclusters: recent advances, grand challenges, and great opportunities. Nanoscale 13, 602–614 (2021). 10.1039/d0nr07499c [DOI] [PubMed] [Google Scholar]
- 97.X. Yuan, B. Zhang, Z. Luo, Q. Yao, D.T. Leong et al., Balancing the rate of cluster growth and etching for gram-scale synthesis of thiolate-protected Au25 nanoclusters with atomic precision. Angew. Chem. Int. Ed. 53, 4623–4627 (2014). 10.1002/anie.201311177 [DOI] [PubMed] [Google Scholar]
- 98.Z. Wu, J. Suhan, R. Jin, One-pot synthesis of atomically monodisperse, thiol-functionalized Au25nanoclusters. J. Mater. Chem. 19, 622–626 (2009). 10.1039/b815983a [Google Scholar]
- 99.A.C. Dharmaratne, T. Krick, A. Dass, Nanocluster size evolution studied by mass spectrometry in room temperature Au25(SR)18 synthesis. J. Am. Chem. Soc. 131, 13604–13605 (2009). 10.1021/ja906087a [DOI] [PubMed] [Google Scholar]
- 100.S. Ito, S. Takano, T. Tsukuda, Alkynyl-protected Au22(C≡CR)18 clusters featuring new interfacial motifs and R-dependent photoluminescence. J. Phys. Chem. Lett. 10, 6892–6896 (2019). 10.1021/acs.jpclett.9b02920 [DOI] [PubMed] [Google Scholar]
- 101.Z.-J. Guan, F. Hu, J.-J. Li, Z.-R. Wen, Y.-M. Lin et al., Isomerization in alkynyl-protected gold nanoclusters. J. Am. Chem. Soc. 142, 2995–3001 (2020). 10.1021/jacs.9b11836 [DOI] [PubMed] [Google Scholar]
- 102.Y.-J. Zhong, J.-H. Liao, T.-H. Chiu, S. Kahlal, C.-J. Lin et al., A two-electron silver superatom isolated from thermally induced internal redox reaction of A silver(I) hydride. Angew. Chem. Int. Ed. 60, 12712–12716 (2021). 10.1002/anie.202100965 [DOI] [PubMed] [Google Scholar]
- 103.K.K. Chakrahari, J.H. Liao, S. Kahlal, Y.C. Liu, M.H. Chiang et al., Cu13{S2 CNn Bu2}6 (acetylide)4]+: a two-electron superatom. Angew. Chem. Int. Ed. 55, 14704–14708 (2016). 10.1002/anie.201608609 [DOI] [PubMed] [Google Scholar]
- 104.F. Sun, Q. Tang, D-e. Jiang Theoretical advances in understanding and designing the active sites for hydrogen evolution reaction. ACS Catal. 12, 8404–8433 (2022). 10.1021/acscatal.2c02081
- 105.Y. Wang, X. Zheng, D. Wang, Design concept for electrocatalysts. Nano Res. 15, 1730–1752 (2022). 10.1007/s12274-021-3794-0 [Google Scholar]
- 106.X.L. Tian, J.Y. Li, L.X. Zhang, Y. Chen, Structural chemistry in energy relevant-catalysis reaction. Chin. J. Struct. Chem. 41, 2207001–2207001 (2022). 10.14102/j.cnki.0254-5861.2022-0160
- 107.Z.W. Seh, J. Kibsgaard, C.F. Dickens, I. Chorkendorff, J.K. Nørskov et al., Combining theory and experiment in electrocatalysis: insights into materials design. Science 355, eaad4998 (2017). 10.1126/science.aad4998 [DOI] [PubMed]
- 108.O.S. Bushuyev, P. De Luna, C.T. Dinh, L. Tao, G. Saur et al., What should we make with CO2 and how can we make it? Joule 2, 825–832 (2018). 10.1016/j.joule.2017.09.003 [Google Scholar]
- 109.M. Lu, M. Zhang, J. Liu, Y. Chen, J.-P. Liao et al., Covalent organic framework based functional materials: important catalysts for efficient CO2 utilization. Angew. Chem. Int. Ed. 61, e202200003 (2022). 10.1002/anie.202200003 [DOI] [PubMed] [Google Scholar]
- 110.Y. Quan, J. Zhu, G. Zheng, Electrocatalytic reactions for converting CO2 to value-added products. Small Sci. 1, 2100043 (2021). 10.1002/smsc.202100043 [Google Scholar]
- 111.Y. Li, T.-B. Lu, Coupling electrochemical CO2 reduction with value-added anodic oxidation reactions: progress and challenges. Mater. Chem. Front. 8, 341–353 (2024). 10.1039/d3qm00625e [Google Scholar]
- 112.G. Ma, L. Qin, Y. Liu, H. Fan, L. Qiao et al., A review of CO2 reduction reaction catalyzed by atomical-level Ag nanomaterials: atom-precise nanoclusters and atomically dispersed catalysts. Surf. Interfaces 36, 102555 (2023). 10.1016/j.surfin.2022.102555 [Google Scholar]
- 113.G. Wang, J. Chen, Y. Ding, P. Cai, L. Yi et al., Electrocatalysis for CO2 conversion: from fundamentals to value-added products. Chem. Soc. Rev. 50, 4993–5061 (2021). 10.1039/d0cs00071j [DOI] [PubMed] [Google Scholar]
- 114.H. Liu, Y. Zhu, J. Ma, Z. Zhang, W. Hu, Atomically thin catalysts: recent advances in atomic-level engineering of nanostructured catalysts for electrochemical CO2 reduction. Adv. Funct. Mater. 30, 2070107 (2020). 10.1002/adfm.202070107 [Google Scholar]
- 115.B. Pan, Y. Wang, Y. Li, Understanding and leveraging the effect of cations in the electrical double layer for electrochemical CO2 reduction. Chem Catal. 2, 1267–1276 (2022). 10.1016/j.checat.2022.03.012 [Google Scholar]
- 116.L. Zhou, R. Lv, Rational catalyst design and interface engineering for electrochemical CO2 reduction to high-valued alcohols. J. Energy Chem. 70, 310–331 (2022). 10.1016/j.jechem.2022.02.033 [Google Scholar]
- 117.S. Ji, Y. Chen, X. Wang, Z. Zhang, D. Wang et al., Chemical synthesis of single atomic site catalysts. Chem. Rev. 120, 11900–11955 (2020). 10.1021/acs.chemrev.9b00818 [DOI] [PubMed] [Google Scholar]
- 118.J. Duan, T. Liu, Y. Zhao, R. Yang, Y. Zhao et al., Active and conductive layer stacked superlattices for highly selective CO2 electroreduction. Nat. Commun. 13, 2039 (2022). 10.1038/s41467-022-29699-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Y. Zheng, A. Vasileff, X. Zhou, Y. Jiao, M. Jaroniec et al., Understanding the roadmap for electrochemical reduction of CO2 to multi-carbon oxygenates and hydrocarbons on copper-based catalysts. J. Am. Chem. Soc. 141, 7646–7659 (2019). 10.1021/jacs.9b02124 [DOI] [PubMed] [Google Scholar]
- 120.J. Wang, H.-Y. Tan, Y. Zhu, H. Chu, H.M. Chen, Linking the dynamic chemical state of catalysts with the product profile of electrocatalytic CO2 reduction. Angew. Chem. Int. Ed. 60, 17254–17267 (2021). 10.1002/anie.202017181 [DOI] [PubMed] [Google Scholar]
- 121.S. Liu, B. Zhang, L. Zhang, J. Sun, Rational design strategies of Cu-based electrocatalysts for CO2 electroreduction to C2 products. J. Energy Chem. 71, 63–82 (2022). 10.1016/j.jechem.2022.03.041 [Google Scholar]
- 122.Y. Fang, L. Han, S. Che, Electrocatalytic reduction of CO2 on chiral Cu surfaces. Chin. J. Struct. Chem. 42, 100107 (2023). 10.1016/j.cjsc.2023.100107 [Google Scholar]
- 123.S. Nitopi, E. Bertheussen, S.B. Scott, X. Liu, A.K. Engstfeld et al., Progress and perspectives of electrochemical CO2 reduction on copper in aqueous electrolyte. Chem. Rev. 119, 7610–7672 (2019). 10.1021/acs.chemrev.8b00705 [DOI] [PubMed] [Google Scholar]
- 124.T.K. Todorova, M.W. Schreiber, M. Fontecave, Mechanistic understanding of CO2 reduction reaction (CO2RR) toward multicarbon products by heterogeneous copper-based catalysts. ACS Catal. 10, 1754–1768 (2020). 10.1021/acscatal.9b04746 [Google Scholar]
- 125.D. Karapinar, C.E. Creissen, J.G. Rivera de la Cruz, M.W. Schreiber, M. Fontecave, Electrochemical CO2 reduction to ethanol with copper-based catalysts. ACS Energy Lett. 6, 694–706 (2021). 10.1021/acsenergylett.0c02610 [Google Scholar]
- 126.A. Ma, J. Wang, Y. Wang, Y. Zuo, Y. Ren et al., Atomically precise M15 (M = Au/Ag/Cu) alloy nanoclusters: structural analysis, optical and electrocatalytic CO2 reduction properties. Polyoxometalates 3, 9140054 (2024). 10.26599/pom.2024.9140054
- 127.Y.-F. Lu, L.-Z. Dong, J. Liu, R.-X. Yang, J.-J. Liu et al., Predesign of catalytically active sites via stable coordination cluster model system for electroreduction of CO2 to ethylene. Angew. Chem. Int. Ed. 60, 26210–26217 (2021). 10.1002/anie.202111265 [DOI] [PubMed] [Google Scholar]
- 128.C.P. Wan, J.D. Yi, R. Cao, Y.B. Huang Conductive metal/covalent organic frameworks for CO2 electroreduction. Chin. J. Struct. Chem. 41, 2205001–2205014 (2022). 10.14102/j.cnki.0254-5861.2022-0075
- 129.K. Yao, J. Li, H. Wang, R. Lu, X. Yang et al., Mechanistic insights into OC-COH coupling in CO2 electroreduction on fragmented copper. J. Am. Chem. Soc. 144, 14005–14011 (2022). 10.1021/jacs.2c01044 [DOI] [PubMed] [Google Scholar]
- 130.Y.-L. Yang, Y.-R. Wang, L.-Z. Dong, Q. Li, L. Zhang et al., A honeycomb-like porous crystalline hetero-electrocatalyst for efficient electrocatalytic CO2 reduction. Adv. Mater. 34, 2206706 (2022). 10.1002/adma.202206706 [DOI] [PubMed] [Google Scholar]
- 131.Y.-M. Wang, J. Cai, Q.-Y. Wang, Y. Li, Z. Han et al., Electropolymerization of metal clusters establishing a versatile platform for enhanced catalysis performance. Angew. Chem. Int. Ed. 61, e202114538 (2022). 10.1002/anie.202114538 [DOI] [PubMed] [Google Scholar]
- 132.G. Ma, F. Sun, L. Qiao, Q. Shen, L. Wang et al., Atomically precise alkynyl-protected Ag20Cu12 nanocluster: structure analysis and electrocatalytic performance toward nitrate reduction for NH3 synthesis. Nano Res. 16, 10867–10872 (2023). 10.1007/s12274-023-5885-6 [Google Scholar]
- 133.S. Wu, N. Salmon, M.M.-J. Li, R. Bañares-Alcántara, S.C.E. Tsang, Energy decarbonization via green H2 or NH3? ACS Energy Lett. 7, 1021–1033 (2022). 10.1021/acsenergylett.1c02816 [Google Scholar]
- 134.A. Valera-Medina, F. Amer-Hatem, A.K. Azad, I.C. Dedoussi, M. de Joannon et al., Review on ammonia as a potential fuel: from synthesis to economics. Energy Fuels 35, 6964–7029 (2021). 10.1021/acs.energyfuels.0c03685 [Google Scholar]
- 135.C. Lv, J. Liu, C. Lee, Q. Zhu, J. Xu et al., Emerging p-block-element-based electrocatalysts for sustainable nitrogen conversion. ACS Nano 16, 15512–15527 (2022). 10.1021/acsnano.2c07260 [DOI] [PubMed] [Google Scholar]
- 136.X. Zhang, Y. Wang, C. Liu, Y. Yu, S. Lu et al., Recent advances in non-noble metal electrocatalysts for nitrate reduction. Chem. Eng. J. 403, 126269 (2021). 10.1016/j.cej.2020.126269 [Google Scholar]
- 137.L. Gu, H. Luo, Y. Zhang, Y. Cong, M. Kuang et al., Engineering interfacial architectures toward nitrate electrocatalysis and nitrogen neutral cycle. Mater. Chem. Front. 8, 1015–1035 (2024). 10.1039/d3qm01038d [Google Scholar]
- 138.S. Zhang, Y. Zha, Y. Ye, K. Li, Y. Lin et al., Oxygen-coordinated single Mn sites for efficient electrocatalytic nitrate reduction to ammonia. Nano-Micro Lett. 16, 9 (2023). 10.1007/s40820-023-01217-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.G. Luo, M. Song, Q. Zhang, L. An, T. Shen et al., Advances of synergistic electrocatalysis between single atoms and nanoparticles/clusters. Nano-Micro Lett. 16, 241 (2024). 10.1007/s40820-024-01463-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.T. Wang, H.-J. Wang, J.-S. Lin, J.-L. Yang, F.-L. Zhang et al., Plasmonic photocatalysis: Mechanism, applications and perspectives. Chin. J. Struct. Chem. 42, 100066 (2023). 10.1016/j.cjsc.2023.100066 [Google Scholar]
- 141.K. Sun, Y. Qian, H.-L. Jiang, Metal-organic frameworks for photocatalytic water splitting and CO2 reduction. Angew. Chem. Int. Ed. 62, e202217565 (2023). 10.1002/anie.202217565 [DOI] [PubMed] [Google Scholar]
- 142.J. Chen, L. Lin, P. Lin, L. Xiao, L. Zhang et al., A direct Z-scheme Bi2WO6/La2Ti2O7 photocatalyst for selective reduction of CO2 to CO. Chin. J. Struct. Chem. 42, 100010 (2023). 10.1016/j.cjsc.2022.100010 [Google Scholar]
- 143.C. Li, J. Wang, L. Tong, Y. Wang, P. Zhang et al., Recent progress and challenges of photocatalytic CO2 conversion into value-added multi-carbon products. Coord. Chem. Rev. 502, 215623 (2024). 10.1016/j.ccr.2023.215623 [Google Scholar]
- 144.M. Sayed, J. Yu, G. Liu, M. Jaroniec, Non-noble plasmonic metal-based photocatalysts. Chem. Rev. 122, 10484–10537 (2022). 10.1021/acs.chemrev.1c00473 [DOI] [PubMed] [Google Scholar]
- 145.L. Zhang, J. Zhang, H. Yu, J. Yu, Emerging S-scheme photocatalyst. Adv. Mater. 34, e2107668 (2022). 10.1002/adma.202107668 [DOI] [PubMed] [Google Scholar]
- 146.C. Ban, Y. Duan, Y. Wang, J. Ma, K. Wang et al., Isotype heterojunction-boosted CO2 photoreduction to CO. Nano-Micro Lett. 14, 74 (2022). 10.1007/s40820-022-00821-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.M. Gao, F. Tian, X. Zhang, Z. Chen, W. Yang et al., Improved plasmonic hot-electron capture in Au nanoparticle/polymeric carbon nitride by Pt single atoms for broad-spectrum photocatalytic H2 evolution. Nano-Micro Lett. 15, 129 (2023). 10.1007/s40820-023-01098-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.M. Lu, M. Zhang, J. Liu, T.-Y. Yu, J.-N. Chang et al., Confining and highly dispersing single polyoxometalate clusters in covalent organic frameworks by covalent linkages for CO2 photoreduction. J. Am. Chem. Soc. 144, 1861–1871 (2022). 10.1021/jacs.1c11987 [DOI] [PubMed] [Google Scholar]
- 149.S. Tao, S. Wan, Q. Huang, C. Li, J. Yu et al., Molecular engineering of g-C3N4 with dibenzothiophene groups as electron donor for enhanced photocatalytic H2-production. Chin. J. Struct. Chem. 41, 48–54 (2022). 10.14102/j.cnki.0254-5861.2022-0068
- 150.J.G. Yu, X. Li,, Z.L. Jin,, H.Tang,, E.Z. Liu, Preface to solar photocatalysis. Chin. J. Struct. Chem. 41, 2206001–2206002 (2022). 10.14102/j.cnki.0254-5861.2022-0158
- 151.R. Jin, G. Li, S. Sharma, Y. Li, X. Du, Toward active-site tailoring in heterogeneous catalysis by atomically precise metal nanoclusters with crystallographic structures. Chem. Rev. 121, 567–648 (2021). 10.1021/acs.chemrev.0c00495 [DOI] [PubMed] [Google Scholar]
- 152.Y. Du, H. Sheng, D. Astruc, M. Zhu, Atomically precise noble metal nanoclusters as efficient catalysts: a bridge between structure and properties. Chem. Rev. 120, 526–622 (2020). 10.1021/acs.chemrev.8b00726 [DOI] [PubMed] [Google Scholar]
- 153.S. Wang, L. Tang, B. Cai, Z. Yin, Y. Li et al., Ligand modification of Au25 nanoclusters for near-infrared photocatalytic oxidative functionalization. J. Am. Chem. Soc. 144, 3787–3792 (2022). 10.1021/jacs.2c01570 [DOI] [PubMed] [Google Scholar]
- 154.Y. Liu, Y. Wang, N. Pinna, Atomically precise metal nanoclusters for photocatalytic water splitting. ACS Mater. Lett. 6, 2995–3006 (2024). 10.1021/acsmaterialslett.4c00622 [Google Scholar]
- 155.K.P.S. Cheung, S. Sarkar, V. Gevorgyan, Visible light-induced transition metal catalysis. Chem. Rev. 122, 1543–1625 (2022). 10.1021/acs.chemrev.1c00403 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.O. Reiser Shining light on copper: unique opportunities for visible-light-catalyzed atom transfer radical addition reactions and related processes. Acc. Chem. Res. 49, 1990–1996 (2016). 10.1021/acs.accounts.6b00296 [DOI] [PubMed]
- 157.A. Hossain, A. Bhattacharyya, O. Reiser, Copper’s rapid ascent in visible-light photoredox catalysis. Science 364, eaav9713 (2019). 10.1126/science.aav9713 [DOI] [PubMed]
- 158.Y. Yun, L. Li, M. Zhou, M. Li, N. Sun et al., Atomically precise coreless AuCu bimetallic nanoclusters for Ullmann C-O coupling. Nano Res. 16, 10756–10762 (2023). 10.1007/s12274-023-5755-2 [Google Scholar]
- 159.Z.-J. Guan, J.-J. Li, F. Hu, Q.-M. Wang, Structural engineering toward gold nanocluster catalysis. Angew. Chem. Int. Ed. 61, e202209725 (2022). 10.1002/anie.202209725 [DOI] [PubMed] [Google Scholar]
- 160.S. Zhao, R. Jin, R. Jin, Opportunities and challenges in CO2 reduction by gold- and silver-based electrocatalysts: from bulk metals to nanoparticles and atomically precise nanoclusters. ACS Energy Lett. 3, 452–462 (2018). 10.1021/acsenergylett.7b01104 [Google Scholar]
- 161.Z. Wu, D.-E. Jiang, A.K.P. Mann, D.R. Mullins, Z.-A. Qiao et al., Thiolate ligands as a double-edged sword for CO oxidation on CeO2 supported Au25(SCH2CH2Ph)18 nanoclusters. J. Am. Chem. Soc. 136, 6111–6122 (2014). 10.1021/ja5018706 [DOI] [PubMed] [Google Scholar]
- 162.Z. Lei, Q.-M. Wang, Homo and heterometallic gold(I) clusters with hypercoordinated carbon. Coord. Chem. Rev. 378, 382–394 (2019). 10.1016/j.ccr.2017.11.001 [Google Scholar]
- 163.A. Ghosh, O.F. Mohammed, O.M. Bakr, Atomic-level doping of metal clusters. Acc. Chem. Res. 51, 3094–3103 (2018). 10.1021/acs.accounts.8b00412 [DOI] [PubMed] [Google Scholar]
- 164.R. Chinchilla, C. Nájera, Recent advances in sonogashira reactions. Chem. Soc. Rev. 40, 5084–5121 (2011). 10.1039/c1cs15071e [DOI] [PubMed] [Google Scholar]
- 165.B.V. Rokade, J. Barker, P.J. Guiry, Development of and recent advances in asymmetric A3 coupling. Chem. Soc. Rev. 48, 4766–4790 (2019). 10.1039/c9cs00253g [DOI] [PubMed] [Google Scholar]
- 166.C. Zhang, C. Tang, N. Jiao, Recent advances in copper-catalyzed dehydrogenative functionalization via a single electron transfer (SET) process. Chem. Soc. Rev. 41, 3464–3484 (2012). 10.1039/c2cs15323h [DOI] [PubMed] [Google Scholar]
- 167.R.G. Pearson, Hard and soft acids and bases, HSAB, part 1: fundamental principles. J. Chem. Educ. 45(9), 581 (1968). 10.1021/ed045p581 [Google Scholar]
- 168.T. He, Z. Huang, S. Yuan, X.-L. Lv, X.-J. Kong et al., Kinetically controlled reticular assembly of a chemically stable mesoporous Ni(II)-pyrazolate metal–organic framework. J. Am. Chem. Soc. 142, 13491–13499 (2020). 10.1021/jacs.0c05074 [DOI] [PubMed] [Google Scholar]
- 169.P. Hervés, M. Pérez-Lorenzo, L.M. Liz-Marzán, J. Dzubiella, Y. Lu et al., Catalysis by metallic nanoparticles in aqueous solution: model reactions. Chem. Soc. Rev. 41, 5577–5587 (2012). 10.1039/c2cs35029g [DOI] [PubMed] [Google Scholar]
- 170.X. Kong, J. Liu, Influence of alumina binder content on catalytic performance of Ni/HZSM-5 for hydrodeoxygenation of cyclohexanone. Plos one 9, e101744 (2014). 10.1371/journal.pone.0101744 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Y. Kobayashi, S. Tada, M. Kondo, K. Fujiwara, H. Mizoguchi, Intermetallic YIr2 nanoparticles with negatively charged Ir active sites for catalytic hydrogenation of cyclohexanone to cyclohexanol. Catal. Sci. Technol. 12, 3088–3093 (2022). 10.1039/d2cy00198e [Google Scholar]
- 172.C. Sun, N. Mammen, S. Kaappa, P. Yuan, G. Deng et al., Atomically precise, thiolated copper-hydride nanoclusters as single-site hydrogenation catalysts for ketones in mild conditions. ACS Nano 13, 5975–5986 (2019). 10.1021/acsnano.9b02052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Y. Li, Z. Tang, P.N. Prasad, M.R. Knecht, M.T. Swihart, Peptide-mediated synthesis of gold nanoparticles: effects of peptide sequence and nature of binding on physicochemical properties. Nanoscale 6, 3165–3172 (2014). 10.1039/c3nr06201e [DOI] [PubMed] [Google Scholar]
- 174.G. Liao, Y. Gong, L. Zhong, J. Fang, L. Zhang et al., Unlocking the door to highly efficient Ag-based nanoparticles catalysts for NaBH4-assisted nitrophenol reduction. Nano Res. 12, 2407–2436 (2019). 10.1007/s12274-019-2441-5 [Google Scholar]
- 175.S.-F. Yuan, Z.-J. Guan, Q.-M. Wang, Identification of the active species in bimetallic cluster catalyzed hydrogenation. J. Am. Chem. Soc. 144, 11405–11412 (2022). 10.1021/jacs.2c04156 [DOI] [PubMed] [Google Scholar]
- 176.X. Chen, Z. Cai, X. Chen, M. Oyama, AuPd bimetallic nanoparticles decorated on graphene nanosheets: their green synthesis, growth mechanism and high catalytic ability in 4-nitrophenol reduction. J. Mater. Chem. A 2, 5668–5674 (2014). 10.1039/c3ta15141g [Google Scholar]
- 177.F. Hu, J.-J. Li, Z.-J. Guan, S.-F. Yuan, Q.-M. Wang, Formation of an alkynyl-protected Ag112 silver nanocluster as promoted by chloride released in situ from CH2Cl2. Angew. Chem. Int. Ed. 59, 5312–5315 (2020). 10.1002/anie.201915168 [DOI] [PubMed] [Google Scholar]
- 178.Q. Zhu, F. Zhang, Y. Huang, H. Xiao, L. Zhao et al., An all-round AI-Chemist with a scientific mind. Natl. Sci. Rev. 9, nwac190 (2022). 10.1093/nsr/nwac190 [DOI] [PMC free article] [PubMed]
- 179.D. Lockey, C. Mathis, H.N. Miras, L. Cronin, Investigating the autocatalytically driven formation of Keggin-based polyoxometalate clusters. Matter 5, 302–313 (2022). 10.1016/j.matt.2021.11.030 [Google Scholar]
- 180.X. Li, S. Takano, T. Tsukuda, Ligand effects on the hydrogen evolution reaction catalyzed by Au13 and Pt@Au12: alkynyl vs thiolate. J. Phys. Chem. C 125, 23226–23230 (2021). 10.1021/acs.jpcc.1c08197 [Google Scholar]
- 181.J.-I. Nishigaki, R. Tsunoyama, H. Tsunoyama, N. Ichikuni, S. Yamazoe et al., A new binding motif of sterically demanding thiolates on a gold cluster. J. Am. Chem. Soc. 134, 14295–14297 (2012). 10.1021/ja305477a [DOI] [PubMed] [Google Scholar]
- 182.Z. Liu, H. Tan, B. Li, Z. Hu, D.-E. Jiang et al., Ligand effect on switching the rate-determining step of water oxidation in atomically precise metal nanoclusters. Nat. Commun. 14, 3374 (2023). 10.1038/s41467-023-38914-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.S. Yoo, S. Yoo, G. Deng, F. Sun, K. Lee et al., Nanocluster surface microenvironment modulates electrocatalytic CO2 reduction. Adv. Mater. 36, e2313032 (2024). 10.1002/adma.202313032 [DOI] [PubMed] [Google Scholar]
- 184.W.-Q. Shi, L. Zeng, R.-L. He, X.-S. Han, Z.-J. Guan et al., Near-unity NIR phosphorescent quantum yield from a room-temperature solvated metal nanocluster. Science 383, 326–330 (2024). 10.1126/science.adk6628 [DOI] [PubMed] [Google Scholar]
- 185.S. Hossain, Y. Niihori, L.V. Nair, B. Kumar, W. Kurashige et al., Alloy clusters: precise synthesis and mixing effects. Acc. Chem. Res. 51, 3114–3124 (2018). 10.1021/acs.accounts.8b00453 [DOI] [PubMed] [Google Scholar]
- 186.K. Kwak, W. Choi, Q. Tang, M. Kim, Y. Lee et al., A molecule-like PtAu24(SC6H13)18 nanocluster as an electrocatalyst for hydrogen production. Nat. Commun. 8, 14723 (2017). 10.1038/ncomms14723 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.S. Li, D. Alfonso, A.V. Nagarajan, S.D. House, J.C. Yang et al., Monopalladium substitution in gold nanoclusters enhances CO2 electroreduction activity and selectivity. ACS Catal. 10, 12011–12016 (2020). 10.1021/acscatal.0c02266 [Google Scholar]
- 188.S. Li, A.V. Nagarajan, S. Zhao, G. Mpourmpakis, R. Jin, Understanding the single atom doping effects in oxygen reduction with atomically precise metal nanoclusters. J. Phys. Chem. C 125, 24831–24836 (2021). 10.1021/acs.jpcc.1c08356 [Google Scholar]
- 189.Z. Qin, S. Hu, W. Han, Z. Li, W.W. Xu et al., Tailoring optical and photocatalytic properties by single-Ag-atom exchange in Au13Ag12(PPh3))10Cl8 nanoclusters. Nano Res. 15, 2971–2976 (2022). 10.1007/s12274-021-3928-4
- 190.W.L. Zhou, J.J. Jiang, W.R. Cheng, H. Su, Q.H. Liu, In-situ synchrotron radiation infrared spectroscopic identification of reactive intermediates over multiphase electrocatalytic interfaces. Chin. J. Struct. Chem. 41, 2210004–2210015 (2022). 10.14102/j.cnki.0254-5861.2022-0083
- 191.L. Xie, C. Huang, Z. Liang, H. Wang, Z. Jiang et al., In-situ HP-STM and operando EC-STM studies of heterogeneous catalysis at interfaces. Chin. J. Struct. Chem. 41, 29–44 (2022). 10.14102/j.cnki.0254-5861.2022-0136
- 192.Z. Xiao, Y.-C. Huang, C.-L. Dong, C. Xie, Z. Liu et al., operando identification of the dynamic behavior of oxygen vacancy-rich Co3O4 for oxygen evolution reaction. J. Am. Chem. Soc. 142, 12087–12095 (2020). 10.1021/jacs.0c00257 [DOI] [PubMed] [Google Scholar]
- 193.Y. Yang, S. Louisia, S. Yu, J. Jin, I. Roh et al., operando studies reveal active Cu nanograins for CO2 electroreduction. Nature 614, 262–269 (2023). 10.1038/s41586-022-05540-0 [DOI] [PubMed] [Google Scholar]
- 194.M. Zhou, T. Higaki, G. Hu, M.Y. Sfeir, Y. Chen et al., Three-orders-of-magnitude variation of carrier lifetimes with crystal phase of gold nanoclusters. Science 364, 279–282 (2019). 10.1126/science.aaw8007 [DOI] [PubMed] [Google Scholar]
- 195.D.-E. Jiang, W. Chen, R.L. Whetten, Z. Chen, What protects the core when the thiolated Au cluster is extremely small? J. Phys. Chem. C 113, 16983–16987 (2009). 10.1021/jp906823d [Google Scholar]
- 196.F. Sun, L. Qin, Z. Tang, G. Deng, M.S. Bootharaju et al., -SR removal or -R removal? A mechanistic revisit on the puzzle of ligand etching of Au25(SR)18 nanoclusters during electrocatalysis. Chem. Sci. 14, 10532–10546 (2023). 10.1039/d3sc03018k [DOI] [PMC free article] [PubMed] [Google Scholar]