Shiyan
Wang†
a,
Chaopeng
Liu†
a,
Weiyao
Hao
a,
Yanling
Zhuang
a,
Jianmei
Chen
a,
Xianjun
Zhu
a,
Longlu
Wang
*a,
Xianghong
Niu
b,
Jianjun
Mao
*c,
Dongwei
Ma
*d and
Qiang
Zhao
*a
aCollege of Electronic and Optical Engineering & College of Flexible Electronics (Future Technology), State Key Laboratory of Flexible Electronics, Nanjing University of Posts and Telecommunications, Nanjing 210023, China. E-mail: wanglonglu@njupt.edu.cn; iamqzhao@njupt.edu.cn
bCollege of Science, Nanjing University of Posts and Telecommunications, Nanjing 210023, China
cDepartment of Physics, Shanghai Normal University, Shanghai 200232, China. E-mail: jjmao@shnu.edu.cn
dSchool of Physics and Electronic Information, Huaibei Normal University, Huaibei 235000, China. E-mail: dwmachina@126.com
First published on 20th March 2025
The structural evolution of metal single-atoms and clusters has been recognized as the new frontier in catalytic reactions under operative conditions, playing a crucial role in key aspects of catalytic behavior, including activity, selectivity, stability, and atomic efficiency as well as precise tunability in heterogeneous catalysis. Accurately identifying the structural evolution of metal single-atoms and clusters during real reactions is essential for addressing fundamental issues such as active sites, metal–support interactions, deactivation mechanisms, and thereby guiding the design and fabrication of high-performance single-atom and cluster catalysts. However, how to evaluate the dynamic structural evolution of metal species during catalytic reactions is still lacking, hindering their industrial applications. In this review, we discuss the behaviors of dynamic structural evolution between metal single-atoms and clusters, explore the driving force and major factors, highlight the challenges and inherent limitations encountered, and present relevant future research trends. Overall, this review provides valuable insights that can inspire researchers to develop novel and efficient strategies for accurately identifying the structural transformations of metal single-atoms and clusters.
Despite extensive research on the design, synthesis, and application of SACs, their commercialization remains limited. Efforts to enhance activity and selectivity often result in compromised stability, a crucial factor impeding large-scale application.10,28–34 The instability of SACs arises from the high surface energy of individual metal atoms, which causes them to aggregate into more stable clusters. This leads to the structural evolution of metal SAs into clusters under operational conditions.35–37 For instance, under a negative potential, H adsorption drives dynamic structural evolution of Cu SAs agglomerations to clusters in Cu–N–C catalysts, enhancing C2H5OH and NH3 production during NO3− and CO2 reduction reaction (NO3RR and CO2RR).38–41 Under pyrolysis conditions, the rare-earth, transition, and noble metal clusters undergo dynamic structural evolution into SAs, which are then stabilized by neighboring N-dopants. This process results in the formation of thermodynamically stable M–N–C structures.42,43 Under different gas atmospheres, similar dynamic behavior is observed that Pt SAs stabilize in the O2 atmosphere and can further undergo dynamic structural evolution to Pt clusters in either H2 or a mixture of CO + O2 atmosphere on alumina during CO oxidation.44 Additionally, a notable structural transformation was observed in CeO2 and TiO2. Li's group developed a laser ablation method to control the structural evolution of Pt clusters into SAs on CeO2, leading to enhanced CO oxidation efficiency.45 Utilizing ab initio molecular dynamics (AIMD) simulations and a microkinetic model, Wang et al.46,47 explored the dynamic behavior at metal/oxide interface. They demonstrated that, under varying temperatures and oxygen partial pressure conditions, Au SAs detach from Au clusters to catalyze CO oxidation and subsequently reintegrate into the Au clusters after the reaction. These dynamic transformations do not deactivate the catalysts but create actual active sites, promoting efficient chemical reactions. Unfortunately, it remains an open question whether the original anchoring sites for metal SAs are stable on a substrate, and if not, What is the structure of the genuine active site for this demanding chemical process? Consequently, a comprehensive assessment of structural evolution between metal SAs and clusters is crucial, as it will be instructive and potentially accelerate the development of research in this field.
In this review, we delve into the dynamic structural transformations between metal SAs and clusters under reaction conditions and explore the driving factors behind these changes. Subsequently, we discuss recent advances in the aggregation of metal SAs into clusters and the reverse transformation from metal clusters into SAs. Finally, we summarize the dynamic structural transformation behaviors observed in catalytic processes, discuss the challenges and inherent limitations encountered, and present relevant future research trends. Overall, the catalytic activity is closely linked to the dynamic structural changes of metal species, as revealed through experimental and theoretical studies. Real-time and spatial monitoring of these changes is vital for uncovering the molecular-level mechanisms of catalysis, determining the structure–activity relationship, and identifying the underlying causes of catalyst deactivation.
![]() | ||
Fig. 1 (a) Schematic representation of structural evolution between metal SAs and clusters.48 (b) Strategies for the structural evolution of metal SAs and clusters under operating conditions of different catalytic reactions. |
![]() | ||
Fig. 2 (a) Schematic representation of the potential-induced structural evolution of Cu–N–C SACs during ORR and NO3RR.36 (b) The aggregation of Cu SAs into clusters is driven by the applied potential switching from 0.00 to −1.00 V vs. RHE, accompanied by NH3 production.38 (c) A novel pathway involving the evolution of Cu–(CO)x moieties driven by the synergistic adsorption of CO and H was proposed for dynamic Cu sintering in Cu–N–C catalysts under the CO2RR condition. (d) The leaching barriers of Cu at −1.0 VSHE were determined through constant-potential CI-NEB calculations. (e) The bond lengths of Cu–N in CuN4, CuN4H2, and CuN4H3 were monitored during a constant-potential AIMD simulation for 10 ps at U = −1.0 VSHE. (f) The optimized structures of CuN4H3 and Cu(CO)2N4H3. (g) The constant-potential Cu leaching barriers on Cu(CO)2N4H3 at U = −0.5, −1.0, and −1.5 VSHE. (h) Cu–Cu bond lengths in Cu(CO)2N4H3 were observed during an additional 10 ps of the AIMD simulation following the leaching of the Cu–(CO)2 moiety.39 (i) The mechanism underlying the dynamic reversible transformation between Cu SAs and Cu4 clusters is driven by potential-induced structural evolution.40 (j) Schematic diagram of electrochemical reconstruction pathway of Cu3P/Cu–N/C catalyst to Cux/Cu–N/C during the CO2RR.41 |
To confirm the existence of a genuine dynamic structural transformation behavior in Cu–N–C catalysis, constant-potential AIMD simulations have revealed that synergistic adsorption of H and CO intermediates can promote the sintering process, leading to the agglomeration of Cu SAs into clusters under CO2RR conditions (Fig. 2c).39 To better understand the kinetics of Cu sintering, the energy barriers for Cu migration from the initial N-coordinated site to a neighboring C site at −1.0 VSHE were calculated using potential-dependent climbing image nudged elastic band (CI-NEB) simulations. As shown in Fig. 2d, the leaching barriers for Cu migration from the N-coordinated site in CuN4 to an adjacent C site decreased notably upon the adsorption of 0, 1, 2, and 3H atoms, with values dropping to 3.66, 1.35, 0.18, and 0.23 eV, respectively. During constant-potential AIMD simulations at 300 K for 10 ps, the Cu–N bond lengths in CuN4, CuN4H2, and CuN4H3 were monitored. As depicted in Fig. 2e, the Cu–N bond length in CuN4 remained stable at approximately 1.9 Å, indicating the Cu atom's resistance to leaching. However, bond breaking in CuN4H2 and CuN4H3 occurred at 614 fs and 670 fs, respectively, once the bond length surpassed 2.95 Å. After bond breaking, the desorbed Cu atoms retain the ability to coordinate with adjacent H-adsorbed N atoms, thereby impeding their swift migration. Furthermore, they found that adding CO intermediate in the form of Cu–(CO)x moieties can significantly accelerate the spillover and clustering of the Cu SAs in Cu–N–C catalysts (Fig. 2f–h). This finding was further supported by theoretical calculations from Bai et al.,40 H adsorption is crucial in driving the reversible dynamic transformation between Cu SAs and Cu4 clusters by applying potential-driven structural evolution in Cu–N–C catalysis under the CO2RR conditions (Fig. 2i). Experimentally, Zang et al.41 demonstrated that the electrochemical structural evolution of Cu3P particles under increasing negative potential resulted in the gradual formation of Cu clusters through dispersion and assembly (Fig. 2j). These Cu clusters exhibited an ethanol faradaic efficiency of approximately 40% and a partial current density of around 350 mA cm−2 during CO2RR. Similar potential-driven transformations in coordination and electronic structures, such as the transition from M–N4 to coordination-unsaturated M–Nx, were also observed in Ni–N–C, Co–N–C, and other catalytic systems.58–60
Large periodic holes in g-C3N4 provide sufficient space, rendering it an ideal substrate for supporting catalysts.61–65 As a result, the incorporation of metal species can significantly alter the electronic structures, with the synergistic interaction between the metal species and g-C3N4 playing a pivotal role in improving the catalytic performance of g-C3N4.66–73 As mentioned above, Cu SAs often undergo leaching and aggregation processes into clusters during chemical reactions. Using operando attenuated total reflectance surface-enhanced infrared absorption spectroscopy, Zhang et al.74 conducted experiments in a specialized electrochemical cell to quantitatively track the electrochemical transformation of Cu SAs into clusters on a g-C3N4 substrate under CO2RR conditions, as illustrated in Fig. 3a. The reconstruction of Cu SAs is strongly influenced by the applied potential, with the rate of Cu clusters evolution at −1.2 V being roughly two orders of magnitude greater than at −0.6 V during the CO2RR (Fig. 3b). Furthermore, DFT calculations show that the coordination environment can be adjusted by modulating the interaction between the Cu SAs and the catalyst substrates.
![]() | ||
Fig. 3 (a) Schematic representation of the electrochemical reconstruction of Cu SAs to clusters on g-C3N4 substrate under CO2RR condition. (b) Evolution rates of Cu single sites are dependent on the applied potential.74 (c) Free energy diagram of original Cu SAs migrating to its adjacent equivalent position of Cu1/g-C3N4. (d) Schematic of the aggregation process of Cu clusters on the Cu1/g-C3N4 SACs. (e) Kinetic barriers of *CO → *CHO and *CO → *OCCO on Cu3/g-C3N4. (f) Activity origin of Cu species toward CO2 electroreduction to C2+ products.75 |
To evaluate the dynamic structural transformation behavior of Cu SAs to clusters on g-C3N4, constant-potential AIMD simulations with a slow-growth approach were performed. As shown in Fig. 3c, the Cu1 atom of Cu1/g-C3N4 can readily migrate from the original position to its adjacent equivalent position with a kinetic barrier of only 0.28 eV. Further, the AIMD simulations demonstrate that the Cu1 can be readily aggregated into Cu2, and then into Cu3 and Cu4 clusters by utilizing available free Cu1 species, as depicted in Fig. 3d. As we all know, the C–C coupling is crucial for C2 products, Zhang et al.75 investigated its reaction process. As illustrated in Fig. 3e, the kinetic barrier of *CO coupling significantly decreases to merely 0.41 eV when the number of atoms in the Cu clusters reaches three. Fig. 3f depicts the mechanism of activity origin of Cu species toward CO2RR to C2+ products: not one, not two, but at least three atoms can efficiently promote CO–CO coupling, as determined by kinetic analyses. Similarly, Bai et al.76 designed Cu4 clusters loaded on the natural pore structure of g-C3N4, featuring a Cu0–Cux atomic interface, for highly efficient and selective CO2RR into C2H5OH.
In addition to the two-dimensional materials of M–N–C and g-C3N4 mentioned above, the structural transformation of metal clusters formation process has also been studied on the surfaces of transition-metal (TM),77–79 metal oxide of ZnO2 (ref. 49) and γ-Al2O.44 Yang et al.77 explored the dynamic changes in composition, morphology, and structure under the CO2RR conditions using operando/in situ techniques. As illustrated in Fig. 4a, monodisperse Cu undergoes a structural transformation, where the surface oxide is decreased, followed by ligand desorption, resulting in aggregated Cu clusters. These changes correlate with the generating catalytically active sites that promote higher C2+ formation.
![]() | ||
Fig. 4 (a) Monodisperse Cu NPs structural transformation to the active metallic Cu nanograins by operando/in situ methods under the CO2RR condition.77 (b) Schematic diagram of the catalyst structures evolution of the cluster formation on a TM surface under realistic reaction conditions.78 (c) Snapshots from kinetic Monte Carlo simulations illustrating the evolution of a Cu(111) surface under exposure to 0.3-torr CO at room temperature, taken at intervals throughout a total simulation period of 105.0 min. (d) Schematic representation of the CO2 hydrogenation reaction on various Cu species, highlighting their impact on product selectivity.49 (e) Schematic illustration depicting the migration of copper species at varying Cu loadings within ZrO2 during CO2 hydrogenation reaction. (f) Dynamics structural evolution of Pt SAs into clusters on γ-Al2O3 during CO oxidation monitored by the operando XAS, DRIFTS monitoring, and STEM images.44 (g) XANES evolution white-line intensities and (h) EXAFS analysis of Pt–Pt CN during the calcination, reaction, reduction, and subsequent reaction sequence for 0.3 wt% Pt/γ-Al2O3 under COOX10 and COOX2 conditions, and for 1.0 wt% Pt/γ-Al2O3 under COOX10 conditions. |
The formation of active sites on the TM occurs through the surface atom's migration toward the clusters, driven by the reaction. As depicted in Fig. 4b, the adsorbate-induced cluster formation occurs in three steps: (i) step-edge atoms ejection, leading to the formation of adatoms on an adjacent terrace, (ii) diffusion of adatoms, and (iii) the adatoms aggregation to form clusters. Using DFT calculations, Xu et al.78 introduced the adatom formation energy as a descriptor for efficiently screening clustering systems. They further employed kinetic Monte Carlo simulations to investigate the CO-driven cluster formation on a Cu surface (Fig. 4c). In addition, Liu et al.79 developed a reliable Cu–C–O machine learning force field with ab initio accuracy, enabling an in-depth understanding of the reconstruction mechanism and distribution of active sites on Cu surfaces under a CO atmosphere through state-of-the-art deep potential molecular dynamics. Combining cluster distribution statistical analysis with microkinetic simulations, they established a strategy to quantitatively evaluate the overall turnover frequency of catalyst surfaces during dynamic catalytic processes. The results indicate that edge Cu atoms undergo rearrangement, ejection, diffusion, and aggregation under a CO atmosphere, leading to the dynamic formation of cluster active sites. These small clusters in dynamic equilibrium are considered the origin of the high catalytic activity of Cu-based catalysts in low-temperature water–gas shift reactions.
EXAFS characterization revealed that during the CO2 hydrogenation reaction, Cu atoms migrate to the surface of Cu/ZrO2 catalysts, highlighting the impact of metal size on catalytic behavior.49Fig. 4d and e illustrate that varying Cu loading results in catalyst formations of different sizes, directly affecting product selectivity. At <2 wt%, the migrated Cu species did not accumulate but remained distributed in a monodispersed state, resulting in the exclusive production of CH3OH and a linear increase in CO2 conversion with increasing Cu loading. When Cu loading was between 4 and 8 wt%, the migrated Cu species partially aggregated into clusters, increasing CO2 conversion and CO selectivity, while CH3OH selectivity decreased. Above 8 wt%, some of the Cu species aggregated to form large Cu particles that cannot activate CO2. Though the size of active metal species remains unchanged before and after the reaction, migration, and diffusion of metal species may occur during CO2RR process, thereby revealing the actual active sites.
Gas molecular atmosphere significantly influences the dynamic structural transformation from single metal atoms into clusters. Based on the operando XAS, DRIFTS monitoring, and STEM images, Dessal et al.44 reported (Fig. 4f) that Pt SAs stabilize in the O2 atmosphere and can further undergo dynamic structural transformations into Pt clusters in either H2 or a mixture of CO + O2 atmosphere on γ-Al2O3 during CO oxidation. Fig. 4g and h present a quantitative comparison of changes in white line intensities and coordination numbers (CN) of Pt–Pt. Transitioning from 0.3Pt COOX10 to 1Pt COOX10 or 0.3Pt COOX2 led to a decrease in white line intensity, indicating that increased Pt loading and reduced O2 concentration promote Pt reduction. During the initial reaction cooling stage, COOX2 conditions caused a more substantial rise in CN, reaching a peak CN of 5, compared to a CN of 3–4 under COOX10 conditions. Consequently, the more reductive COOX2 environment facilitates the dynamic transformation of Pt SAs into clusters.
Significantly, to monitor the structural evolution of Cu SAs on the CeO2 catalyst during C–N coupling with varying applied potential, Wei et al.80 performed in situ XAS experiments. The Cu R-space EXAFS spectrum reveals that the coordination of the first shell of Cu transition from Cu–O to Cu–Cu bonding, i.e., Cu2+ is gradually reduced to Cu+ and Cu0 at the reduction potential, indicating the formation of Cu clusters. Furthermore, when the applied potential shifts to an open-circuit potential, the Cu clusters undergo a dynamic and reversible transformation back into SAs.
![]() | ||
Fig. 5 (a) Achieve controlled conversion of Pt NPs to SAs on CeO2 using a laser ablation strategy for enhanced CO oxidation efficiency. (b) The chemical potential of Pt at varying oxygen chemical potentials is calculated based on the configuration of the active center.45 (c) Schematic illustration of the anticipated reaction mechanism for CO oxidation at the Au SAs. (d) Snapshots of dynamics structural evolution by the MD simulation for CeO2-supported Au50 cluster to SAs.46 (e) Distribution function P(rcm) of Au atoms relative to the center of mass of the Au-nanoparticle.47 (f) The Mulliken charge of the isolated Au atom plotted against the simulation time after equilibration. (g) Schematic representation of the catalyst structures evolution of the Cu SAs formation process on Gd-doped CeO2 surface.82 (h) Gibbs free energy change calculated for multiple CO adsorptions on Cu10/Gd0.25Ce0.75O2 under CO oxidation conditions. (i) The top and side views of the most stable CO-saturated adsorption configuration on Cu10/Gd0.219Ce0.781O2. |
Fig. 5c presents a schematic illustration of the dynamic behavior at the metal/oxide interface, outlining the following catalytic steps: (i) CO adsorption initiates the formation of isolated Au–CO, which interacts with lattice oxygen ions, producing a CO2 molecule; (ii) Au SAs then reassemble into clusters; (iii) a subsequent CO adsorption generates another isolated Au–CO complex; (iv) O2 adsorbs at the Ov site, and the second CO molecule reacts with this adsorbed O2, resulting in an additional CO2 molecule. Based on AIMD simulations, they also performed the dynamics and structural transformations of Au clusters to SAs on CeO2. In the absence of CO adsorption, a 14 ps AIMD simulation confirmed that creating an isolated Au atom is energetically unfavorable. In contrast, in the presence of CO, the dynamics structural evolution of Au clusters required only 2 ps to SAs, where CO was initially located at a low-coordination interfacial Au site (Fig. 5d).46 In a separate study, the function P(rcm) was employed to quantify the probability distribution of an Au atom positioned at a distance of rcm from the center of mass of Au20 clusters. The peak observed around ∼7 Å in P(rcm) enables the clear identification of separated Au atoms, as shown in Fig. 5e. Upon detachment of the Au–CO unit from the cluster, the charge state of Au shifts from negative within the cluster to positive in the SAs (Fig. 5f). This transition suggests that, similar to CeO2, the active species is best represented as an Oad–Au+–CO unit.47
Recently, Liu et al.82 investigated the transformation of metal Cu clusters to SAs (Fig. 5g), focusing on how Gd-doping influences the electronic interactions between the metal and support. They explored the combined effect of Gd-doped CeO2 and CO adsorption in the formation of single-atom sites. Through DFT calculations (Fig. 5h), they analyzed the Gibbs free energy changes associated with multiple CO molecules adsorbing onto Gd-doped CeO2-supported Cu clusters under CO oxidation conditions. Their ab initio thermodynamic analysis revealed that the adsorption of a sixth CO molecule causes the Cu–CO species to detach from the cluster (Fig. 5i), leading to the formation of active Cu single-atom sites. These findings underscore the importance of dynamic active site formation under reaction conditions and their significant role in CeO2 and TiO2.
On the M–N–C catalysts, Li et al.42 observed a direct conversion of noble metal clusters into SAs at temperatures above 900 °C in an inert atmosphere. As shown in Fig. 6a, high-resolution high-angle annular dark field scanning transmission electron microscopy confirmed the dynamic structural shift from metal clusters to SAs. Statistical data on particle diameters and quantities at various temperatures are displayed in Fig. 6b and c. The average diameter of Pd clusters gradually increased from room temperature up to 900 °C, in contrast, the number of Pd clusters steadily decreased (Fig. 6b), indicating a concurrent process of sintering and atomization during the transformation from metal clusters to SAs. At 1000 °C, a Pd cluster with an initial 6.5 nm diameter decreased to 4.0 nm at 136 s, further to 1.9 nm at 150 s, and completely disappeared by 162 s (Fig. 6c). Additionally, DFT calculations were used to simulate the conversion of clusters to SAs. As shown in Fig. 6d, a more thermodynamically stable Pd single atom formation at the N4 defect from the Pd10 clusters decomposition requires overcoming a kinetic barrier of 1.47 eV, accompanied by a significant exothermicity of 3.96 eV. In a separate study, heating (PtNi)n/Zn-ZIF from room temperature to 600–900 °C in an inert atmosphere resulted in notable changes in the elemental distribution of Ni, Pt, and Zn. They also proposed a novel conversion mechanism for multiphase metal catalysts, involving atomic displacement between SAs and nano-alloys clusters. This process consists of the release and enrichment of Zn atoms from the CN substrate into nano-alloy clusters, and the migration of Ni atoms from nano-alloy clusters into the CN substrate, while Pt atoms remain distributed on the nano-alloy clusters throughout the entire transformation.48
![]() | ||
Fig. 6 (a) Schematic representation of the structural transformation from Pd NPs to SAs, depicted through high-resolution HAADF-STEM images. Average diameter and number of particles (b) vs. temperatures (c) vs. pyrolyzing time at 1000 °C. (d) The kinetic barriers, and the corresponding initial and final configurations of the Pd atom from the Pd10 cluster to the Pd–N4 defect.42 (e) Scheme of the energy barrier for the structural transformation of metal NPs to SAs. (f) The thermally driven atomization process transforming NPs to SAs is facilitated by adjacent N-dopants. (g) Schematic of the balance between achieving atomization and high loading while avoiding unwanted sintering during temperature-driven atomization. (h) The proposed strategy of high N-doping-assisted atomization enables the transformation of MxSy clusters into highly dispersed single metal atoms. (i) The ultrahigh metal loadings achieved for 23 different ultrahigh-density M–N–C SACs. (j) The kinetic barriers, and the corresponding initial and final configurations for the dynamic Ni atom dissociation from bulk NiS.43 |
The preparation of SAs by heat-driven direct transformation of metal clusters has been widely studied in above mentioned works. In theory, the structural transformation must overcome a kinetic barrier (Fig. 6e), thereby necessitating high temperatures to provide the requisite energy. Furthermore, methods for converting metal clusters to SAs are still challenged by low metal loading, primarily due to limited stabilization for mobile metal atoms. Introducing high metal content often leads to reaggregation from elevated surface free energy (Fig. 6f). High temperatures can intensify competitive atomization and sintering reactions, resulting in a tradeoff between achieving high metal atomic loading and mitigating sintering (Fig. 6g). Lu et al.43 present a high N-doping-assisted atomization strategy that directly transforms MxSy clusters into ultra-high density (UHD) SACs. As illustrated in Fig. 6h, they developed N-doped carbon supports with substantial nitrogen content to selectively control the migration pathway of metal atoms, favoring atomization over sintering. This approach effectively bypasses the usual trade-off between high loading and aggregation, thereby enhancing M–N–C loading. Among the 23 UHDM–N–C SACs, 17 metals exhibit loadings exceeding 20%, with 6 metals (Sn, Ce, Er, Dy, Pt, and Lu) showing loadings greater than 30% (Fig. 6i). CI-NEB calculations reveal that releasing a Ni atom requires a kinetic barrier of 1.06 eV (Fig. 6j). Theoretical results suggest that increasing the temperature facilitates the transformation of MxSy clusters into metal SAs, which aligns with the experimental observation of the thermal-driven decomposition of NiS clusters during the preparation of UHDNi–N–C SACs.
Zhang et al.83 utilized Co/Zn-ZIF as a precursor for pyrolysis, reporting the first observation of the evolution of Co SAs alongside the formation of a porous N-doped carbon support. Their study revealed that Co species underwent a dynamic cycle of agglomeration, dispersion, reagglomeration, and sublimation. In situ TEM and XAFS characterizations confirmed the stability of Co SAs on the porous CNx carriers, which demonstrated an exceptional catalytic performance for the selective oxidation of ethylbenzene into acetophenone.
Notably, the above working preparation process requires additional energy to break strong metal–metal bonds and inhibit agglomeration to maintain atomic-level dispersion, which makes it difficult to control the purity and uniformity of SAs samples. Very recently, Wang et al.84 demonstrated a novel, efficient, and simple strategy of electrochemically knocking-down Zn clusters into SAs (Fig. 7a): metal organic framework precursor derived from defects rich in Zn clusters carbon material as a starting point, through the controlled multi-cycle electrochemical lithiation/delithiation process, makes Zn clusters on carbon substrates to SAs. Using synchrotron radiation-based X-ray absorption near edge spectroscopy (XANES), they explore the coordination structure of Zn atoms to build a model that accurately represents real experimental conditions. The Zn K-edge XANES analysis depicted in Fig. 7b indicates that, compared with Zn SA–carbon, the position of the absorption edge for the Zn SA/cluster-carbon slightly shifts to large photon energy, revealing that the presence of a mixed oxidation state of Zn ranging between 0 and +2 originates from the Zn clusters and SAs. Subsequently, they used EXAFS spectra to quantitatively determine the coordination of the Zn atoms. As shown in Fig. 7c, apart from the primary Zn–N peak, the second most prominent peak in Zn SA/cluster-carbon is the Zn–Zn peak, which indicates the coexistence of Zn SAs and clusters.
![]() | ||
Fig. 7 (a) Schematic illustration of a novel, simple electrochemical knock-down approach for inducing structural evolution from Zn clusters into SAs.84 (b) Zn K-edge XANES spectra and (c) FT-EXAFS spectra of Zn SA/cluster-carbon and Zn SA–carbon, along with ZnO and ZnPc standard samples. (d) The calculated adsorption energy of Li+ on all possible adsorption sites. (e) The conversion pathway of metal Zn clusters into individual SAs during the alloying/dealloying process. |
To further validate the observed conversion of Zn clusters to SAs, they constructed a model incorporating defects, Zn–N4 coordination, and Zn clusters for DFT calculations. Adsorption energies of Li+ at different sites are described in Fig. 7d, Li+ would preferentially adsorb on the defects and Zn–N4 sites compared with that of Zn cluster sites, the active sites in the model are initially saturated with Li+ to facilitate a more detailed analysis of the specific interactions between Li+ and Zn clusters. As shown in Fig. 7e, the strong adsorption energy of Li+ indicates that Li+ can easily combine with Zn clusters, facilitating an alloying reaction. In addition, upon the interaction of the fifth Li+ with Zn clusters, the Zn atom appears to escape significantly, indicating that the Zn–Zn bond can be destroyed after enough Zn–Li bonds are established. These liberated Zn atoms are efficiently captured by numerous Zn–N4 defects, leading to the creation of new Zn monatomic sites. This significant finding marks the behavior of metal cluster dynamics, paving the way for creating innovative materials through the direct transformation of metal clusters into SAs.
An efficient strategy for producing high-purity and high-loading metal SAs involves starting with a well-defined substrate and predepositing metal atoms using conventional techniques, without the need for precise control. As illustrated in Fig. 8a–c, Liu et al.85 discovered that under an applied voltage, there is a sufficiently wide electrochemical potential window in which all metal species, except for the tightly bound SAs, are oxidized and leached out. For this method to be effective, the substrate must exhibit a strong enough binding to the specific SAs, ensuring that the chemical potential of the metal in the SAs is lower than in the metal clusters. Provided this condition is met, their theoretical results suggest that the strategy can successfully produce high-loading and high-purity SAs, even when the substrate surface contains clusters of varying sizes.
![]() | ||
Fig. 8 Schematic illustrations depicting (a) pre-deposited Pt on the substrate in the form of clusters and SAs, (b) SAs remaining on the substrate following selective leaching within the electrochemical potential window, and (c) typical trends of chemical potentials (∂G/∂ni) with particle size under three typical conditions. Threshold leaching potentials Un for Ptn (n = 20–1) support in (d) NC and (e) Fe2O3.85 |
The threshold potential (Un) required for leaching each metal atom was subsequently calculated. As shown in Fig. 8d, at U = 0.67 V, all 18 Pt atoms forming the Pt clusters at single vacancies leach into the aqueous solution as Pt2+, while two Pt SAs remain at double vacancies coordinated by four nitrogen atoms. Leaching of these two Pt SAs requires U > 1.80 V on NC. Similarly, Fig. 8e illustrates a chemical potential window of 0.23 V (between leaching potentials of 0.82 and 1.05 V) that facilitates the simple preparation of Pt/Fe2O3 SACs. The findings indicate that, apart from the strongly bound SAs, all forms of Pt undergo electrochemical leaching and dissolution from the NC and Fe2O3 surfaces into the electrolyte, providing an efficient strategy for preparing high-purity and high-loading Pt SACs without the need for precise control of Pt chemical potential. Additionally, the strategy's versatility is demonstrated by modeling the synthesis of Pd and Ni SACs anchored on NC and Fe2O3.
Dynamic structural evolution of metal SAs and clusters presents a significant challenge in identifying the catalytically active species. Ex situ characterizations may fail to capture true active sites, as metal SAs can aggregate into clusters or reaction-induced clusters may convert into isolated atoms by the reaction's end. Therefore, operando characterizations under realistic working conditions are essential to prevent inaccurate or misleading conclusions.
Researchers have made significant efforts to identify the true active sites in catalysts, but there remains disagreement about the contributions of metal clusters and single atomic sites to catalyst activity. Recently, an effective approach has been to combine metal sites with distinct structures onto a catalyst substrate, where these varied sites work in coordination to facilitate the overall reaction process through a division-of-labor strategy. The site with the highest intrinsic activity functions as the primary reaction site, while another metal species serves as a dopant to modulate the electronic structure of the active site. The synergistic effect between the composite catalysts consisting of metal clusters and SAs can tune the electronic structure of the active site, thereby enhancing its inherent activity. By incorporating two sites optimized for different catalytic reactions, bifunctional catalysis can be exploited to improve the efficiency of multi-step catalytic reactions.86–114 This underscores the importance of synergistic catalysis between metal clusters and SAs in optimizing catalytic efficiency. However, to date, no studies have addressed the dynamic structural evolution of this synergistic effect. A comprehensive theory to explain or predict the behavior of metal catalysts with varying particle sizes under different reaction conditions remains unexplored, which will be the focus of future research.
However, there are inherent limitations for real-time simulation of the catalytic reactions under operative conditions: (i) the elementary steps may occur at different locations of the catalyst, and accurately identifying the structure and dynamic evolution of the active site of the electrocatalyst during the catalytic reaction is a necessary condition for exploring the sources of the activity of the catalyst and understanding the mechanism of catalytic reaction. However, monitoring the real-time evolution of catalyst structural complexity is a significant challenge. (ii) The dynamic change of the structure can affect the kinetics of catalytic reactions, leading to different pathways and products. Therefore, developing a theoretical approach that integrates structural dynamics with reaction kinetics is crucial for simulating catalytic reactions. (iii) Establishing a reliable structure–activity relationship from the onset of the reaction to catalyst deactivation presents a significant challenge. In principle, this relationship is determined by comparing two different and clearly defined structures under the same reaction conditions. However, this task becomes extremely difficult due to the dynamical evolution of the catalyst structure and the complexity of its changes with the reaction conditions.
In summary, this review highlights the dynamic evolution of metal SAs and clusters in various catalytic reactions, showcasing the complexity and diversity of active site structures in catalysis. It emphasizes the significance of dynamic structural changes of metal species under operating conditions and identifies the challenges current theoretical methods face in simulating these changes. To gain a comprehensive understanding of reaction mechanisms and design efficient catalysts with optimized active site structures, it is crucial to focus on the dynamic evolution of the active center. This requires integrating catalyst structural dynamics with reaction kinetics through the development of novel theoretical frameworks and advanced in situ/operando characterization techniques.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |