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

Dual activation of CO2 and N2 facilitated by single Ta4+ clusters

Yifan Gaoab, Ran Chengab, Klavs Hansenc and Zhixun Luo*ab
aState Key Laboratory for Structural Chemistry of Unstable and Stable Species, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China. E-mail: zxluo@iccas.ac.cn
bUniversity of Chinese Academy of Sciences, Beijing 100049, China
cCentre for Joint Quantum Studies, Department of Physics, School of Science, Tianjin University, 92 Weijin Road, Tianjin 300072, China

Received 7th September 2025 , Accepted 25th November 2025

First published on 28th November 2025


Abstract

Governing molecular activation and reaction selectivity is a fundamental strategy for advancing catalytic performance. Metal clusters show great promise for tailoring reactions due to their unique electronic structures, yet the microscopic mechanisms of how their active sites cooperate are still not fully elucidated. This work examines the reactivity of Tan+ clusters with CO2 and N2, demonstrating the dual activation of CO2 and N2 by individual Ta4+ clusters, resulting in N–O coupling. Through electron localisation function (ELF) and multi-centre bonding analyses, we elucidate how delocalised electrons and charge distribution of such metal clusters govern the reaction selectivity and stabilise intermediates along the reaction pathways. This work enhances our comprehension of cluster catalysis and offers a framework to design efficient catalysts for dual activation of CO2 and N2.


Introduction

In light of the challenges posed by global climate change and the continued energy requirement, the reduction of carbon dioxide (CO2)1–5 and nitrogen (N2),6–12 as pivotal subjects within green chemistry, necessitates the urgent advancement of efficient and sustainable catalytic technologies. Conventional catalytic methods, exemplified by the Haber–Bosch process,13 while prevalent in industry, face limitations in long-term viability due to their substantial energy consumption and environmental pollution concerns. Up to now, the experimental pursuit of highly efficient and low-energy catalytic systems has emerged as a hot topic of research.14–19 Meanwhile, the general mechanisms have been well established for N2 reduction (e.g., direct dissociative mechanism vs. associative distal/alternating pathway)20 and CO2 reduction (e.g., direct CO2 adsorption and activation vs. proton-coupled electron transfer, PCET).21,22 Also, insights into the metal identity in diverse catalysts, electronic and geometric structures, metal–support interactions,23,24 key intermediates, reaction path selectivity, etc. have been extensively studied.

Recent investigations indicate increasing interest in the dual activation of CO2 and N2.25–30 However, ambiguities persist concerning the active locations and catalytic processes. Remarkable research on innovative nanocatalysts and single-atom catalysts has shown the formation of C–N or N–O coupling products during the concurrent activation of CO2 and N2, contingent upon the properties of the catalyst's active sites and the evolving structures during the reaction.31,32 While the importance of N–C coupling is well known, the potential for direct N–O coupling interactions between CO2 and N2 may provoke new understanding of the fundamental mechanisms underlying NOx production in atmospheric haze.33 To comprehensively elucidate the catalytic mechanisms and establish a basis for the rational design of high-performance catalysts for dual activation of CO2 and N2, transition metal clusters are anticipated to function as optimal model systems.34,35 Catalysts made of atomically precise clusters feature well-defined structures, tuneable compositions, specific coordination environments, and adjustable multi-site synergy, enabling a direct link between the electronic properties of individual atoms and nanoparticles.36–40 This shows promise for achieving precise control and high performance in both CO2 reduction and N2 activation.41–50

This study investigates the reaction dynamics of tantalum (Ta) clusters with CO2 and N2. Joint experimental and theoretical analyses demonstrate that both localised and delocalised electrons (e.g., 3c–2e multicentre bonds) in the Ta clusters are associated with the activation of CO2 and N2, dictating variable C[double bond, length as m-dash]O and N[triple bond, length as m-dash]N dissociation pathways. CO2 selectively adsorbs at electron-rich Ta–Ta sites, undergoing C[double bond, length as m-dash]O cleavage via twisted deformation; N2 dissociates with low energy barriers on the resulting Ta4O1,2+ intermediates. Despite weaker passivation than the nitride Ta4N2+, the intermediates of cluster oxide Ta4O1,2+ preserve active sites for N–O coupling to form stable Ta4N2O2+. By revealing the structure–activity relationships between electronic structures and reaction pathways, our findings lay the groundwork for optimising energy-efficient and selective CO2 and N2 conversions. This study clarifies how charge distribution and electron localisation or delocalisation control reaction selectivity and intermediate stability, thereby enabling more sustainable chemical processes.

Results and discussion

Fig. 1a depicts the mass spectrum of nascent Tan+ clusters (n = 3–11), produced by a magnetron sputtering (MagS) cluster source in tandem with a multi-ion laminar flow tube reactor and a triple quadrupole mass spectrometer (MIFT-TQMS).51 Fig. 1b illustrates the recorded mass spectrum resulting from the CO2 reaction, showing a series of products including TanO+, TanO2+, TanO3+, and TanO4+. The observation of these products indicates that the Tan+ clusters readily react with CO2 and are favourably transformed into metal oxides, along with CO removal. Fig. 1c illustrates the subsequent reactions by further introducing N2 into the flow tube reactor. The TanO0–4+ clusters then reacted with N2 to yield a series of TanNxOy+ products. The inset shows the products from a typical Ta4+ cluster, such as Ta4NO+, Ta4N2O+, and Ta4N2O2+. The observation of these TanNxOy+ products suggests dual activation of N2 and CO2, likely involving N–O coupling processes. Notably, this experiment did not exhibit a significant etching effect, likely due to the efficient cooling of the reaction products caused by the presence of enough He gas in the laminar flow tube.
image file: d5sc06899a-f1.tif
Fig. 1 (a) Mass spectrum of the nascent Tan+ clusters. The stars above the mass peaks indicate oxygen contamination. (b) Mass spectra of the Tan+ clusters after reactions with CO2 (1% in He) and (c) then N2 (1% in He) in the same flow tube reactor (∼0.5 Torr), with a controllable flow rate of ∼2 sccm for each.

The reactions of Tan+ with CO2 and N2 to form TanNxOy+ products were also validated using 15N isotope-labelled nitrogen gas (Fig. 2) and 18O-labelled carbon dioxide (Fig. S1, SI). When the Tan+ clusters react with a mixed reactant containing isotope-labelled 15N2 and normal CO2, they form TanNxOy+ products similar to those generated by adding CO2 and N2 sequentially. Upon the introduction of a minor flow of the 15N2/CO2 mixed gas, a range of nitrogen oxide products, such as TanNO+, TanN2O+, and TanN2O2+, emerge in the mass spectra. When introducing a larger dose of the reactant gas, bigger clusters were etched and depleted, resulting in the persistence of the relatively stable Ta4+ cluster. Concurrently, the relative intensity of nitrogen oxide species associated with the clusters rose, and the diversity of nitrogen- and oxygen-containing products likewise increased. The degree of reactivity tends to decrease as the cluster size grows. This is because larger clusters have greater heat capacity and more flexible structures to dissipate energy, whereas smaller clusters have limited energy dispersal and reduced collision cooling, thus leading to shorter lifetimes of intermediates. Meanwhile, thermal dissociation threshold values are usually small for the larger clusters, allowing for incidental fragmentation.52,53 Apart from the fruitful products of the Ta4+ cluster under different conditions (Fig. S2–S7, SI), distinct nitride and nitrogen oxide products of relatively weaker mass abundances were also detected on the other Tan+ clusters. These Tan+ clusters exhibited notably elevated reaction rates (larger than 10−9 cm3 s−1 per molecule) with the mixed reactant of 15N2/CO2 under ambient conditions (Fig. S8, SI).


image file: d5sc06899a-f2.tif
Fig. 2 (a–c) Mass spectra of the Tan+ cluster before and after the reactions with a mixed reactant of N2 and CO2 (both 1% in He) at different flow rates (ca. 2 sccm and 5 sccm), introduced from the inlet of the MagS source to leverage plasma assistance. An isotopically pure reagent of 15N2 was used to provide unambiguous mass assignments. The stars on the peaks indicate oxygen contamination.

Notably, N2 readily dissociates in reacting with the bare Tan+ clusters (Fig. S2, SI), demonstrating high activity of such clusters for N2 activation. This observation aligns well with previous theoretical and experimental studies, which have reported similar reactivity for small Ta clusters.54–57 Also, the production of TanNxOy+ products is reproduced with different N2/CO2 ratios or different size distributions of the nascent Tan+ clusters (Fig. S3, SI). Our results demonstrate that the end-on coordinated N2 bends toward a neighbouring Ta atom. This forms an intermediate where N2 bridges two Ta atoms along a tetrahedral Ta4 edge, followed by N–N bond dissociation to yield two µ3-N moieties on the neighbouring Ta3 facets. A comparison of thermodynamic data reveals that the energy released during the adsorption of CO2 on the Ta4+ cluster (ΔH = −0.31 eV) and the subsequent CO removal to form Ta4O+H = −2.23 eV) are typically smaller than that for N2 adsorption (ΔH = −0.78 eV) and the formation of N–N dissociative Ta4N2+H = −4.77 eV). However, enhanced stability of the N-capped Ta4N2+ diminishes the likelihood of subsequent activation of CO2 molecules (ΔH = −1.34 eV), indicating that the µ3-N exerts an effective passivation effect on the Ta4+ cluster. This result rebuts the traditional idea that N2 must be added first because its stronger bond (N[triple bond, length as m-dash]N) makes it harder to activate than CO2 (Table 1).

Table 1 The energy differences of Ta4+ in reacting with CO2 and N2, calculated quantum chemically at the tpsstpss/def2tzvp level of theory
Ta4+ + CO2 → Ta4CO2+, ΔH = −0.31 eV Ta4+ + N2 → Ta4·N2+, ΔH = −0.78 eV
Ta4+ + CO2 → Ta4O+ + CO, ΔH = −2.23 eV Ta4+ + N2 → Ta4N2+, ΔH = −4.77 eV
Ta4O+ + CO2 → Ta4O·CO2+, ΔH = −0.35 eV Ta4N2+ + CO2 → Ta4N2+·CO2, ΔH = −2.16 eV
Ta4O+ + CO2 → Ta4O2+ + CO, ΔH = −1.97 eV Ta4N2+ + CO2 →Ta4N2O+ + CO, ΔH = −1.34 eV
Ta4O+ + N2 → Ta4O·N2+, ΔH = −1.38 eV Ta4O2+ + N2 → Ta4O2·N2+, ΔH = −0.87 eV
Ta4O+ + N2 → Ta4ON2+, ΔH = −4.23 eV Ta4O2+ + N2 → Ta4O2N2+, ΔH = −4.35 eV


In comparison, when the Ta4+ cluster first reacts with CO2 to form a Ta4O+ intermediate (ΔH = −2.23 eV), the subsequent N2 adsorption energy is markedly exothermic (ΔH = −1.38 eV). This indicates that nitrogen adheres strongly to the cluster oxide Ta4O+ and is less prone to desorption compared to the nascent Ta4+ cluster, which is beneficial for stable adsorption and subsequent N2 activation. The N[triple bond, length as m-dash]N dissociation energy on the Ta4O+ intermediate (ΔH = −4.23 eV) is comparable with that on the pure Ta4+ cluster, establishing an energetic basis for the amalgamation of nitrogen and oxygen atoms within clusters. These energy changes promote the coupling of nitrogen and oxygen atoms, leading to the formation of N–O coupling products, of which the geometric and electronic structures are given in the SI (Fig. S9–S16).

Fig. 3 presents the energy diagrams comparing the potential reaction pathways for a Ta4+ cluster with N2 and two CO2 molecules. When Ta4+ initially reacts with N2, it forms a stable N2-dissociated Ta4N2+ intermediate (Fig. 3A), which makes the subsequent CO2 activation dynamically less favourable, with an energy barrier of 5.02 eV for N–O coupling (TS5). In comparison, the reaction of CO2 first (Fig. 3B) suffers from a small single-step energy barrier for the first CO2 dissociation (0.67 eV, TS1). Nonetheless, the formation of Ta4O+ along with a CO removal will result in significant energy gain (−2.23 eV, IM2). The second CO2 adsorption and C[double bond, length as m-dash]O bond dissociation event is more exothermic, forming a Ta4O2+ product with two isomeric configurations (IM4). The energy gains are −4.06 eV for the ortho-site bridged isomer (Route 1) and −4.20 eV for the para-site bridged isomer (Route 2). The large energy gain for Ta4+ to react with two CO2 molecules facilitates the subsequent N2 activation. Notably, the energies for N2 dissociation on Ta4O+ (Fig. S17, SI) and Ta4O2+ (Fig. 3B) are lower than that on pure clusters. The significant decrease in energy promotes the subsequent N2-adsorption and N[triple bond, length as m-dash]N dissociation, and enables N–O coupling at a relatively smaller single-step energy barrier (TS6). The other reaction pathways for different sequences of CO2 and N2 adsorbed on the Ta4+ cluster all lead to similar results with a product of Ta4(NO)1,2+ and the pathways for N–N dissociation and NO release ultimately are thermodynamically favourable even on a graphene support (Fig. S18–S24, SI). Additionally, the selectivity of the two reaction pathways (i.e., Routes 1 and 2) was evaluated by referring to RRKM theory.58–61 The results demonstrate a pronounced kinetic preference for Route 1, as evidenced by its significantly higher rate constant of 8.85 × 1012 s−1—roughly 122 times larger than the value calculated for Route 2 (c.a. 7.23 × 1010 s−1).


image file: d5sc06899a-f3.tif
Fig. 3 (A and B) Energy reaction diagram of a Ta4+ cluster reacting with an N2 and two CO2 molecules at different sequences, ultimately yielding Ta4NO+ and a released NO molecule. Insets depict the LUMOs and HOMOs of the intermediates involved in the reaction process.

The modified electronic configurations profoundly influence the cluster's reactivity. Nitrogen's stronger electron affinity disrupts the bonding network, enhancing passivation and inhibiting further reactions. Fig. 4 presents the electron localisation function (ELF)32,33 analysis for all the intermediate states along Route 1 (for details see Fig. S25–S34, SI), revealing the electronic coupling between small molecules and Ta4+ clusters throughout the whole reaction process. ELF analysis reveals uniformly distributed charges across all remaining Ta–Ta bonds at this stage. The distinctive electronic configuration of Ta4+ facilitates charge-driven interactions with CO2. Initially, CO2 interacts with Ta4+ to form an adsorption intermediate (IM1). Following this, the coordination of CO2 causes distortion of the electron cloud in TS1 (µ2-O), facilitating the cleavage of a C–O bond to form Ta4O+ (IM2). The incorporation of oxygen modifies the cluster's electronic configuration, resulting in greater charge localisation within the Ta–Ta bonds adjacent to the Ta–O–Ta units (specifically, Ta1–Ta3). The enhanced bonding network promotes further CO2 adsorption, leading to the formation of the Ta4O2+ intermediate (IM4). In addition to the considerable energy advantage of CO removal in the formation of Ta4O2+, large charge accumulation at the metal sites of the Ta4O2+ intermediate facilitates subsequent N2 adsorption and activation. The increased charge density facilitates subsequent N2 adsorption across multiple Ta–Ta sites (Fig. S36, SI). Subsequent to N2 adsorption, charge polarisation facilitates N2 dissociation, wherein one nitrogen atom transitions to an oxygen site via a transition state to yield NO. The residual nitrogen is drawn to adjacent charges, resulting in further NO production and the transformation of the cluster into Ta4NO+. Likewise, the reaction pathway for Route 2 was delineated using ELF analysis (Fig. S35, SI), revealing heightened charge density between neighbouring Ta–Ta bonds (Ta2–Ta3) and the connection (Ta3–Ta4). This method emphasises the synergistic effects of cluster charge redistribution and orbital overlap interactions in the catalytic dual activation of CO2 and N2.


image file: d5sc06899a-f4.tif
Fig. 4 The 2D and 3D electron localization functional (ELF) analyses for the intermediates (IM) and transition states (TS) during the reaction process (Route 1 in Fig. 3). In 2D-ELF, active sites are highlighted with white circles, while changes in covalent bonding are indicated by grey circles. The numerical labels on the atoms correspond to the atomic numbering within the metal cluster. In 3D-ELF, the red arrows indicate active sites, with the atoms 1, 2, 3, and 4 being kept in the same position for all.

This interplay between heteroatom incorporation and electronic restructuring underscores the utility of adaptive natural density partitioning (AdNDP) analysis.62 Notably, the frontier orbitals of N2 and CO2 align well with those of the nascent Ta4+ and subsequent intermediates (IM2 and IM4), promoting the formation of coordination bonds once adsorbed. Previous studies have demonstrated the mechanisms governing such small cluster stability and structure–property relationships, including geometric structure, charge population, and surface Lewis acid/base sites, as well as electronic configuration including superatomic features, and the resulting energetics and dynamics within the surrounding environment.57 Here, the finding of unique cluster catalysis of Ta4+ is largely contributed by its identical Lewis acid sites and the reduced sizes.54–56 To fully elucidate the reaction mechanism, we analysed the bonding modes and their evolution in this small cluster reaction using AdNDP analysis, which offers a comprehensive approach to describe chemical bonding, encompassing both localised electron pairs and delocalised multi-centre bond interactions (Fig. S37–S43, SI). As illustrated in Fig. 5, the nascent Ta4+ cluster initially features five 2c–2e bonds and four 3c–2e bonds, along with a 2c–1e Ta–Ta bond due to the positive charge. The Ta4O+ and Ta4O2+ intermediates preserve the multi-centre bonds and electronic structure of Ta4+; however, the addition of oxygen weakens the 2c–1e bond occupancy and decreases the occupation numbers (ON) of delocalised 3c–2e bonds, indicating electron transfer to the oxygen atoms. In contrast, the Ta4N2+ and Ta4NO+ clusters exhibit a more substantial disruption of 3c–2e bonding, indicative of a more pronounced effect than in Ta4O1–4+ clusters. This emphasises nitrogen's stronger ability to withdraw electrons and its more significant influence on the electronic accommodation during the reaction process.


image file: d5sc06899a-f5.tif
Fig. 5 AdNDP bonding analysis of the Ta4+, Ta4O1–4+, Ta4N2+ and Ta4NO+ clusters. The occupation numbers (ON) are given below each pattern.

In principle, localised electrons, due to their concentrated electron density, tend to act as key active sites or reactive centres. Electron-rich regions, such as lone pairs and π-bonds, can initiate nucleophilic reactions; conversely, electron-deficient areas, like the positive end of a polar bond, can be targeted by electron-rich reagents. The well-defined bond energy of localised σ-bonds within a cluster, like Ta4+, allows for homolytic cleavage and radical formation. In contrast, delocalised electrons (or multicentre bonds) may function to stabilise the overall metal cluster structure, with a more even charge distribution throughout the cluster. A balance of local bonding and delocalised electrons allows for distinctive reaction pathways by stabilising crucial intermediates,63 thereby generating energetically favourable sites for further electrophilic or nucleophilic attacks, which is particularly beneficial when multiple reactant molecules are involved.

Experimental

Experimental details

The experiments were conducted using a homemade MIFT-TQMS instrument,64 in tandem with a customised MagS source. The MagS source was used to produce pure tantalum clusters. These clusters were then introduced into a flow tube reactor (with a diameter of 60 mm and a length of 1 meter) using high-purity helium gas (purity > 99.999%). Inside the reactor, the clusters reacted with specific reactant gases. The flow rate of the reactant gases was controlled using a gas flowmeter (Alicat) throughout the experiment. To ensure sufficient collisional interactions of the Tan+ clusters, the pressure within the MagS source was kept at approximately 2.6 Torr, while the flow tube was kept at about 0.6 Torr. This was achieved by using a laminar flow of helium buffer gas from the upstream MagS source and a downstream Roots pump with a pumping speed of 142 litres per second. After the reaction, the resulting products were transferred into a differentially pumped five-stage vacuum system, equipped with different-sized apertures and linear/conical octuple ion guides, respectively. Finally, the products were analysed using a custom quadrupole mass spectrometer.

Computational details

The initial structures of the tantalum clusters were designed by examining the experimental and theoretical studies that had been previously reported in the literature while also considering the variations in spin multiplicity and geometric structure. The Gaussian 16 program65 was employed to conduct the structural optimisation and energetics calculations at the TPSStpss/def2-TZVP level of theory.66 The energies were corrected by zero-point vibrations, and all the calculated low-lying isomers and intermediates were verified to ensure that there was no imaginary frequency. All transition states (TS) were confirmed to possess one imaginary frequency and correlated with the reaction pathway, which was also validated by intrinsic reaction coordinate (IRC) computations.67 The projected density of states (pDOS) and natural population analysis (NPA) studies were conducted utilizing the Multiwfn software.68 The VMD (Visual Molecular Dynamics) program was employed to create the figures.69

Conclusions

In conclusion, we have prepared well-resolved Tan+ clusters (n = 4–10) and studied their gas-phase reactions with N2 and CO2 using a customised flow tube reactor and mass spectrometer. Our study demonstrates a direct relationship between their atomic structure and their varying reactivity with the two inert gas molecules. Interestingly, the sequential activation of CO2 followed by N2 by the bare Ta4+ cluster is facilitated by an oxygen-induced electronic reconfiguration, resulting in N–O coupling. The incorporation of one or two oxygen atoms selectively creates new, highly reactive sites on the cluster surface, which are crucial for the subsequent N2 cleavage and coupling reactions. Conversely, the pre-formed Ta4N2+ cluster demonstrates reduced reactivity, which is rooted in its thermodynamically stable dual µ3-N structure. This structural motif passivates the active sites and releases more energy upon production of Ta4N2+ in contrast to the formation of a Ta4O2+ equivalent with µ2-O bridge bonds. Utilising ELF and AdNDP studies, we clarify how delocalised electrons and charge distribution within these metal clusters dictate reaction selectivity and stable intermediates across the reaction pathways. The selectivity of these Ta clusters is shaped by their electronic configuration, which is affected by their geometric structures. The essential role of multi-centre bonds in facilitating dual activation of N2 and CO2 presents a fundamental framework for rational design of efficient Ta-based catalysts, emphasising the balance between electronic accommodation and surface passivation.

Author contributions

The manuscript was written through the contributions of all authors.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: Fig. S1–S43, experimental and computational details, including mass spectrometry observations, rate constants, geometric and electronic structures, energetics, DOSs, frontier orbitals, reaction pathways, and ELF and AdNDP analyses. See DOI: https://doi.org/10.1039/d5sc06899a.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Grant No. 92261113), the CAS Project for Young Scientists in Basic Research (Grant No. YSBR-050), the CAS Key Research Program of Frontier Sciences (QYZDBSSW-SLH024), and Beijing Natural Science Foundation (Grant No. F251006).

Notes and references

  1. W. Zhou, K. Cheng, J. Kang, C. Zhou, V. Subramanian, Q. Zhang and Y. Wang, Chem. Soc. Rev., 2016, 48, 3193–3228 RSC.
  2. X. Kang, Q. Zhu, X. Sun, J. Hu, J. Zhang, Z. Liu and B. Han, Chem. Sci., 2016, 7, 266–273 RSC.
  3. J. Wu, Y. Huang, W. Ye and Y. Li, Adv. Sci., 2017, 4, 1700194 CrossRef PubMed.
  4. L. G. Dodson, M. C. Thompson and J. M. Weber, Annu. Rev. Phys. Chem., 2018, 69, 231–252 CrossRef CAS PubMed.
  5. L. Wang, W. Chen, D. Zhang, Y. Du, R. Amal, S. Qiao, J. Wu and Z. Yin, Chem. Soc. Rev., 2019, 48, 5310–5349 RSC.
  6. M. Jiang, M. Zhu, M. Wang, Y. He, X. Luo, C. Wu, L. Zhang and Z. Jin, ACS Nano, 2023, 17, 3209–3224 CrossRef CAS.
  7. D. V. Yandulov and R. R. Schrock, Science, 2003, 301, 76–78 CrossRef CAS PubMed.
  8. D. Bao, Q. Zhang, F.-L. Meng, H.-X. Zhong, M.-M. Shi, Y. Zhang, J.-M. Yan, Q. Jiang and X.-B. Zhang, Adv. Mater., 2017, 29, 1604799 CrossRef PubMed.
  9. J. Deng, J. A. Iñiguez and C. Liu, Joule, 2018, 2, 846–856 CrossRef CAS.
  10. M.-A. Legare, G. Belanger-Chabot, R. D. Dewhurst, E. Welz, I. Krummenacher, B. Engels and H. Braunschweig, Science, 2018, 359, 896–899 CrossRef CAS PubMed.
  11. C. Liu, Q. Li, C. Wu, J. Zhang, Y. Jin, D. R. MacFarlane and C. Sun, J. Am. Chem. Soc., 2019, 141, 2884–2888 CrossRef CAS.
  12. G. Qing, R. Ghazfar, S. T. Jackowski, F. Habibzadeh, M. M. Ashtiani, C.-P. Chen, M. R. Smith III and T. W. Hamann, Chem. Rev., 2020, 120, 5437–5516 CrossRef CAS PubMed.
  13. C. M. Goodwin, P. Lömker, D. Degerman, B. Davies, M. Shipilin, F. Garcia-Martinez, S. Koroidov, J. Katja Mathiesen, R. Rameshan, G. L. S. Rodrigues, C. Schlueter, P. Amann and A. Nilsson, Nature, 2024, 625, 282–286 CrossRef CAS PubMed.
  14. T. Wu, X. Zhu, Z. Xing, S. Mou, C. Li, Y. Qiao, Q. Liu, Y. Luo, X. Shi, Y. Zhang and X. Sun, Angew. Chem., Int. Ed., 2019, 58, 18449–18453 CrossRef CAS PubMed.
  15. M.-M. Shi, D. Bao, B.-R. Wulan, Y.-H. Li, Y.-F. Zhang, J.-M. Yan and Q. Jiang, Adv. Mater., 2017, 29, 1606550 CrossRef.
  16. H. Cheng, L.-X. Ding, G.-F. Chen, L. Zhang, J. Xue and H. Wang, Adv. Mater., 2018, 30, 1803694 CrossRef PubMed.
  17. Y. Guo, T. Wang, Q. Yang, X. Li, H. Li, Y. Wang, T. Jiao, Z. Huang, B. Dong, W. Zhang, J. Fan and C. Zhi, ACS Nano, 2020, 14, 9089–9097 CrossRef CAS PubMed.
  18. J. Gu, C.-S. Hsu, L. Bai, H. M. Chen and X. Hu, Science, 2019, 364, 1091–1094 CrossRef CAS PubMed.
  19. L.-H. Mou, Y. Li, G.-P. Wei, Z.-Y. Li, Q.-Y. Liu, H. Chen and S.-G. He, Chem. Sci., 2022, 13, 9366–9372 RSC.
  20. C. Cui, H. Zhang, R. Cheng, B. Huang and Z. Luo, ACS Catal., 2022, 12, 14964–14975 CrossRef CAS.
  21. G. Liu, P. Poths, X. Zhang, Z. Zhu, M. Marshall, M. Blankenhorn, A. N. Alexandrova and K. H. Bowen, J. Am. Chem. Soc., 2020, 142, 7930–7936 CrossRef CAS PubMed.
  22. X.-Y. He, Y.-Z. Liu, J.-J. Chen, X. Lan, X.-N. Li and S.-G. He, J. Phys. Chem. Lett., 2023, 14, 6948–6955 CrossRef CAS PubMed.
  23. M. Xu, M. Peng, H. Tang, W. Zhou, B. Qiao and D. Ma, J. Am. Chem. Soc., 2024, 146, 2290–2307 CrossRef CAS PubMed.
  24. H. Miura, K. Imoto, H. Nishio, A. Junkaew, Y. Tsunesada, Y. Fukuta, M. Ehara and T. Shishido, J. Am. Chem. Soc., 2024, 146, 27528–27541 CrossRef CAS PubMed.
  25. C. Chen, X. Zhu, X. Wen, Y. Zhou, L. Zhou, H. Li, L. Tao, Q. Li, S. Du, T. Liu, D. Yan, C. Xie, Y. Zou, Y. Wang, R. Chen, J. Huo, Y. Li, J. Cheng, H. Su, X. Zhao, W. Cheng, Q. Liu, H. Lin, J. Luo, J. Chen, M. Dong, K. Cheng, C. Li and S. Wang, Nat. Chem., 2020, 12, 717–724 CrossRef CAS.
  26. M. Yuan, J. Chen, Y. Xu, R. Liu, T. Zhao, J. Zhang, Z. Ren, Z. Liu, C. Streb, H. He, C. Yang, S. Zhang and G. Zhang, Energy Environ. Sci., 2021, 14, 6605–6615 RSC.
  27. M. Yuan, J. Chen, Y. Bai, Z. Liu, J. Zhang, T. Zhao, Q. Wang, S. Li, H. He and G. Zhang, Angew. Chem., Int. Ed., 2021, 60, 10910–10918 CrossRef CAS PubMed.
  28. M. Wang, L.-Y. Chu, Z.-Y. Li, A. M. Messinis, Y.-Q. Ding, L. Hu and J.-B. Ma, J. Phys. Chem. Lett., 2021, 12, 3490–3496 CrossRef CAS PubMed.
  29. P. Xing, S. Wei, Y. Zhang, X. Chen, L. Dai and Y. Wang, ACS Appl. Mater. Interfaces, 2023, 15, 22101–22111 CrossRef CAS PubMed.
  30. X. Song, C. Basheer, Y. Xia, J. Li, I. Abdulazeez, A. A. Al-Saadi, M. Mofidfar, M. A. Suliman and R. N. Zare, J. Am. Chem. Soc., 2023, 145, 25910–25916 CrossRef CAS PubMed.
  31. Y. Yao, Z. Sun, T. Li, Z. Zhao, Z. Li, X. Lu, Y. Wan, Y. Fan and Z. Chen, ACS Nano, 2025, 19, 18947–18975 CrossRef CAS PubMed.
  32. F. Hu, L. Yang, Y. Jiang, C. Duan, X. Wang, L. Zeng, X. Lv, D. Duan, Q. Liu, T. Kong, J. Jiang, R. Long and Y. Xiong, Angew. Chem., Int. Ed., 2021, 60, 26122–26127 CrossRef CAS PubMed.
  33. Q. Yi, C. Cui, D. Ma and Z. Luo, Inorg. Chem., 2025, 64, 4082–4089 CrossRef CAS.
  34. Z. Luo and A. Shehzad, ChemPhysChem, 2024, 25, e202300715 CrossRef CAS PubMed.
  35. A. E. Green, J. Justen, W. Schöllkopf, A. S. Gentleman, A. Fielicke and S. R. Mackenzie, Angew. Chem., Int. Ed., 2018, 57, 14822–14826 CrossRef CAS PubMed.
  36. G. Frenking, I. Fernández, N. Holzmann, S. Pan, I. Krossing and M. Zhou, JACS Au, 2021, 1, 623–645 CrossRef CAS PubMed.
  37. Y. Lu, W.-Z. Yang, X.-X. Ding, S.-Q. Nie, Z.-G. Jiang and C.-H. Zhan, Dalton Trans., 2023, 52, 13063–13067 RSC.
  38. C. Cui, Y. Jia, S. Lin, L. Geng and Z. Luo, Small, 2024, 20, 2404638 CrossRef CAS PubMed.
  39. L. Xiao, Q. Zheng, S. Luo, Y. Ying, R. Zhou, S. Zhou, X. Li, X. Ye, Z. Yu, Q. Xu, H. Liao and J. Xu, Sci. Adv., 2024, 10, eadn2707 CrossRef CAS PubMed.
  40. M. K. Mohanta and P. Jena, Nanoscale, 2025, 17, 8505–8514 RSC.
  41. B. Li, M. Chen, Q. Hu, J. Zhu, X. Yang, Z. Li, C. Hu, Y. Li, P. Ni and Y. Ding, Appl. Catal., B, 2024, 346, 123733 CrossRef CAS.
  42. Y. Du, C. Li, Y. Dai, H. Yin and M. Zhu, Nanoscale Horiz., 2024, 9, 1262–1278 RSC.
  43. W.-P. Chen, K.-P. Bai, M.-T. Lv, S. Ni, C. Huang, Q.-Y. Yang and Y.-Z. Zheng, Angew. Chem., Int. Ed., 2025, 64, e202424805 CrossRef CAS PubMed.
  44. H. Tao, C. Choi, L.-X. Ding, Z. Jiang, Z. Han, M. Jia, Q. Fan, Y. Gao, H. Wang, A. W. Robertson, S. Hong, Y. Jung, S. Liu and Z. Sun, Chem, 2019, 5, 204–214 CAS.
  45. J. K. Li, J. P. Dong, S. S. Liu, Y. Hua, X. L. Zhao, Z. Li, S. N. Zhao, S. Q. Zang and R. Wang, Angew. Chem., Int. Ed., 2024, 63, e202412144 CrossRef CAS PubMed.
  46. 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.
  47. L. Qin, F. Sun, X. Ma, G. Ma, Y. Tang, L. Wang, Q. Tang, R. Jin and Z. Tang, Angew. Chem., Int. Ed., 2021, 60, 26136–26141 CrossRef CAS PubMed.
  48. Y. Sun, Z. Luo and J. Qiu, Angew. Chem., Int. Ed., 2024, 63, e202406879 CrossRef CAS PubMed.
  49. Q. J. Wu, D. H. Si, P. P. Sun, Y. L. Dong, S. Zheng, Q. Chen, S. H. Ye, D. Sun, R. Cao and Y. B. Huang, Angew. Chem., Int. Ed., 2023, 62, e202306822 CrossRef CAS PubMed.
  50. S. Zhuang, D. Chen, L. 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.
  51. H. Zhang, H. Wu, Y. Jia, L. Geng, Z. Luo, H. Fu and J. Yao, Rev. Sci. Instrum., 2019, 90, 073101 CrossRef PubMed.
  52. B. Yoon, P. Koskinen, B. Huber, O. Kostko, B. von Issendorff, H. Hakkinen, M. Moseler and U. Landman, ChemPhysChem, 2007, 8, 157–161 CrossRef CAS PubMed.
  53. M. Yang, H. Wu, B. Huang and Z. Luo, J. Phys. Chem. A, 2019, 123, 6921–6926 CrossRef CAS PubMed.
  54. C. Geng, J. Li, T. Weiske and H. Schwarz, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 11680–11687 CrossRef CAS.
  55. D. V. Fries, M. P. Klein, A. Steiner, M. H. Prosenc and G. Niedner-Schatteburg, Phys. Chem. Chem. Phys., 2021, 23, 11345–11354 RSC.
  56. D. V. Fries, M. P. Klein, A. Strassner, M. E. Huber, M. Luczak, C. Wiehn and G. Niedner-Schatteburg, J. Chem. Phys., 2023, 159, 164303 CrossRef CAS PubMed.
  57. R. Cheng, Y. Gao, C. Cui and Z. Luo, J. Phys. Chem. Lett., 2025, 16, 454–459 CrossRef CAS PubMed.
  58. R. A. Marcus, J. Chem. Phys., 1952, 20, 359–364 CrossRef CAS.
  59. S. K. Gray, S. A. Rice and M. J. Davis, J. Phys. Chem., 1986, 90, 3470–3482 CrossRef CAS.
  60. J. Westergren, H. Grönbeck, S.-G. Kim and D. Tománek, J. Chem. Phys., 1997, 107, 3071–3079 CrossRef CAS.
  61. B. Huang, W. Gan, K. Hansen and Z. Luo, J. Phys. Chem. A, 2022, 126, 4801–4809 CrossRef CAS PubMed.
  62. T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS.
  63. S. Lin, D. Li, D. Zhang, L. Geng, Y. Jia, W. Wang, L. Cheng, S. N. Khanna and Z. Luo, Chem. Sci., 2025, 16, 11619–11625 RSC.
  64. B. Huang, H. Wu, M. Yang and Z. Luo, Rev. Sci. Instrum., 2022, 93, 113307 CrossRef CAS PubMed.
  65. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16 Rev. B.01, Gaussian, Inc., Wallingford, CT, 2016 Search PubMed.
  66. J. Wu, J. Hu, Q. Liu, Y. Tang, Y. Liu, W. Xiang, S. Sun and Z. Suo, J. Mol. Model., 2024, 30, 33 CrossRef CAS PubMed.
  67. C. Gonzalez and H. B. Schlegel, J. Chem. Phys., 1989, 90, 2154–2161 CrossRef CAS.
  68. T. Lu, J. Chem. Phys., 2024, 161, 082503 CrossRef CAS PubMed.
  69. W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graph., 1996, 14, 33–38 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2026
Click here to see how this site uses Cookies. View our privacy policy here.