Zhimin Chen‡
a,
Dongjie Zuo‡b,
Lancheng Zhao‡c,
Yuping Chena,
Fang Suna,
Likai Wang*c,
Hui Shen
*b and
Qing Tang
*a
aSchool of Chemistry and Chemical Engineering, Chongqing Key Laboratory of Chemical Theory and Mechanism, Chongqing University, Chongqing 401331, China. E-mail: qingtang@cqu.edu.cn
bSchool of Energy Materials and Chemistry, Inner Mongolia University, Hohhot 010021, China. E-mail: shen@imu.edu.cn
cSchool of Chemistry and Chemical Engineering, Shandong University of Technology, Zibo, Shandong 255049, China. E-mail: lkwangchem@sdut.edu.cn
First published on 28th April 2025
Surface ligands play an important role in dictating the structure and catalytic properties of metal nanoclusters. Recently, a novel class of Au clusters protected by N-heterocyclic carbenes (NHCs) and halogens has been synthesized; however, the dynamic stability of the Au–NHCs/Au–halogen interface in real electrochemical environments as well as the influence of the ligand layer on the catalytic process remains obscure. Herein, we combined first-principles simulations with experiments to investigate the metal–ligand interface interaction of the classical [Au13(NHCMe)9Cl3]2+ cluster and its unique potential to promote electrocatalytic CO2 reduction to syngas. Our simulations revealed the facile shedding of chlorine ligands from the surface of the Au13 core upon electrochemical biasing, and the more negative the applied potential, the faster the kinetics of the Au–Cl bond cleavage. By contrast, the Au–NHC interface is highly stable, indicating the greater stability of Au–C bonds over the Au–Cl bonds under electrochemical conditions. Interestingly, the exposed icosahedral Au in dechlorinated [Au13(NHCMe)9Cl2]3+ cluster is capable of efficiently catalyzing electrochemical CO2 reduction to generate CO and H2 with comparable barriers in a wide potential range, showcasing its strong potential for syngas formation. Our predictions are further corroborated by experimental electrochemical data, where X-ray photoelectron spectroscopy (XPS) verified halogen stripping under acid or neutral media, and the activated Au13 cluster demonstrated enhanced catalytic efficacy for syngas formation with a CO:
H2 ratio of approximately 0.8 to 1.2 across a broad potential range of −0.50 to −1.20 V. This work reveals an exciting frontier in the understanding of ligand etching dynamics in NHC-protected metal nanoclusters, and particularly, the catalytic preference for syngas production is revealed for the first time in gold-based nanoclusters, which is distinctive from previously reported Au nanoclusters that mainly produce CO.
Among the numerous types of ligand-protected AuNCs, the structural and catalytic properties of AuNCs protected by thiolate, alkynyl and phosphine ligands have been extensively investigated, while the N-heterocyclic carbene (NHCs)-protected AuNCs are much less explored.21–27 The NHC ligands exhibit strong σ-donating and moderate π-accepting abilities, and have been increasingly appreciated as a promising alternative to thiols for surface stabilization of metal surfaces and nanoclusters. In 2019, Crudden and co-workers made a seminal breakthrough in obtaining the first atomically precise [Au13(NHC)9Cl3]2+ nanoclusters protected by NHCs and halogen ligands via the direct reduction of metal–carbene complexes.28,29 Since then, great progress has been made in the experimental synthesis of other NHC-protected AuNCs with metal core sizes of Au6, Au10, Au11, Au16, Au17, Au23, Au24, Au25 and Au44.4,30,31 Particularly, in recent experimental studies, Au13 clusters protected by bis-N-heterocyclic carbene (bisNHC) ligands, such as [Au13(bisNHC)4I4]+ and [Au13(bisNHC)5Cl2]3+, have also been characterized.15,32 According to the superatom model,33,34 the icosahedral Au135+ core has stable eight bonding electrons with a 1S21P6 electronic structure. The stability of these Au13 nanoclusters is enhanced by the presence of NHCs, but these ligands passivate the surface by imposing spatial constraints on the accessibility of the reactant molecules, thus blocking the availability of surface-active atoms. In this regard, the selective removal of ligands at local sites is considered to be crucial in boosting the catalytic activity of NHC-protected AuNCs. For example, Crudden et al. recently demonstrated that applying thermal treatment to the [Au13(bisNHC)5Cl2]3+ nanocluster results in gradual and partial removal of bisNHC ligands and leads to improved catalytic efficiency in electrocatalytic CO2 reduction compared to the diphosphine-stabilized analogues.15 This suggests that partial removal of the NHC ligand could effectively activate the Au13 clusters.
Besides thermal annealing, electrochemical biasing is another promising strategy to partially etch the surface ligands and expose the catalytically active metal sites.35 For example, recent studies have revealed that the thiolate-protected AuNCs such as Au25(SR)18 exhibit enhanced electrocatalytic activity after electrochemical activation,11,13,16 where the surface thiolate ligands can be selectively removed under electrochemical reduction conditions.36 Compared with the widely studied thiolate- and alkynyl ligand- protected AuNCs, the applications of emerging NHC-protected AuNCs in electrocatalysis have only recently become a topic of investigation. Particularly, the dynamic stability of the gold–NHCs interface in realistic electrochemical environments and the impact of the ligand layers on the active sites of the catalytic process have not been explored. To enhance the development of NHC-based AuNCs for future electrocatalytic applications, it is necessary to probe how the NHC-protected AuNCs interact with the electrolyte environment and provide a deep molecular-level understanding of the electrocatalytic mechanisms.
Motivated by these considerations, in this study, we selected the representative NHC and halogen co-protected [Au13(NHCMe)9Cl3]2+ (N,N′-dimethylbenzimidazolylidene, simplified as NHCMe) cluster as our research model, where the side groups on the N atom of NHC were simplified to –CH3 to reduce the computational cost. Considering solvation and potential effects, we first examined the dynamics of ligand etching on [Au13(NHCMe)9Cl3]2+ under electrochemical treatment, followed by identification of the active sites. After a series of ab initio molecular dynamics (AIMD) simulations, we found that the halogen chlorine atoms are favorably detached from the surface of the gold nucleus under applied reduction potential, such that the dechlorinated bare Au atoms serve as the active sites. In contrast, the Au–NHC interface is highly stable under electrochemical biasing, and the NHC ligands are stably bonded to the Au13 core during the AIMD simulations. A strong dynamic preference for breaking the Au–Cl bond is also observed in the phosphine analogue [Au13(PMe3)9Cl3]2+ cluster, where NHC ligands are replaced by phosphine ligands, indicating the stronger stability of Au–C or Au–P bonds over the Au–Cl bonds under electrochemical conditions. The catalytic properties of the dechlorinated Au13 cluster under near neutral conditions (pH = 7) were further examined using constrained dynamic simulations with a slow-growth method. Surprisingly, the computational results demonstrated that the dechlorinated [Au13(NHCMe)9Cl2]3+ cluster can effectively facilitate the electrochemical reduction of CO2 to generate CO and H2 with comparable energy barriers, indicating its great potential for syngas production. Our calculations are further supported by electrochemical experiments, where X-ray photoelectron spectroscopy (XPS) analysis revealed that the chlorine atoms in the [Au13(NHCMe)9Cl3]2+ cluster were stripped off under electrocatalytic conditions. Furthermore, the [Au13(NHCMe)9Cl3]2+ cluster, after ligand stripping at pH = 0 and URHE = −1.60 V, demonstrated enhanced catalytic efficacy for syngas formation in the CO2RR across a broad electrochemical potential range of −0.50 to −1.20 V, exhibiting a stable CO:
H2 ratio of approximately 0.8 to 1.2. Our work provides unprecedented atomic-scale insights into the ligand etching dynamics occurring on the surface of novel [Au13(NHCMe)9Cl3]2+ clusters and offers an effective strategy to facilitate syngas production.
URHE = UAg/AgCl + 0.197V + 0.0591 × pH |
The electrochemical desorption of Cl ligands on [Au13(NHC)9Cl3](PF6)2 NCs was conducted in 0.5 M H2SO4 electrolyte and 0.5 M Na2SO4 electrolyte. Chronopotentiometry tests were performed at voltages of U = −0.30 V (vs. RHE), U = −1.60 V (vs. RHE), U = −0.61 V (vs. RHE), and U = −1.30 V (vs. RHE), with each voltage test lasting for one hour.
In the electrocatalytic CO2RR process, the cathode chamber and anode chamber were separated by an anion exchange membrane. The electrolyte (0.5 M NaHCO3 or 1.0 M KOH) in both the cathode and anode compartments was recirculated using a peristaltic pump, with a rotation speed of 40 rpm. High-purity CO2 (99.9999%) gas was continuously purged into the gas chamber behind the gas-diffusion layer. The gas flow rate was 24 sccm, controlled by a mass flow controller (D08-1F, Beijing Sevenstar Flow Co., Ltd). The gas products were analyzed quantitatively and qualitatively with the help of a gas chromatograph (GC, Huaai 9560). The faradaic efficiencies (FE) for the formation of the products (CO and H2) were calculated as follows:
Linear sweep voltammetry (LSV) of the CO2RR was conducted in 1 M KOH or 0.5 M NaHCO3 solution saturated with either N2 or CO2, using a scan rate of 50 mV s−1. The electrochemical active surface area (ECSA) of the samples was quantified using Cdl obtained from cyclic voltammetry (CV) curves, with a scanning rate ranging from 40 to 120 mV s−1 in a sealed and gas-circulated H-cell in a 0.5 M KHCO3 solution.
For the neutral model at pH = 7, about 0, 2 and 4 additional electrons are introduced to reach the URHE values of 0.45 V, −0.61 V, and −1.30 V, respectively. As illustrated in Fig. 1a, our AIMD snapshots demonstrate that the [Au13(NHCMe)9Cl3]2+ cluster, when exposed to liquid water and potential at room temperature (300 K) undergoes spontaneous halogen shedding at lower reduction potentials. Moreover, the number of detached chlorine atoms increases under more negative potentials. In order to gain more detailed information of the ligand etching process, we focused on the atomic-level dynamics in the region surrounding the broken bond and labeled the key atoms involved (see Fig. S1 and S2a† for a detailed illustration of labelling). In the [Au13(NHCMe)9Cl3]2+ cluster, the metal core is protected by nine NHC ligands and three halogen atoms (Au8–Cl14, Au6–Cl16, and Au9–Cl15). The structure of the [Au13(NHCMe)9Cl3]2+ cluster remained unaltered after 7 ps of dynamic simulations at URHE = 0.45 V (Fig. 1a, left), which is consistent with the high stability of this cluster under solvation conditions at room temperature, as observed in the experiment.28 Interestingly, as the potential shifts to a negative value of URHE = −0.61 V (Fig. 1a, middle), it takes about 1.53 ps to reach an elongated Cl16⋯Au6 distance of 4.0 Å from an initial Au–Cl bond of about 2.5 Å (marked by the green star in Fig. 1b), which can be considered as the complete detachment of the Cl16 atom from the cluster surface (Fig. 1d). As the potential further decreases to URHE = −1.31 V (Fig. 1a, right), the time required for the large separation between Au6 and Cl16 is dramatically reduced to 0.27 ps (Fig. 1c), and meanwhile, another halogen Cl14 atom subsequently detaches from the surface Au8 atom at a time of 1.47 ps (Fig. 1e). This interesting phenomenon highlights the critical role of applied potential in affecting the detachment dynamics of halogen ligands, where the more negative the applied potential, the easier and faster the Au–Cl bond breaks. The relative distance between the key atoms involved in the halogen leaching indicates that the detached Cl atom quickly migrates away from the surface Au to the water layer and that the structure of the resultant dechlorinated [Au13(NHCMe)9Cl2]3+ cluster remains essentially unchanged throughout the simulations. In addition, we investigated the possibility of re-adsorption of leached chlorine ligands onto Au atoms under pH = 7 and URHE = −0.16 V condition using the slow-growth method in constrained AIMD simulations (SG-MD). The free energy curve demonstrated that the energy barrier (Ea) to be surmounted for the entire migratory adsorption process was 0.99 eV (Fig. S3†), suggesting that re-coordination of the leached chlorine ligand with the Au atom is challenging.
Furthermore, to elucidate the influence of the electrolyte environment on the ligand leaching dynamics in the [Au13(NHCMe)9Cl3]2+ cluster, we also performed constant potential AIMD simulations under acidic conditions (pH = 0) to monitor the structural evolution of the cluster at room temperature. Under acidic conditions, the introduction of 0e−, 2e− and 4e− would lead to electrode potentials (URHE) of around −0.09 V, −0.30 V and −1.60 V, respectively. The final snapshots after 7 ps of AIMD simulations at 300 K demonstrate that the cluster structure [Au13(NHCMe)9Cl3]2+ remains intact at URHE = −0.09 V (Fig. 2a, left), indicating the high structural stability of the cluster in the acid environment when the interfacial potential is small. However, upon the application of a more negative potential of −0.30 V and −1.60 V, it was observed that the [Au13(NHCMe)9Cl3]2+ cluster undergoes halogen leaching from the gold core, resulting in the removal of one (Fig. 2a, middle) and two (Fig. 2a, right) chlorine atoms from the surface shell layer, respectively. The detailed dynamic process of halogen detachment can be traced from the change in the relative distance between the Au and Cl atoms (Fig. 2b–e). When URHE is −0.30 V (Fig. 2b), the proton (H286) from H3O+ spontaneously attacks the electronegative Cl16 atom and weakens the interaction between Cl16 and Au6 atoms, separating the detached Cl16 atom from the Au13 core to form an HCl molecule at 0.46 ps which subsequently diffuses into the aqueous solution (Fig. 2d). Note that the generated HCl molecule would then release the proton and be stably adsorbed on the exposed surface Au6 atom at about 1.43 ps with a Au–H bond length of 1.68 Å. This facile adsorption of protons can be attributed to the modification of the charge state of the exposed Au6 site after chlorine removal, which will be discussed later. When the potential decreases to URHE = −1.60 V (Fig. 2c), the Cl16 atom would be first detached in a shorter time of 0.43 ps compared to the 0.46 ps taken at URHE = −0.30 V, accompanied by proton adsorption at the exposed Au6 site at 1.39 ps (Fig. 2e). Moreover, the second Cl14 atom separates from the surface Au8 at 1.99 ps and was completely detached from the cluster at about 6.00 ps. Furthermore, the introduction of 5e− in the acidic model resulted in an electrode potential of approximately −1.80 V. Following a 5 ps AIMD simulation, three chlorine ligands were found to completely detach from the surface of Au13 nuclei at 0.25 ps, 0.45 ps, and 2.45 ps, respectively (Fig. S4†). The shedding dynamics are analogous to those at lower potentials, where H3O+ in acidic solvents interacts with the chlorine atom to facilitate the leaching of the chlorine ligands. The process also involves proton transfer, with the excess proton finally adsorbed on the undercoordinated Au6 site. The above AIMD results under acidic conditions also revealed that the more negative the applied potential, the faster the kinetics of Au–Cl bond cleavage, where proton attack on halogens can accelerate the leaching process. Note that although the Cl ligands can be easily detached into solution, the NHC ligands remained tightly bound to the surface Au atoms of the Au13 core throughout the AIMD simulations. This indicates that the interaction between Au and the NHC ligand is very strong and dynamically very stable under electrochemical conditions.
For comparison, we additionally carried out the constant potential AIMD simulations of the phosphine analogue cluster [Au13(PMe3)9Cl3]2+ at 300 K under both acidic (pH = 0) and neutral conditions (pH = 7), where the NHC ligands in the pristine [Au13(NHCMe)9Cl2]3+ cluster are replaced by the phosphine ligands. In neutral conditions, the dynamic process of ligand leaching from this cluster displays great similarities to that observed in [Au13(NHCMe)9Cl2]3+ (Fig. S2c, d and S5†). Specifically, at URHE = 0.28 V, the ligand remains stably attached to the [Au13(PMe3)9Cl3]2+ cluster. As the applied potential becomes increasingly negative, a gradual detachment of one Cl (URHE = −0.50 V) and two Cl atoms (URHE = −0.67 V) from the surface Au is observed. Under acidic conditions (pH = 0), the statistical analysis of relative distances between representative atoms in the equilibrium AIMD simulations indicate that the protons from H3O+ can promote the leaching of Cl from the surface Au (Fig. S6a†). At URHE = −0.48 V, the proton (H389) first attacks Cl16, weakening the Au6–Cl16 bond and rapidly breaking it at 0.14 ps. Noteworthily, unlike what we have observed for the above [Au13(NHCMe)9Cl2]3+ cluster under acidic conditions, the proton does not transfer to the gold site (Au6) exposed by Cl removal in the [Au13(PMe3)9Cl3]2+ cluster but remains in the H3O+ form in the water layer (Fig. S6b and S6d†). When the reduction potential increased to −1.09 V (Fig. S6c and S6e†), we observed the facile removal of two chlorine atoms (Cl16 and Cl14) from the [Au13(PMe3)9Cl3]2+ cluster at 0.29 and 0.39 ps, respectively. In this case, a sufficiently negative applied potential can facilitate the proton transfer to the activated Au6, ultimately leading to the proton adsorption on Au6 at 0.56 ps. Throughout the overall AIMD simulations, the structure of the [Au13(PMe3)9Cl3]2+ cluster remains intact, and no elongation in the distance between the P atom of the phosphine ligand and the surface Au is observed, indicating the high dynamic stability of the Au–P bonds.
Therefore, our constant-potential AIMD simulations of [Au13(NHCMe)9Cl2]3+ and [Au13(PMe3)9Cl3]2+ clusters indicate that there is a strong dynamic preference for breaking the Au–Cl bond rather than the Au–C or Au–P bonds under electrochemical conditions. Noteworthily, previous studies have indicated that the electron cloud overlap between the d orbitals of the peripheral Au atom and the coordinated P or C atom is greater than that between Au and Cl, suggesting that the Au–C and Au–P bonds are stronger than the Au–Cl bonds.36 This also provides a rationale for the preferential removal of halogen over NHC and phosphine ligands. Moreover, under acidic conditions, the minimal spatial hindrance around the chlorine ligand allows enough space for protons to interact with Cl. Besides, the high electronegativity of chlorine also results in a strong affinity for proton. These, in turn, lead to the weakening of the Au–Cl bond, thereby promoting halogen shedding.
We further analyzed the changes in the Bader charge of the dechlorinated gold atoms after constant-potential AIMD simulations. Our analysis indicates that the protected gold atoms (Au6, Au8 and Au9) initially coordinated with chlorine exhibit a positive charge (0–0.1|e|), while the uncapped gold atoms, after the leaching of halogen ligands, undergo a charge transition from positive to negative (<−0.1|e|) (Fig. 3). This observation indicates that the dechlorinated gold atoms could serve as the catalytic active centers, potentially facilitating the activation of reactants in electrocatalytic processes such as the CO2RR and HER. This also provides a logical rationale for the observed adsorption affinity of the dechlorinated gold for proton in under acidic conditions. The Bader charges of the other Au atoms are provided in Tables S1 and S2.† It is noteworthy that the Au atoms coordinated to the C atom of the NHC ligands in the [Au13(NHCMe)9Cl3]2+ cluster are positively charged, whereas the Au atoms coordinated to the P atom of phosphine ligands in the [Au13(PMe3)9Cl3]2+ cluster are negatively charged. The difference in the charge states is consistent with the decreasing atomic electronegativity from C (2.5), Au (2.4), to P (2.1). Moreover, the Au1 atom situated at the core of the icosahedral Au13 nucleus exhibits a net negative charge under all electrochemical conditions. Overall, the electrochemical etching process results in a reduction in the coverage of surface Cl ligands and exposes the surface Au atoms, providing the possible catalytic sites. As Häkkinen et al. demonstrated, for the phosphine–halide-protected Au11(PH3)7Cl3 cluster, the removal of halogen atoms constitutes a pivotal step in the cluster activation process.54 Following the identification of the active site, the effect of halogen chloride removal on the catalytic properties of the [Au13(NHCMe)9Cl3]2+ cluster in hydrogen evolution and CO2 electroreduction is of particular interest and will be the subject of our subsequent investigations.
To understand how the applied potential influences the CO2 activation on the dechlorinated [Au13(NHCMe)9Cl2]3+ catalyst, we used constrained AIMD simulations by the slow-growth method (SG-MD) to construct the free energy profile of CO2 adsorption at different electrode potentials (Fig. 4a) at pH = 7 (the collective variable (CV) setting for CO2 activation was shown in Fig. S8a†). CO2 activation was deemed to be complete when the O–C–O angle reached equilibrium throughout the simulations (Fig. S9†). The CO2 molecule preferentially adsorbs at the top site of the exposed Au atom. Obviously, the energy barrier (ΔG‡) of CO2 adsorption has a strong dependence on the electrode potential. As the potential U shifted negatively from 0.48 V to −0.39 V, the energy barrier (ΔG‡) decreased from 0.82 eV to 0.056 eV, indicating that the negative electrode potential could promote the kinetic adsorption of CO2. In the inset of Fig. 4a, the conformation of adsorbed CO2 deviates from its pristine linear shape, bending into a chemisorbed V-shaped conformation under the applied potential. Furthermore, since CO2 activation entails the migration of reactants from the solvent to the Au surface, the difficulty of CO2 activation can be assessed based on the position of the transition state (TS) (Fig. 4b). At a potential of 0.48 V, the C–Au bond length at the TS is 1.89 Å, indicating that CO2 must move a long distance from solution to the active Au site to be activated. However, at a more negative potential of −0.39 V, the corresponding Au–C distance of TS increases to 2.56 Å, indicating that CO2 can be activated farther from Au, which is possibly due to the higher electron density on the surface of the [Au13(NHCMe)9Cl2]3+ cluster at lower potentials. Interestingly, we fitted the corresponding potential U and the reaction energy barriers ΔG‡ to get a linear correlation with R2 = 0.90, whereas a higher R2 value of 0.98 is achieved in a quadratic correlation (Fig. 4c). This observation could be ascribed to the configurational asymmetries in the potential energy surfaces (PES) of the initial and final states of CO2 activation, resulting in a significant deviation from the typically anticipated linear potential dependence.55 Such a quadratic type potential-dependent ΔG‡ relationship indicates that CO2 can be easily and effectively activated at potentials below −0.11 V, where the activation barrier is lower than 0.20 eV.
Similarly, we also examined the free energies of the subsequent CO2RR steps at different applied potentials, including *COOH formation (*CO2 + H2O + e− → *COOH + OH−, Fig. 5a), *CO formation (*COOH + e− → *CO + OH−, Fig. 5b) and CO desorption (Fig. 5c). Unlike CO2 activation, the conversion of *CO2 to *COOH involves a solvent-mediated proton transfer. As illustrated in the inset of Fig. 5a, the neighboring water molecules are involved in the formation of *COOH under neutral conditions. The water-mediated proton transfer also occurs during the formation of *CO (Fig. 5b), which is analogous to the process observed in the *CO2 to *COOH conversion. The CV settings along the reaction coordinate in *COOH and *CO formation are provided in Fig. S8b and S8c.† The last elementary step is the desorption of *CO from the active Au site, which determines its effective release from the active site and thus affects the overall catalytic performance. The corresponding CV setting of CO desorption in our calculation is defined as the distance between the center of mass of CO and the Au site (Fig. S8d†). In Fig. 5c, the length of the Au–C bond during steady-state CO adsorption is approximately 2.14 Å (IS). Throughout the free energy sampling, it was found that the free energy along the *CO desorption process increases with the elongation of the C–Au bond, which could be attributed to the diffusive resistance of the aqueous layer subsequent to CO desorption from the active site. Accordingly, an elongated C–Au bond length of 3 Å is deemed to signify the complete desorption of CO from the catalyst surface. The fitted potential-dependent energy barrier relationships for each elementary reaction are shown in Fig. 5d. Obviously, as the applied potential becomes more negative, the energy barriers for all the reaction steps reduce. Moreover, by considering all the reaction steps together, the critical rate-determining step and overall reaction difficulty of the CO2RR process can be identified. Among the four elementary steps, the conversion of *COOH to *CO presents a significant energy barrier (ΔG‡ > 0.8 eV) at typical working potentials (−1.5–0 V), making it a crucial rate-limiting step in the overall reaction process of CO2 electroreduction.
Furthermore, in competition with the CO2 reduction reaction, the hydrogen evolution reaction (HER) significantly impacts the selectivity of the NHC-stabilized Au13 catalyst. To probe this influence, an in-depth investigation of the free energy barrier for the HER under neutral conditions (pH = 7) was carried out using the same SG-MD approach. As shown in Fig. 6a, the energy barrier for proton adsorption from water cleavage in the Volmer step (H2O + e− + * → *H + OH−) shows a marked dependence on the applied potential. The free energy barrier, ΔG‡, decreases progressively from 1.29 eV to 0.34 eV as the electrode potential shifts from −0.28 V to −1.09 V. Analysis of the FS snapshot reveals that hydrogen adsorbs at the top site of the unsaturated, dechlorinated gold atoms. A quadratic model (R2 = 0.98) provides a better fit to the potential U and the energy barrier ΔG‡ compared to a linear model (R2 = 0.92) (Fig. 6b). Considering the involvement of water cleavage, the energy barrier for the initial Volmer step of *H formation is substantially higher than that observed for CO2 adsorption. In neutral or alkaline environments, the subsequent Heyrovsky step in the HER involves the combination of H from water cleavage with the adsorbed hydrogen to form hydrogen gas (*H + H2O + e− → * + H2 + OH−). The potential-dependent free energy change for the Heyrovsky step is illustrated in Fig. 6c. It is evident that as the applied potential becomes more negative, the activation barrier for this step decreases from 1.34 eV at 0.18 V to 0.77 eV at −1.73 V, indicating a direct correlation between the potential and the reaction kinetics. In this case, the quadratic correlation (R2 = 0.99) also provides a better fit to the potential U and the energy barrier ΔG‡ compared to the linear model (R2 = 0.95) (Fig. 6d). Upon collating the fitted quadratic plots of the two elementary steps (Fig. 6e), the rate-determining step for the HER becomes evident. Interestingly, the Volmer step is identified as the rate-determining step when the applied potential is higher than −0.33 V, whereas the Heyrovsky step becomes as the rate-determining step when the potential is lower than −0.33 V. This transition underscores the pivotal role of electrode potential in modulating the catalytic mechanism of the HER on the [Au13(NHCMe)9Cl2]3+ cluster catalyst.
Based on the established fitting quadratic relationships, a comparative analysis of the CO2RR and HER at different potentials under neutral codition is illustrated in Fig. 6f. This analysis reveals a crossover potential at Ucross = −0.58 V, where the free energy barrier of the rate-determining step for the CO2RR equals to that for the HER (ΔG‡ (CO2RR) = ΔG‡ (HER)). At URHE > −0.58 V, the energy barrier of the rate-determining step in the CO2RR is lower than that for the HER. While conversely, at URHE < −0.58 V, the energy barrier for the CO2RR becomes higher than that of the HER. As the potential deviates further into the negative (URHE < −0.58 V) or extreme positive range (URHE > −0.33 V), the disparity in the energy barrier between these two reactions is exacerbated, indicating a higher selectivity for HER when URHE < −0.58 V and a higher selectivity for CO2RR when URHE > −0.33 V. Intriguingly, we found that the energy barriers for CO2 reduction and hydrogen evolution become comparable within a specific potential range, suggesting that the [Au13(NHCMe)9Cl2]3+ catalyst holds great promise for syngas production. By defining a threshold for the barrier difference between the CO2RR and HER as |ΔG‡ (CO2RR) − ΔG‡ (HER)| ≤ 0.15 eV, the corresponding potential is located roughly between −1.0 V and −0.26 V, as illustrated in the shaded region of Fig. 6f. Within this potential range, the selective production of syngas or a mixture of CO and H2 can be precisely controlled by modulating the applied potential.
Moreover, to further verify our theoretical predictions regarding syngas production, we investigated whether the synthesized [Au13(NHC)9Cl3](PF6)2 cluster is efficient in the electrochemical reduction of CO2 to generate CO and H2. Fig. 8a shows the linear sweep voltammetry (LSV) curves of the [Au13(NHC)9Cl3](PF6)2 cluster under CO2 and N2 gas saturation conditions before and after activation at different potentials in 0.5 M sulfuric acid for one hour. The LSV results showed that the pristine and activated [Au13(NHC)9Cl3](PF6)2 clusters only facilitate the HER in a N2-saturated environment, whereas a shifted onset potential is observed in a CO2 atmosphere for both the unactivated and activated Au13 clusters, suggesting their capability for electrocatalytic CO2 reduction. The electrochemical tests were conducted in a 1 M KHCO3 flow cell (pH = 7). Noteworthily, the pretreated Au13 clusters at potentials of −0.30 V and −1.60 V under acidic conditions achieve significantly higher current density than that of the untreated one when tested in a CO2-saturated KHCO3 solution, indicating that the electrochemical activation may induce ligand removal to dramatically increase the catalytic activity. The gaseous products were analyzed by gas chromatography (GC), which demonstrated that CO and H2 are the primary products, while no liquid product was identified from nuclear magnetic resonance (NMR) spectroscopy. Fig. 8b and c display the faradaic efficiencies (FE) of gaseous products, CO and H2, at different potentials. The results indicate that Au13 pre-activated under URHE = −1.60 V exhibits higher CO2RR activity in the potential range of −0.70 to −1.20 V. In particular, compared to activation at URHE = −0.30 V, Au13 activated at URHE = −1.60 V also demonstrates a more balanced activity for both HER and CO2RR. Fig. 8d presents the CO:
H2 ratio distribution of CO2RR for Au13 treated at URHE = −1.60 V, where the stable CO
:
H2 ratio was about 0.8 to 1.1 in the potential range of −0.70 to −1.20 V (the shaded area in Fig. 8d), which represents an ideal ratio of syngas for the hydroformylation reaction in the manufacture of fine chemicals. Moreover, the fractional current densities for CO (jCO) and H2 (jH2) (Fig. 8e) clearly showed that H2 vs. CO selectivity increases with decreasing reduction potential (more negative potential), and jH2 gradually increases more rapidly than jCO, indicating enhanced selectivity towards the HER, which is consistent with the computational results shown in Fig. 6f. Fig. S11† shows the turnover frequency (TOF) of CO and H2 products. Notably, in the potential range of −0.7 to −1.1 V, the Au13 cluster activated at URHE = −1.60 V exhibits similar TOF values for both CO and H2. However, as the potential becomes more negative, the TOF of H2 shows a more pronounced increase, indicating that hydrogen evolution gradually becomes the dominant process. The intrinsic properties of the catalyst were also measured as shown in Fig. S12,† where the electrochemical active surface area (ECSA) was determined from Cdl. Obviously, the Au13 cluster treated at URHE = −1.60 V exhibits the largest Cdl value (0.99 mF cm−2), indicating a larger active surface area, which is consistent with its enhanced electrocatalytic performance. Additionally, the stability test (Fig. 8f) demonstrated that the Au13 catalyst pre-activated at URHE = −1.60 V could maintain a stable current density and CO
:
H2 ratio for over 10 h at −0.9 V without significant performance degradation. As evidenced by the transmission electron microscopy (TEM) analysis in Fig. S13,† the Au13 nanocluster remained well-dispersed and stably anchored on the carbon nanotubes after prolonged reaction, also indicating the structural robustness of this nanocluster under the reaction conditions. Furthermore, the Au13 cluster that had shed three chlorine ligands and was pre-treated at pH = 0 and URHE = −1.82 V was subjected to electrocatalytic testing in a 0.5 M KHCO3 solution. As demonstrated in Fig. S14,† the HER activity of the catalyst enhanced by up to 90%, and the CO
:
H2 ratio was significantly reduced to approximately 0.5. This suggests that the HER is more advantageous than CO2 reduction under the condition of the removal of three chlorine ligands.
Besides the electrochemical performance in a near-neutral environment, the electrocatalytic activity of the Au13 nanocluster under strong alkaline conditions was also investigated in 1 M KOH solution. Similarly, Fig. S15† shows that the Au13 after activation at URHE = −1.60 V exhibits higher faradaic efficiency compared to the initial and treated cluster at URHE = −0.30 V, indicating the enhanced CO2RR activity. The CO:
H2 ratio distribution shows that the Au13 cluster activated at URHE = −1.60 V maintains a CO
:
H2 ratio between 0.81 and 1.20 in the potential range of −0.50 to −1.10 V, making it suitable for syngas production in hydroformylation reactions. These electrochemical results under strong alkaline conditions are consistent with the performance under neutral conditions.
Overall, our experiments indicated that the activated [Au13(NHC)9Cl3](PF6)2 cluster demonstrated enhanced CO2RR activity towards syngas performance over a wide electrochemical potential, in good agreement with the theoretical predictions. Note that Au-based nanoclusters (e.g., Au25, Au38, and Au144)13,56 typically have high selectivity towards CO formation, often producing CO as the main reduction product in the electrocatalytic CO2RR process. To our best knowledge, this work revealed, for the first time, the high catalytic preference for syngas production in the atomically precise Au13 nanocluster. The distinctive electrocatalytic selectivity could be correlated to the unique icosahedral Au13 core, where the exposed surface Au atoms after electrochemical dechlorination exhibit high curvature, enhancing their reactivity and facilitating water dissociation, thus leading to the facile formation of H2 in addition to the typical CO product. In addition to the [Au13(NHC)9Cl3]2+ prototype cluster explored in this work, other atomically and compositionally monodisperse NHC-protected gold nanoclusters with different core sizes and atomic arrangements have been reported, and the dependence of their electrocatalytic performance on factors such as size, structure, shape and composition as well as the mechanistic insights deserves future investigations.
Footnotes |
† Electronic supplementary information (ESI) available: Additional simulation models, Bader charge analysis, reaction coordinates for SG-MD simulations, free energy profiles, and electrochemical characterization data. See DOI: https://doi.org/10.1039/d5sc00896d |
‡ These three authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |