Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Atomic-level engineering of single Ag1+ site distribution on titanium–oxo cluster surfaces to boost CO2 electroreduction

Ru-Xin Meng a, Lan-Cheng Zhao b, Li-Pan Luo c, Yi-Qi Tian a, Yong-Liang Shao d, Qing Tang *c, Likai Wang *b, Jun Yan a and Chao Liu *a
aCollege of Chemistry and Chemical Engineering, Central South University, Changsha 410083, Hunan, P. R. China. E-mail: chaoliu@csu.edu.cn
bSchool of Chemistry and Chemical Engineering, Shandong University of Technology, Zibo 255049, Shandong, P. R. China. E-mail: lkwangchem@sdut.edu.cn
cCollege of Chemistry and Chemical Engineering, Chongqing University, Chongqing, 400044, P. R. China. E-mail: qingtang@cqu.edu.cn
dSchool of Chemistry and Chemical Engineering, Lanzhou University, Lanzhou, 730000, P. R. China

Received 23rd October 2024 , Accepted 6th January 2025

First published on 22nd January 2025


Abstract

Precise control over the distribution of active metal sites on catalyst surfaces is essential for maximizing catalytic efficiency. Addressing the limitations of traditional cluster catalysts with core-embedded catalytic sites, this work presents a strategy to position catalytic sites on the surfaces of oxide clusters. We utilize a calixarene-stabilized titanium–oxo cluster (Ti12L6) as a scaffold to anchor Ag1+in situ, forming the unique nanocluster Ti12Ag4.5 with six surface-exposed Ag1+ sites. The in situ transformation from Ti12L6 into Ti12Ag4.5 clusters was traced through mass spectrometry, revealing a solvent-mediated dynamic process of disintegration and reassembly of the Ti12L6 macrocycle. The unique Ti12Ag4.5 cluster, featuring a surface-exposed catalytic site configuration, efficiently catalyzes the electroreduction of CO2 to CO over a broad potential window, achieving CO faradaic efficiencies exceeding 82.0% between −0.4 V and −1.8 V. Its catalytic performance surpasses that of bimetallic Ti2Ag2, which features a more conventional design with Ag1+ sites embedded within the cluster. Theoretical calculations indicate that the synergy between the titanium–oxo support and the single Ag1+ sites lowers the activation energy, facilitating the formation of the *COOH intermediate. This work reveals that engineered interactions between active surface metal and the oxide support could amplify catalytic activity, potentially defining a new paradigm in catalyst design.


Introduction

Within the intricate sphere of catalysis, the sophisticated interplay of size, morphology, and structural attributes of materials is recognized for its profound influence on catalytic efficacy.1–4 Recently developed single-atom catalysts (SACs), with isolated metal atoms dispersed on solid supports, are garnering significant attention due to their exceptional catalytic efficiency and ability to maximize the utilization of metal atoms.5–9 Oxide-supported SACs, in particular, are among the most thoroughly investigated systems.10 Advances in the fabrication of these materials are noteworthy, but achieving uniform dispersion of catalytic sites on oxide substrates remains a significant challenge.11 This requires not only advanced synthesis techniques but also sophisticated characterization methods.12 The complexity is further exacerbated by the absence of detailed atomic structures, which obscures the analysis of the metal coordination environment—a pivotal factor in determining the performance of these catalyst systems.13–15

To effectively tackle these challenges, it is crucial to develop molecular analogues for oxide-supported single-atom materials. Recent advances in TiO2 analogues,16–19 particularly through the development of crystalline titanium oxide clusters (TOCs), have facilitated comprehensive investigations of TiO2 structures and reactivities at the molecular level.20–24 These TOCs, when modified with catalytically active single metals, hold promise as effective molecular mimics for oxide-supported SACs. However, heterometal-doped TOCs are typically created via a one-pot solvothermal method, with heterometal sites embedded deep within the cluster core, often in fully coordinated states.25–29 Such configurations limit the interactions with reactants, exposing a significant shortfall in their capacity to act as true molecular proxies. This insight highlights the urgent need for a new class of crystalline, cluster-stabilized single-atom materials, designed to more precisely emulate the structural and reactive properties of oxide-supported SACs at a molecular level.30,31

Here, we present an approach designed to strategically place catalytic Ag1+ sites on the surfaces of oxide clusters (Scheme 1). Employing thiacalix[4]arene (TC4A) as a protective ligand,32–38 we meticulously engineered a titanium–oxo macrocycle, Ti12L6, through a one-step solvothermal method, which is enriched with surface O and S sites. We utilized Ti12L6 as a scaffold for the in situ loading of Ag1+ ions, revealing a solvent-mediated assembly process through mass spectrometry. In N,N-dimethylformamide (DMF) solution, the scaffold undergoes decomposition under the cleavage of Ag1+ ions; however, in acetonitrile, the scaffold remains stable and coordinates six accessible single Ag1+ sites on the cluster surface, ultimately transforming into a Ti12Ag4.5 cluster. Ti12Ag4.5 has proven to be an exceptional catalyst for the electrochemical reduction of CO2, exhibiting superior reactivity and CO selectivity compared to its bimetallic counterpart, Ti2Ag2, which possesses a more conventional structure with Ag1+ sites embedded within the cluster matrix. Notably, Ti12Ag4.5 exhibits high selectivity for CO across a wide voltage range, with CO faradaic efficiency (FECO) consistently exceeding 82.0% from −0.4 V to −1.8 V vs. RHE, reaching a peak FECO of 92.7%. At an overpotential of ∼−1.4 V, the system remained stable for continuous electrolysis over 11 h with a CO partial current density exceeding 100.0 mA cm−2, while the FECO consistently remains above 85.0%. We elucidated the reaction path using in situ ATR-SEIRAS technology and comprehensively calculated the Gibbs free energy changes for each elementary step of CO2 conversion to CO, highlighting the critical contribution of exposed Ag centers to the observed catalytic prowess.


image file: d4sc07186g-s1.tif
Scheme 1 Schematic representation of the cluster assembly featuring a Ti–oxo core with surface-exposed Ag1+ catalytic sites.

Results and discussion

One-step synthesis of titanium–oxo macrocyclic carriers

The synthetic route commenced with the reaction of Ti(OiPr)4 with TC4A in a solvent blend of CH3CN/CHCl3, maintained at 80 °C for three days. This process yielded a turbid yellow liquid. Extending the reaction duration by introducing acetone and continuing incubation at 80 °C for an additional three days led to the crystallization of yellow rhomboid crystals of Ti12L6. Interestingly, this reaction exhibits a high sensitivity to temperature changes. Upon heating to 130 °C, certain sulfur sites in TC4A oxidize, leading to the emergence of two distinct TOCs, Ti8L6 and Ti8L5. SCXRD analysis revealed that the Ti12L6 cluster has the composition of {H6Ti12O18(HTC4A)6}, featuring a core–shell architecture where a central Ti–oxo macrocycle is encapsulated by six TC4A molecules. Ti12L6 consists of six {Ti2O(TC4A)} units interconnected by twelve μ2-O2− (Fig. 1A). TC4A adopts a mono-cone configuration, coordinating with two Ti4+ through three phenoxide groups and two S atoms, thereby forming the {Ti2O(TC4A)} units. These units are geometrically arranged in a triangular pattern, bridged by five μ2-O linkages to form a trimeric {Ti6(TC4A)3}. This trimer further aggregates into the hexameric {Ti12(TC4A)6} structure via the integration of two additional μ2-O2− ions (Fig. 1B). This configuration yields a “donut” shape with an outer diameter of 3 nm, representing the largest TC4A-stabilized TOC reported to date (Fig. 1C). The structure of Ti8L6 closely resembles that of Ti12L6, with a central Ti8O8 ring encircled by six TC4A molecules (Fig. 1D). In contrast, the core of Ti8L5 consists of a Ti8O10 cluster, surrounded by five TC4A molecules arranged in a pentagonal bipyramidal configuration. In both structures, some of the TC4A molecules undergo oxidation, with one or two sulfur sites being converted to S[double bond, length as m-dash]O groups. Additionally, each structure has two Na+ ions adsorbed on its surface.
image file: d4sc07186g-f1.tif
Fig. 1 (A) The overall structure of Ti12L6; (B) the Ti12O18 core in Ti12L6; (C) the polyhedral arrangement of six calix moieties in Ti12L6; (D) the overall structure of Ti8L6; (E) the overall structure of Ti8L5.

Solvent-mediated assembly of Ti12Ag4.5 clusters

The surfaces of this series of calixarene-stabilized TOCs are endowed with abundant O/S sites. Specifically, in Ti8L6 and Ti8L5, these sites facilitate the adsorption of Na+ onto the cluster surfaces. This characteristic has inspired us to utilize these clusters as carriers for positioning Ag1+ ions. In our preliminary studies, we attempted to directly introduce Ag(I) salt into the reaction mixture for these clusters, but this approach resulted in the formation of crystals of Ti2Ag2 (Fig. 2A). Ti2Ag2 is a typical bimetallic cluster, with two Ag1+ and two Ti4+ ions arranged in a quadrilateral configuration, sandwiched between two oxidized-TC4A ligands (Fig. 2B). Each Ag1+ site maintains a coordination number of six, with the Ag1+ seamlessly integrated into the cluster core, effectively shielded from external exposure (Fig. 2C). The composition of Ti2Ag2 was determined using electrospray ionization mass spectrometry (ESI-MS) (Fig. S28), which revealed a +1 signal at m/z = 1794.18, corresponding to the species [HTi2Ag2O(O-TC4A)2]+.
image file: d4sc07186g-f2.tif
Fig. 2 (A) The overall structure of Ti2Ag2; (B) coordination mode of oxidized TC4A in Ti2Ag2. (C) Comparison of the positions of the Ag and Ti sites in Ti2Ag2.

Interestingly, using Ti12L6 crystals in reactions with Ag(I) salts yielded distinctly different outcomes. Introducing Ag2SO4 into DMF solution along with Ti12L6 crystals and allowing the mixture to react at 80 °C for three days resulted in the formation of the known clusters Ti2Ag2-DMF and Ti2Ag4.34 Subsequently, by switching the solvent to CH3CN, rhombic Ti12Ag4.5 crystals were obtained, emphasizing the critical role of solvent selection in steering the chemical pathway. Ti2Ag2-DMF and Ti2Ag4 have typical bimetallic configurations, with two or four Ag1+ ions embedded between two {Ti(TC4A)} units. In contrast, in Ti12Ag4.5, multiple Ag1+ ions are effectively anchored onto the surface of the Ti12 core without altering its intrinsic structure. SCXRD analysis revealed the structure of Ti12Ag4.5 to be [H1.5Ti12Ag4.5O18(HTC4A)6(CH3CN)4], with the cluster containing three crystallographically distinct sites for Ag. Structural analysis demonstrated that the Ti–oxo core of Ti12Ag4.5 closely mirrors the macrocyclic structure of Ti12L6, albeit with minor deviations (Fig. 3A). The Ag1+ sites are symmetrically divided and positioned on both the upper and lower facets of the Ti–oxo macrocycle (Fig. 4B). Ag1 and Ag2, coordinated through phenoxide, sulfur, and μ-O22−, are placed between two TC4A ligands, exhibiting coordination numbers of 3 and 4, respectively. The distances between Ag1/Ag2 and the Ti–oxo core range from 2.392 to 2.607 Å. Notably, the Ag3 site, with an O3N2 coordination environment, is defined by two phenoxide groups, one μ-O2, and two CH3CN, with Ag–O bond lengths ranging from 2.649 to 2.774 Å and Ag–N distances of 2.056 and 2.215 Å. The spatial arrangement of the Ag sites, especially the Ag3 sites proximal to the cluster surface as depicted in Fig. 3C, highlights their potential as catalytically active sites. This is particularly significant given the labile nature of the CH3CN ligand, which may facilitate dynamic catalytic processes. Sites Ag1 and Ag2 exhibit full occupancy, while Ag3 displays a partial occupancy of 0.25. As a result, the average number of Ag sites in the cluster is 4.5, which is confirmed by the mass spectrometry analysis of Ti12Ag4.5. Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS) analysis reveals six distinct peaks with a mass difference of 107.86, corresponding to the ion {H4Ti12Ag6O16(TC4A)6(CH2Cl2)}+ (Fig. 3D) (x = 1–6). This pattern suggests that the cluster contains six Ag sites, which can sequentially detach under ionization conditions. Additionally, the atomic ratio of Ag to Ti in the cluster, as determined by ICP analysis, is 2.63, which is in excellent agreement with the theoretical value of 2.66 (Table S2).


image file: d4sc07186g-f3.tif
Fig. 3 (A) The overall structure of Ti12Ag4.5; (B) space-filling model of the structure of Ti12Ag4.5, highlighting the exposed Ag atoms on the surface; (C) bonding mode of six Ag1+ sites to the titanium oxide core; (D) positive-mode MALDI-TOF-MS of Ti12L6 and Ti12Ag4.5 in CH2Cl2 solution.

image file: d4sc07186g-f4.tif
Fig. 4 Time-dependent ESI-MS for the reaction of the Ti12L6 crystals and Ag2SO4 in CH3CN (A) or DMF (B) solution at 80 °C at 0 h, 1 h and 6 h.

From the above results, it is evident that the solvent choice significantly influences the reaction pathway of Ti12L6 with Ag(I) salts; CH3CN primarily leads to Ti12Ag4.5, whereas solvents like DMF result in different structures of Ti2Ag2-DMF and Ti2Ag4. The significant impact of seemingly minor variations in solvent choice on the synthesis outcomes is fascinating. To elucidate the formation mechanism of Ti12Ag4.5, its evolutionary process was monitored using ESI-MS.39 Time-resolved ESI-MS analysis of the reaction mixture, containing Ti12L6 crystals and Ag2SO4 in CH3CN, captured data at various intervals (Fig. 4A). Initially, ESI-MS detected two principal signal sets corresponding to the +2 and +3 charge states of {HxTi12O18−y(TC4A)6}, indicating that Ti12L6 maintained its integrity in the early stages of the reaction. After heating at 80 °C for one hour, new signals emerged, with a strong peak corresponding to {HTi2(TC4A)2}+, resulting from the fragmentation of Ti12L6. Additionally, subtle signals in the m/z range of 2000–2500, likely representing {Ti12AgxO20−z(TC4A)6}2+, were observed. By the six-hour mark, the spectrum displayed distinct signal sets aligned with the +2 and +3 charge states of the {HxTi12AgyO20−z(TC4A)6} species, indicating a complete transformation of Ti12L6 into Ti12Ag4.5.

Conversely, in a DMF environment, ESI-MS revealed a completely different reaction scenario (Fig. 4B). Initially, Ti12L6 also maintained its structure in DMF (Fig. S34). One hour post Ag(I) salt introduction, ESI-MS revealed the breakdown of the macrocycle, detecting {HxTi6Oz(TC4A)3}+ and {HxTi2Oz(TC4A)2}+ species. Notably, the emergence of {HxTi2AgyOz(TC4A)2}+ (x = 2–4) clarified the formation of Ti2Ag4 and Ti2Ag2-DMF clusters. By the six-hour mark, these signals intensified significantly, highlighting the progressive formation of these clusters in the DMF solution.

These findings indicate that the synthesis of Ti12Ag4.5 is not merely a simple process of Ag1+ adsorption onto the Ti12L6 carrier. Instead, it involves a complex sequence of fragmentation and subsequent reassembly (Fig. 5). Under solvothermal conditions, the Ti12L6 framework undergoes a fragmentation process, resulting in the formation of {Ti2O(TC4A)} units. A dynamic equilibrium forms between the larger {Ti12O12(TC4A)6} structure and these smaller {Ti2O(TC4A)} units. In the presence of CH3CN solvent, during the reassembly phase, these {Ti2O(TC4A)} fragments coordinate with Ag1+ ions, facilitating the formation of Ti12Ag4.5. Conversely, using DMF as the solvent leads to a transformation of the {Ti2O(TC4A)} units into more complex {Ti2O(TC4A)2} structures. These structures then complex with Ag1+ ions to form Ti2Ag2-DMF and Ti2Ag4, highlighting a distinct synthetic pathway that is significantly influenced by the choice of solvent.


image file: d4sc07186g-f5.tif
Fig. 5 Schematic diagram illustrating the solvent-mediated self-assembly process transitioning from Ti12L6 to Ti12Ag4.5.

Electrochemical CO2 reduction

Ag-based nanoclusters are optimal catalysts for electrocatalytic CO2 reduction reactions.40–42 Tuning the coordination environment of Ag sites within the clusters is a key strategy to boost their catalytic activity and tune the selectivity for CO production.43–51 From the structural comparison provided, it is clear that the distribution of Ag sites in the two Ti–Ag clusters, Ti12Ag4.5 and Ti2Ag2, is significantly different. For bimetallic Ti2Ag2, Ag1+ and Ti4+ sites are indistinguishably mixed within the core of the cluster. In contrast, Ti4+ sites of Ti12Ag4.5 are embedded in the Ti–oxo core, while Ag1+ sites are strategically distributed across the cluster surface. These structural differences between Ti12Ag4.5 and Ti2Ag2 provide a unique platform for investigating how variations in the Ag environments affect their catalytic activity.52,53

Powder X-ray diffraction (PXRD) analysis shows that the Ti12Ag4.5 crystal maintains its crystalline structure even after exposure to a highly alkaline 1 M KOH solution for 24 hours (Fig. S35). Given the resilience of Ti12Ag4.5, its eCO2RR activity was assessed in a three-compartment flow cell with 1 M KOH as the electrolyte. Linear sweep voltammetry (LSV) results indicated a significantly higher current density and more positive onset potential for both the Ti12Ag4.5 and Ti2Ag2 clusters in the CO2 flow electrolyzer compared to the N2-purged system, confirming their CO2 reduction capability, as illustrated in Fig. 6A. Gas chromatography detected only CO and H2 as products, with no other liquid products identified by 1H NMR spectroscopy (Fig. S39). Control experiments under N2-saturated conditions yielded no carbon-reduction products. Isotopic tracing experiments with 13CO2 through GC-MS confirmed the production of 13CO (m/z = 29) (Fig. S40). Additionally, the primary electrocatalytic product of the Ti12L6 catalyst was identified as H2, indicating that the Ag1+ components of these clusters predominantly drive the catalytic activity (Fig. S37).


image file: d4sc07186g-f6.tif
Fig. 6 (A) LSV of samples in N2 or CO2 saturated 1 M KOH solution; (B) FE values of Ti12Ag4.5 at different voltages; (C) FE values of Ti2Ag2 at different voltages; (D) CO partial current density (jCO); (E) It test and FECO values of Ti12Ag4.5 at −1.4 V vs. RHE in different time periods; (F) the ATR-SEIRAS results from 2500 to 1000 cm−1 of Ti2Ag2 and Ti12Ag4.5.

Fig. 6B and C illustrate the trends in product distribution across varying potentials for two distinct clusters. In a CO2-saturated environment, Ti12Ag4.5 consistently exhibits a high FECO production, maintaining over 82.0% across an extensive potential range from −0.4 V to −1.8 V vs. RHE, peaking at 92.7% at a higher potential of −1.4 V (Table S3). However, for Ti2Ag2, the predominant reaction between −0.6 V and −1.0 V is CO2 reduction, reaching its highest FECO of 85.1% at −0.8 V. Beyond this, the reaction is largely overtaken by H2 evolution in the potential range from −1.2 to −1.8 V, leading to a progressive decline in FECO as potential increases. At −1.8 V, the FECO of Ti2Ag2 plummets to merely 21.2%, whereas Ti12Ag4.5 still manages to maintain an FECO of 83.25%. Additionally, the CO partial current density (JCO) for both clusters was also analyzed (Fig. 6D). The JCO for Ti12Ag4.5 reached an impressive 130.1 mA cm−2 at −1.8 V, which is above 2.4 times greater than that of Ti2Ag2. This comparative analysis emphasizes the pronounced differences in catalytic efficiencies between two clusters, attributing Ti12Ag4.5's superior performance to its unique structural configuration and effective Ag1+ site utilization.

Electrocatalytic stability is a key indicator for evaluating the performance of electrocatalysts in the eCO2RR. We employed PXRD to analyze the reduction of Ag+ in Ti12Ag4.5 after reaction at different applied voltages (Fig. S44). PXRD analysis reveals that the catalyst remains stable when the applied voltage is below −1.8 V. However, when the voltage exceeds −1.8 V, signals corresponding to Ag nanoparticles appear, indicating the reduction of Ag+ to metallic Ag. To further evaluate the durability of the catalyst, we conducted a rigorous 11-hour chronoamperometric test at −1.4 V. During this process, the current density remained above 100 mA cm−2, and the FE for CO remained stable above 85.0% (Fig. 6E). Additionally, ESI-MS analysis of the catalysts after electrolysis showed a signal for the {HxTi12Ag6O18−z(TC4A)6}2+ species, confirming that the structural integrity of the Ti12Ag4.5 catalyst was maintained (Fig. S45). Further characterization by transmission electron microscopy, X-ray photoelectron spectroscopy, and infrared spectroscopy showed that the catalyst maintained its chemical composition and structural stability during the electrolysis process (Fig. S47–S49). Differential pulse voltammetry (DPV) measurements reveal that the electrochemical gap of Ti12Ag4.5 is 1.91 V (Fig. S50), which is greater than the 1.32 V observed for Ti2Ag2. This larger electrochemical gap further suggests enhanced stability of Ti12Ag4.5 during the electrocatalytic process. The stability of Ag+ in the Ti12Ag4.5 clusters can be attributed to the strong electronic and structural stabilization provided by the Ti–oxo support. This unique characteristic not only preserves the active sites but also prevents the competitive reduction of Ag+, thereby enhancing the suitability of these clusters for CO2 reduction.

The hypothesized mechanism for CO2 reduction to CO using Ag-based catalysts follows the pathway: CO2(g) → *COOH → *CO → CO(g).54–57 To validate this mechanism, we utilized in situ electrochemical attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS). This technique allows for real-time monitoring of the absorption of evanescent waves by substances on the catalyst surface, providing direct insight into the reaction dynamics. We set the potential range from −0.5 V to −1.3 V and monitored changes in absorption peaks with Ti12Ag4.5 and Ti2Ag2 as the electrocatalysts (Fig. 6F). The spectra display similarities, including a pronounced peak at 1225 cm−1, attributed to C–OH stretching in *COOH, which intensifies with increasing voltage. Another peak at 1711 cm−1, corresponding to C[double bond, length as m-dash]O stretching in *COOH, also increases in intensity from −0.5 V to −1.3 V, suggesting a rise in the surface coverage of *COOH species with increasing voltage.58,59 Additionally, a weak signal at 2127 cm−1, attributed to the Ag–*CO vibration mode, indicates the presence of *CO adsorbed on the catalyst surface, especially at lower potentials where the *CO band intensity shows a slight increase. Notably, the intermediate characteristic peak of Ti2Ag2 becomes distinctly observable only from an electrode potential of approximately −1.1 V, while the characteristic peaks of Ti12Ag4.5 are already evident at −0.5 V. This observation indicates that Ti12Ag4.5 demonstrates a higher reaction activity and stronger catalytic proficiency for the eCO2RR, aligning with experimental findings.

Density Functional Theory (DFT) calculations were conducted to elucidate the reactivity of two specifically designed clusters, Ti12Ag4.5 and Ti2Ag2. These models were optimized to mirror their actual crystal structures, simplified by replacing the tert-butyl groups on the TC4A with H for faster computational convergence. The results of these optimizations are depicted in the Gibbs free energy diagrams for both the eCO2RR and the HER, as presented in Fig. 7A. DFT calculations specifically focused on the energetics of each step, revealing that the formation of the *COOH intermediate is the rate-determining step in the CO2RR process. A critical finding from our study is the calculated Gibbs free energy for the formation of *COOH on the Ag3 site in Ti12Ag4.5, which was notably low at 0.55 eV. This value contrasts with the corresponding energy of 0.88 eV for the same process within Ti2Ag2. This substantial difference highlights that the exposed Ag sites on Ti12Ag4.5 are much more energetically favorable for catalyzing the conversion of CO2 to CO compared to those on bimetallic Ti2Ag2 (Fig. 7B). Additionally, the Gibbs free energies for hydrogen adsorption (*H) were calculated, showing high values of 2.46 eV for both clusters, indicating that neither Ti2Ag2 nor Ti12Ag4.5 is favorable for the formation of H2. This finding is important as it underscores the higher selectivity of both clusters for the CO2RR-to-CO pathway compared to the HER.


image file: d4sc07186g-f7.tif
Fig. 7 (A) Free energy diagrams for the eCO2RR and HER on Ti12Ag4.5 and Ti2Ag2; (B) schematic diagram of the eCO2RR process on Ti12Ag4.5.

By comparing the Gibbs free energy diagrams of the eCO2RR and HER, the exceptional catalytic efficiency of Ti12Ag4.5 is clearly linked to its unique Ag coordination. This coordination is supported by a titanium–oxo core, with Ag sites evenly distributed across the cluster surface. Specifically, the Ag active sites in Ti12Ag4.5 exhibit a d-band center (εd) that is closer to the Fermi level at −3.04 eV. As a comparison, the bimetallic cluster Ti2Ag2 exhibits a lower εd value of −3.47 eV, likely attributed to the absence of the unique support effects found in Ti12Ag4.5. Furthermore, the projected density of states (PDOS) for Ti12Ag4.5 shows higher and narrower peaks at energies near the Fermi level, suggesting a higher and more localized density of electronic states (Fig. S53). This enhanced electronic configuration facilitates stronger interactions with adsorbate *COOH molecules, improving the activation and subsequent transformation of CO2. These electronic characteristics are critical for reactions requiring complex electron interactions, thus positioning Ti12Ag4.5 as a more effective catalyst for CO2 reduction.

Conclusions

In summary, this study has designed and synthesized a new type of crystalline oxide cluster-stabilized single-atom material that can more precisely emulate the structural and reactive characteristics of oxide-supported single-atom catalysts at the molecular level. We have discovered a solvent-mediated assembly process that transforms the calixarene-stabilized titanium oxide cluster, Ti12L6, into the oxide-supported cluster, Ti12Ag4.5. This transformed cluster prominently features six accessible single Ag1+ sites on its surface. The distinctive structure of Ti12Ag4.5 has demonstrated superior catalytic performance in the eCO2RR, outperforming the traditional bimetallic cluster Ti2Ag2 in terms of reactivity and CO selectivity. In situ ATR-SEIRAS and DFT calculations were employed to investigate the catalytic mechanism, complemented by a structural comparison between Ti12Ag4.5 and Ti2Ag2 to clarify the factors underlying their different performances. These insights emphasize the importance of controlling the distribution of active metal sites on the surface of the cluster. This study promotes the intentional design of interactions between active surface sites and their cluster support, potentially establishing a new paradigm in single-atom catalyst design by significantly enhancing catalytic activities.

Data availability

The data that support the findings of this study are available in the main text and the ESI.

Author contributions

C. L. supervised the project and conceived the idea. R. X. M., L. C. Z and L. K. W carried out synthesis, characterization and catalytic experiments of clusters. Q. T. and L. P. L performed the calculations for this article. Y.-L. S. was responsible for the single crystal testing. C. L. wrote the manuscript. All authors discussed the experimental results.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work was supported by the Natural Science Foundation of Hunan Province (2023JJ30650), the Central South University Innovation-Driven Research Programme (2023CXQD061), and the National Natural Science Foundation of China (21901256 and 21805170). We thank Professor Mingzhao Chen and Dr Ning Wang from Guangzhou University for the mass spectrometry measurements.

Notes and references

  1. L. Liu and A. Corma, Chem. Rev., 2018, 118, 4981–5079 Search PubMed.
  2. Y. Guo, S. Mei, K. Yuan, D. J. Wang, H. C. Liu, C. H. Yan and Y. W. Zhang, ACS Catal., 2018, 8, 6203–6215 CrossRef CAS.
  3. M. Murdoch, G. I. N. Waterhouse, M. A. Nadeem, J. B. Metson, M. A. Keane, R. F. Howe, J. Llorca and H. Idriss, Nat. Chem., 2011, 3, 489–492 CrossRef CAS PubMed.
  4. A. vonWeber and S. L. Anderson, Acc. Chem. Res., 2016, 49, 2632–2639 CrossRef CAS.
  5. B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li and T. Zhang, Nat. Chem., 2011, 3, 634–641 CrossRef CAS.
  6. S. Yang, J. Kim, Y. J. Tak, A. Soon and H. Lee, Angew. Chem., Int. Ed., 2016, 55, 2058–2062 CrossRef CAS PubMed.
  7. J. Zhang, X. Wu, W. C. Cheong, W. Chen, R. Lin, J. Li, L. Zheng, W. Yan, L. Gu, C. Chen, Q. Peng, D. Wang and Y. D. Li, Nat. Commun., 2018, 9, 1002 CrossRef.
  8. C. M. Zhao, X. Y. Dai, T. Yao, W. X. Chen, X. Q. Wang, J. Wang, J. Yang, S. Q. Wei, Y. Wu and Y. D. Li, J. Am. Chem. Soc., 2017, 139, 8078–8081 CrossRef CAS PubMed.
  9. J. D. Yi, X. P Gao, H. Zhou, W. Chen and Y. Wu, Angew. Chem., Int. Ed., 2022, e202212329 CAS.
  10. R. Lang, X. Du, Y. Huang, X. Jiang, Q. Zhang, Y. Guo, K. Liu, B. Qiao, A. Wang and T. Zhang, Chem. Rev., 2020, 120, 11986–12043 CrossRef CAS.
  11. X. F. Yang, A. Wang, B. Qiao, J. Li, J. Liu and T. Zhang, Acc. Chem. Res., 2013, 46, 1740–1748 CrossRef CAS PubMed.
  12. Z. Huang, X. Gu, Q. Cao, P. Hu, J. Hao, J. Li and X. Tang, Angew. Chem., Int. Ed., 2012, 51, 4198–4203 CrossRef CAS PubMed.
  13. C. T. Campbell, Acc. Chem. Res., 2013, 46, 1712–1719 CrossRef CAS.
  14. R. Jin, C. Zeng, M. Zhou and Y. Chen, Chem. Rev., 2016, 116, 10346–10413 CrossRef CAS.
  15. P. Liu, Y. Zhao, R. Qin, L. Gu, P. Zhang, G. Fu and N. Zheng, Sci. Bull., 2018, 63, 675–682 Search PubMed.
  16. W. H. Fang, L. Zhang and J. Zhang, Chem. Soc. Rev., 2018, 47, 404–421 Search PubMed.
  17. P. Coppens, Y. Chen and E. Trzop, Chem. Rev., 2014, 114, 9645–9661 CrossRef CAS PubMed.
  18. L. Rozes and C. Sanchez, Chem. Soc. Rev., 2011, 40, 1006–1030 RSC.
  19. Y. Lv, J. Cheng, A. Steiner, L. Gan and D. S. Wright, Angew. Chem., Int. Ed., 2014, 53, 1934–1938 CrossRef CAS PubMed.
  20. C. Zhao, Y. Z. Han, S. Dai, X. Chen, J. Yan, W. Zhang, H. Su, S. Lin, Z. Tang, B. K. Teo and N. Zheng, Angew. Chem., Int. Ed., 2017, 56, 16252–16256 Search PubMed.
  21. G. Zhang, C. Liu, D. L. Long, L. Cronin, C. H. Tung and Y. Wang, J. Am. Chem. Soc., 2016, 138, 11097–11100 CrossRef CAS PubMed.
  22. W. H. Fang, L. Zhang and J. Zhang, J. Am. Chem. Soc., 2016, 138, 7480–7483 CrossRef CAS PubMed.
  23. M. Y. Gao, F. Wang, Z. G. Gu, D. X. Zhang, L. Zhang and J. Zhang, J. Am. Chem. Soc., 2016, 138, 2556–2559 Search PubMed.
  24. J. L. Hou, N. H. Huang, D. Acharya, Y. X. Liu, J. Y. Zhu, J. X. Teng, Z. Wang, K. Q. Qu, X. X. Zhang and D. Sun, Chem. Sci., 2024, 15, 2655–2664 RSC.
  25. M. Zhang, M. Lu, M.-Y. Yang, J.-P. Liao, Y.-F. Liu, H.-J. Yan, J.-N. Chang, T.-Y. Yu, S.-L. Li and L. Y. Qian, eScience, 2023, 3, 100116 CrossRef.
  26. Y. J. Liu, P. Shao, M. Y. Gao, W. H. Fan and J. Zhang, Inorg. Chem., 2020, 59, 11442–11448 CrossRef CAS PubMed.
  27. X. Fan, F. Yuan, D. Li, S. Chen, Z. Cheng, Z. Zhang, S. Xiang, S. Q. Zang, J. Zhang and L. Zhang, Angew. Chem., Int. Ed., 2021, 60, 12949–12954 CrossRef CAS.
  28. M. Y. Gao, K. Wang, Y. Y. Sun, D. J. Li, B. Q. Song, Y. H. Andaloussi, M. J. Zaworotko, J. Zhang and L. Zhang, J. Am. Chem. Soc., 2020, 142, 12784–12790 CrossRef CAS PubMed.
  29. N. Li, J.-M. Lin, R.-H. Li, J.-W. Shi, L.-Z. Dong, J. Liu, J. He and Y. Q. Lan, J. Am. Chem. Soc., 2023, 145, 16098–16108 CrossRef CAS PubMed.
  30. E. M Han, R. X Meng, Y. Q Tian, J. Yan, K. Y Liu and C. Liu, Chem. Commun., 2023, 59, 11097–11100 RSC.
  31. S. Chen, Z. N. Chen, W. H. Fang, W. Zhuang, L. Zhang and J. Zhang, Angew. Chem., Int. Ed., 2019, 58, 10932–10935 CrossRef CAS PubMed.
  32. D. T. Geng, X. Han, Y. F. Bi, Y. C. Qin, Q. Li, L. L. Huang, K. Zhou, L. J. Song and Z. P. Zheng, Chem. Sci., 2018, 9, 8535–8541 RSC.
  33. S. T. Wang, X. H. Gao, X. X. Hang, X. F. Zhu, H. T. Han, W. P. Liao and W. Chen, J. Am. Chem. Soc., 2016, 138, 16236–16239 CrossRef CAS PubMed.
  34. Y. Q. Tian, W. L. Mu, L. L. Wu, X. Y. Yi, J. Yan and C. Liu, Chem. Sci., 2023, 14, 10212–10218 Search PubMed.
  35. Z. Wang, L. Li, L. Feng, Z. Y. Gao, C. H. Tung, L. S. Zheng and D. Sun, Angew. Chem., Int. Ed., 2022, 61, e202200823 Search PubMed.
  36. Z. Wang, H. F. Su, Y. W. Gong, Q. P. Qu, Y. F. Bi, C. H. Tung, D. Sun and L. S. Zheng, Nat. Commun., 2020, 11, 308 CrossRef CAS PubMed.
  37. Z. Wang, F. Alkan, C. M. Aikens, M. Kurmoo, Z. Y. Zhang, K. P. Song, C. H. Tung and D. Sun, Angew. Chem., Int. Ed., 2022, 61, e202206742 Search PubMed.
  38. Z. J. Guan, F. Hu, S. F. Yuan, Z. A. Nan, Y. M. Lin and Q. M. Wang, Chem. Sci., 2019, 10, 3360–3365 Search PubMed.
  39. X.-M. Luo, S. Huang, P. Luo, K. Ma, Z.-Y. Wang, X.-Y. Dong and S.-Q. Zang, Chem. Sci., 2022, 13, 11110–11118 RSC.
  40. L. J. Li, Y. T. Luo, Y. Q. Tian, P. Wang, X. Y. Yi, J. Yan, Y. Pei and C. Liu, Inorg. Chem., 2023, 62, 14377–14384 CrossRef CAS.
  41. Z. Liu, J. Chen, B. Li, D.-e. Jiang, L. Wang, Q. Yao and J. Xie, J. Am. Chem. Soc., 2024, 146(17), 11773–11781 CrossRef CAS PubMed.
  42. S. L. Zhuang, D. Chen, L. W. Liao, Y. Zhao, N. Xia, W. Zhang, C. Wang, J. Yang and Z. Wu, Angew. Chem., Int. Ed., 2020, 59, 3073–3077 CrossRef CAS PubMed.
  43. J. Y. Xu, L. Xiong, X. Cai, S. S. Tang, A. C. Tang, X. Liu, Y. Pei and Y. Zhu, Chem. Sci., 2022, 13, 2778–2782 RSC.
  44. X. Liu, E. D. Wang, M. Zhou, Y. Wan, Y. K. Zhang, H. Q. Liu, Y. Zhao, J. Li, Y. Gao and Y. Zhu, Angew. Chem., Int. Ed., 2022, 61, e202207685 CrossRef CAS PubMed.
  45. S. Li, A. V. Nagarajan, D. R. Alfonso, M. K. Sun, D. R. Kauffman, G. Mpourmpakis and R. Jin, Angew. Chem., Int. Ed., 2021, 60, 6351–6356 CrossRef CAS PubMed.
  46. S. Zhao, N. Austin, M. Li, Y. Song, S. D. House, S. Bernhard, J. C. Yang, G. Mpourmpakis and R. Jin, ACS Catal., 2018, 8, 4996–5001 CrossRef CAS.
  47. L. Tang, Y. Luo, X. Ma, B. Wang, M. Ding, R. Wang, P. Wang, Y. Pei and S. Wang, Angew. Chem., Int. Ed., 2023, e202300553 CAS.
  48. L. J. Li, W. L. Mu, Y. Q. Tian, W. D. Yu, L. Y. Li, J. Yan and C. Liu, Chem. Sci., 2024, 15, 7643–7650 RSC.
  49. G. C. Deng, H. Yun, M. S. Bootharaju, F. Sun, K. Lee, X. L. Liu, S. Yoo, Q. Tang, Y. J. Hwang and T. Hyeon, J. Am. Chem. Soc., 2023, 145(50), 27407–27414 CrossRef CAS PubMed.
  50. L. B. Qin, F. Sun, X. S. Ma, G. Y. Ma, Y. Tang, L. Wang, Q. Tang, R. Jin, Z. Tang and Z. H. Tang, Angew. Chem., Int. Ed., 2021, 60, 26136–26141 CrossRef CAS PubMed.
  51. X. S. Ma, F. Sun, L. B. Qin, Y. G. Liu, X. W. Kang, L. K. Wang, D. E. Jiang, Q. Tang and Z. H. Tang, Chem. Sci., 2022, 13, 10149–10158 RSC.
  52. S. Yoo, S. Yoo, G. C. Deng, F. Sun, K. Lee, H. Jang, C. W. Lee, X. L. Liu, J. Jang, Q. Tang, Y. J. Hwang, T. Hyeon and M. S. Bootharaju, Adv. Mater., 2024, 36, 2313032 CrossRef CAS PubMed.
  53. X. L. Liu, T. Ki, G. C. Deng, S. Yoo, K. Lee, B.-H. Lee, T. Hyeon and M. S. Bootharaju, Nanoscale, 2024, 16, 12329–12344 RSC.
  54. G. Deng, J. Kim, M. S. Bootharaju, F. Sun, K. Lee, Q. Tang, Y. J. Hwang and T. Hyeon, J. Am. Chem. Soc., 2023, 145, 3401–3407 CrossRef CAS PubMed.
  55. J. Wang, F. Xu, Z. Y. Wang, S. Q. Zang and T. C. W. Mak, Angew. Chem., Int. Ed., 2022, 61, e202207492 CrossRef CAS PubMed.
  56. L. J. Liu, Z. Y. Wang, Z. Y. Wang, R. Wang, S. Q. Zang and T. C. Mak, Angew. Chem., Int. Ed., 2022, 61, e202205626 CrossRef CAS PubMed.
  57. Y. F. Lu, L. Z. Dong, J. Liu, R. X. Yang, J. J. Liu, Y. Zhang, L. Zhang, Y. R. Wang, S. L. Li and Y. Q. Lan, Angew. Chem., Int. Ed., 2021, 60, 26210–26217 CrossRef CAS PubMed.
  58. D. Wang, J. J. Mao, C. C. Zhang, J. W. Zhang, J. S. Li, Y. Zhang and Y. F. Zhu, eScience, 2023, 3, 100119 CrossRef.
  59. X. Wang, M. H. Yu and X. L. Feng, eScience, 2023, 3, 100141 CrossRef.

Footnotes

Electronic supplementary information (ESI) available: X-ray crystallographic file in CIF format, and full experimental and computational details. CCDC 2349621–2349625. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4sc07186g
These authors contributed equally.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.