DOI:
10.1039/D5MH00042D
(Review Article)
Mater. Horiz., 2025,
12, 3286-3300
Recent advances in metal single-atom catalysts for ammonia electrosynthesis
Received
9th January 2025
, Accepted 10th February 2025
First published on 12th February 2025
Abstract
Electrochemical ammonia synthesis is a promising alternative to the Haber–Bosch process, offering significant potential for sustainable agricultural production and the development of portable, carbon-free energy carriers. The development of electrocatalytic systems is currently dependent on the exploration of electrocatalysts with high activity, selectivity, and stability. Metal single-atom catalysts (SACs) have become a new attractive frontier for ammonia electrosynthesis, owing to their maximized atom utilization, unsaturated atom coordination, and tunable electronic structure. In this review, we focused on different metal sites inside the single-atom catalysts and summarized recent advances in SACs for ammonia electrosynthesis. The properties of small nitrogenous substances (including N2, NO, NO2−, and NO3−) are summarized. In addition, the SACs for different catalytic systems are reviewed, with a particular focus on the special and common grounds of metal atom sites. Finally, the perspectives and challenges of SACs for ammonia electrosynthesis are comprehensively discussed, aspiring to provide insights into the development of electrochemical ammonia synthesis.
Wider impact
The electrocatalytic conversion of nitrogenous substances (N2, NO, NO2−, and NO3−) to ammonia represents a promising alternative to the Haber–Bosch process. Metal single-atom materials with high activity and ammonia selectivity have attracted significant attention in materials science and electrochemical ammonia synthesis due to the exposure of sufficient active sites and the adjustment of the electronic structure and microenvironment. This review provides a concise comparison of the physicochemical properties of diverse nitrogenous substances (including N2, NO, NO2− and NO3−) and offers a synopsis of the impact of these properties on ammonia electrosynthesis. Additionally, it summarizes the frequently used metal active sites in metal single-atom catalysts and their roles in ammonia electrosynthesis. Finally, the review comprehensively discusses the problems faced by metal single-atom catalysts in ammonia electrosynthesis and future research directions. These advancements have profound implications for the design of electrocatalytic materials at the atomic scale and the study of fundamental mechanisms of catalytic conversions of nitrogenous substances. This review offers potential solutions for green ammonia electrosynthesis and other electrochemical energy conversions.
|
1. Introduction
Ammonia (NH3) plays a pivotal role in the global economy and has numerous applications in various fields, including fertilizer production. It also represents a future carbon–neutral hydrogen carrier with a high energy density.1,2 Currently, approximately 90% of the global NH3 production is produced through the century-old Haber–Bosch (H–B) process, which is regarded as the most significant innovation of the 20th century.3 However, the continued application of the existing energy- and capital-intensive H–B process NH3 production is limited by high temperatures, extremely high pressures, and excessive CO2 emissions.4 These limitations make the H–B process ineffective in meeting the future production demands of NH3. Therefore, exploring a sustainable alternative to the conventional H–B process has attracted considerable attention from researchers.
NH3 electrosynthesis has been identified as a potential solution to the dual challenges of energy crisis and environmental pollution (Fig. 1). This process will complement the H–B process as an additional means of NH3 production. In natural nitrogen cycles, abundant and readily available main forms of nitrogen (N2, NO, NO2−, and NO3−) are considered promising raw materials for NH3 electrosynthesis.5–8 Furthermore, NH3 electrosynthesis is performed under environmental conditions, and it uses clean and renewable electric energy, making it a green and sustainable method.9 However, in the electrocatalytic process, the multitude of H+ and e−-transfer reactions, characterized by slow kinetics, ultimately result in an ultra-low NH3 yield rate. Additionally, the competitive dynamics of the favorable two-electron hydrogen evolution reaction (HER) lead to low Faraday efficiency (FE) and selectivity in the NH3 production.10
 |
| Fig. 1 The transformation of nitrogenous substances into NH3. | |
Recently, the development of highly efficient electrocatalysts has been a crucial research focus for advancing “green NH3” production technology. Several metals (including Pt, Ru, Cu, and Ti) and their compounds exhibit catalytic activity.11–13 With further advancements in research, the design of metal active sites has become critical in optimizing the catalytic activity of electrocatalytic materials.14 Metal single-atom catalysts (SACs) have become an emerging research area, attracting much interest in the field of NH3 electrosynthesis. Especially, SACs (in which metal atoms are atomically dispersed on supports) provide great potential to achieve high NH3 selectivity and a high NH3 yield rate due to the exposure of sufficient active sites and the adjustment of the electronic structure and microenvironment of the overall catalytic materials.15 They often exhibit higher selectivity than bulk metals due to their effective suppression of the HER through the ensemble effect.16 Furthermore, the well-defined atomic-level structure of SACs is promising as suitable candidates for studying the fundamental mechanisms of catalytic conversions of nitrogenous substrates, serving as a model for monitoring structural evolution at the atomic scale and the catalytic behavior of the reaction center.17–21 The metal atoms in SACs usually have an unsaturated coordination environment, making their interactions with the reactants more flexible and efficient.22 The adsorption and activation of small nitrogenous substances can be enhanced through the engineering of the coordination environments of active single atoms. The energy barriers of different proton–electron transfer steps can be selectively adjusted, offering opportunities for further enhancing the selectivity of NH3 production under various conditions.23,24
In this review, we outline the importance of recent developments in these SACs for NH3 electrosynthesis. First, we compare the characteristics of different nitrogenous substances (N2, NO, NO2−, and NO3−), and summarize the advantages of the electroreduction process for NH3 electrosynthesis and the technical difficulties associated with this process. Additionally, we summarize the advantages of SACs in each electrocatalytic reaction. Second, we provide a comprehensive overview of the research progress made by introducing different metal sites in different electrocatalytic reactions, and summarize the possible effect of frequently-used metal active sites in each reaction. Finally, we summarize the problems faced by SACs in NH3 electrosynthesis and present future research directions. This review will provide insight into the development of advanced SACs for NH3 electrosynthesis and other electrochemical energy conversion fields.
2. Comparison of different nitrogenous substances for NH3 electrosynthesis
The efficient NH3 electrosynthesis is influenced by both thermodynamics and kinetics. Thus, the features of the nitrogenous reactants, reaction path, and reaction energy barrier need to be considered. The basic features of small nitrogenous substances are shown in Fig. 2. First, dinitrogen (N2) is derived from the natural atmosphere and can be used without any limitations, making it the optimal reactant.25 Nitrite ion (NO2−) and nitrate ion (NO3−) are major water pollutants.26 The excessive use of nitrogen-rich fertilizers, the improper discharge of untreated municipal sewage and industrial wastewater, leachate from landfill sites, and poorly constructed septic tanks and leach pits mainly contribute to nitrate accumulation in water bodies.27 Nitric oxide (NO) is recognized as a major component of atmospheric pollutants, and its primary anthropogenic source is fuel combustion in machinery, such as power plants, vehicles, aircraft, and internal combustion engines.28 During the combustion process, N2 in the fuel reacts with O2 at high temperatures to form NO, which is subsequently released and then reacts with O2 to form NO2. Furthermore, the industrial manufacturing and agricultural activities associated with nitrogen-containing products, such as the production of nitric acid, nitrates, and gunpowder, also result in the emission of a certain quantity of nitric oxide gases. The uncontrolled release of NO3−, NO2−, and NO into the natural environment can significantly disrupt the nitrogen cycle, resulting in adverse environmental impacts and posing a threat to human health.29 Therefore, converting these nitrogenous substances into NH3 exemplifies the “waste-to-valuables conversion” concept.
 |
| Fig. 2 Comparison of the basic properties of nitrogenous substances and standard reduction potentials (E0) for the major electrocatalytic NRR, NORR, NO2−RR, NO3−RR, and HER in an aqueous electrolyte under standard conditions. | |
Second, NO3− has the highest valence state of +5. Consequently, it entails a more intricate process involving multiple reactions, intermediates, and products that span a wide range of nitrogen valence states (from +5 to −3). Third, the highly inert N2 has the highest bond energy (941 kJ mol−1) from the N
N triple bond (the first bond dissociation energy is 410 kJ mol−1), followed by NO with a lower bond dissociation energy. Compared with NO, NO2− and NO3− have the lowest bond dissociation energy of 204 kJ mol−1.30,31 Therefore, the adsorption and activation of NO2− and NO3− on the catalyst surface may be more favorable because N–N separation is challenging. Finally, the water solubility of these nitrogenous substances has been compared. The results show that NO2− and NO3− are highly soluble, while N2 and NO are insoluble. Therefore, NO2− and NO3− are more favorable for achieving a high reaction rate and rapid NH3 electrosynthesis, and the mass transport limitation of NO and N2 in the electrolyte is a critical issue.
For NH3 electrosynthesis, the reactions and the corresponding standard reduction potentials (E0) for the electroreduction of different nitrogenous substances to NH3 are listed in Fig. 2. The E0 values of the NO reduction reaction (NORR), the NO2− reduction reaction (NO2−RR), and the NO3− reduction reaction (NO3−RR) are more positive than those of the N2 reduction reaction (NRR) [0.09 V vs. reversible hydrogen electrode (RHE)] and HER (0.00 V vs. RHE), indicating that they are thermodynamically easier to achieve than NRR.32 This indicates that the theoretical potential range of NORR is significantly larger than that of NRR in the absence of HER interference. Therefore, the electrocatalytic NORR, NO2−RR, and NO3−RR provide an excellent opportunity for NH3 electrosynthesis.
In the electrochemical process, the key stages of the NH3 electrosynthesis can be broadly divided into three categories: (1) the adsorption of nitrogenous substances, (2) the reaction of intermediates on the catalyst surface (multi-step reduction and oxidation), and (3) the desorption of NH3 molecules. Based on the density functional theory (DFT), the Gibbs free energy of each reaction coordinate state represents the spontaneity of each reaction step.33 Notably, because of the multistep proton–electron transfer process and multiple intermediates of these reactions, the reaction mechanisms are different. In the NRR process, the three main pathways are dissociative, associative distal, and associative altering pathways (Fig. 3a).34,35 The dissociative pathway usually occurs in the H–B process. N2 adsorption takes a side-on configuration on the active site, and then N2 activation entails the N
N bond breakage, generating an isolated active nitrogen (*N) atom on the active site. Since the N
N bond is very stable, this reaction pathway requires significant energy consumption, which is not suitable for sustainable development. Comparatively, more NH3 electrosynthesis reactions catalyzed by SACs have been reported to follow associative reaction mechanisms.36 In the associative mechanism, the N2 molecule is adsorbed on the active site, only breaking down in select hydrogenation stages. In addition, the hydrogenation process involves two steps: distal and altering pathways. The main bottlenecks in the electrochemical NRR arise from the thermodynamic constraints imposed by the intermediates of the reaction. Theoretically, when a sufficiently negative potential is applied to an electrode, the NRR is enabled at room temperature and atmospheric pressure. The equilibrium potential of the electrochemical NRR is comparable to that of the HER. Thus, H2 is the primary byproduct of the NRR process. Nevertheless, the NRR process involves multiple proton–electron transfer reactions, leading to the formation of a multitude of intermediates. Especially, the redox potential of the initial step (which forms the *NNH intermediate) is more negative, indicating the inherent difficulty of the initial active hydrogen (*H) addition. Furthermore, the addition of a second *H is more challenging than the addition of a third *H, leading to higher redox potentials for two- and four-electron reduction processes than for six-electron reduction processes.10 These significant negative potentials of the intermediates demonstrate the thermodynamic difficulty of NRR.
 |
| Fig. 3 Schematic depiction of the as-proposed possible mechanisms for NH3 electrosynthesis from (a) N2, and (b) NO, NO2− and NO3−. | |
The series of intermediate transformations involved in the electrochemical NORR, NO2−RR, and NO3−RR significantly differ from those observed in the NRR process. According to nitrogen chemistry, the electroreductions of NO, NO2− and NO3− are thermodynamically and dynamically more favorable than direct NRR-to-NH3 production. Many studies have shown that the pathways for the electrocatalytic reduction of NO2− and NO3− to NH3 and the reduction of NO to NH3 are identical (Fig. 3b),37 although their rate-limiting steps may be different. In the case of NO3−RR, the electroreduction of NO3− to NO2− is typically regarded as the limiting step in the NO3−RR process. Different from the NO3−RR process, NO2−RR directly commences with NO2− adsorption, indicating that NO2−RR can facilitate the thermodynamic production of NH3. The role of *H in the rapid reduction of intermediates is also significant. In addition to considering the effect of hydrogenation and the rate-determining step, the limited solubility of NO in water influences the performance. Furthermore, NO2−RR and NO3−RR depend on the key divergent center-adsorbed NO (*NO). The conversion of *NO to NH3 involves complex reaction pathways. The hydrogenation process involves the following steps: the solvated protons first adsorb onto catalysts to form adsorbed *H, followed by surface hydrogenation (Tafel-type pathway), or a NO molecule and intermediates are protonated directly (Heyrovsky-type pathway).38 The different reaction pathways determine the nature of the byproducts, which include N2O, N2, and NH2OH. NH2OH is reported as a common byproduct in metal single-atom catalysis systems, which is initially formed through the stepwise hydrogenation of *HNOH. It readily desorbs from the metal active sites when the adsorption is insufficient.39,40 The formation of N2 and N2O byproducts usually requires the dimerization of the absorbed NO intermediate species with the same orientation on the adjacent metal sites.23 Due to the high dispersion of metal sites, N2 and N2O byproducts are rarely found.
3. The characteristics of SACs in the NH3 electrosynthesis
To date, various single metal atoms have been used as the active centers for NH3 electrosynthesis. A periodic table shows the frequency of elements applied in SACs (Fig. 4). Most of these SACs consist of d-block transition metal sites. The underlying physical mechanism can be traced back to the parallelled d orbitals, which enable the electron “donation back donation” process for the molecule adsorption and activation.41 Conversely, s-block metal, alkali, and alkaline-earth metal atom sites have been rarely studied due to the lack of combined empty and occupied orbitals.42 Doping p-block atoms into the metal catalysts can modulate the reaction energetics and significantly enhance the catalytic activity, often surpassing the effects achieved with d-block dopants.43 For NRR, metal Mo and Fe-based atom catalysts have been identified as highly active and are the focus of extensive research studies. In the NORR, Fe and Co are highly active.
 |
| Fig. 4 Frequency of elements applied in SACs towards NH3 electrosynthesis. | |
The metal atoms influence the catalytic functionality in SACs, and depend on the synergy with the ligands incorporated in supports. For instance, metal sites are studied as active centers in carbon substrates.44 In metal alloys or metal compounds, single metal sites are often considered to modulate the electronic structure of the host material, or to enable the single metal sites and metals/compounds to form synergistic interactions.45 In either form, single-atom sites have demonstrated significant advantages for NH3 electrosynthesis although some issues need to be resolved.
4. Small nitrogenous substances for NH3 electrosynthesis
4.1 N2-to-NH3 electrosynthesis
The earth-abundant N2 in the atmosphere is approximately 78%, making it highly abundant and essentially zero-cost raw materials. However, the negative properties of N2, including formidable ionization energy (15.84 eV), highly stable triple N
N bond, and zero dipole moment, hinder high-performance NH3 electrosynthesis.46 Biological N2 fixation in many plants is facilitated by the enzyme nitrogenase under mild conditions, which directly converts N2 to NH3, providing growth nutrients.35,47 However, the slow reaction rate of nitrogenase in the biological nitrogen fixation process has prompted significant efforts to explore alternative sustainable routes. Nitrogenase consists of two multi-subunit proteins: Fe protein and Fe–Mo protein.48,49 Thus, in early studies, attempts were made to synthesize enzyme-like electrocatalysts with Fe and Mo sites to achieve N2 electroreduction to NH3.50 With further research advancements, some other metal active centers have also been extensively studied (Table 1). To date, to the best of our knowledge, the reported SACs for N2-to-NH3 electrosynthesis include Au, Ru, Mo, Fe, and Co atoms supported on carbon substrates and metal compounds, in which metal atoms are anchored to supports exclusively through M–N/C/O/S coordination bonds or metal bonds.
Table 1 Summary of the recently reported SACs for NRR
Single metal atom |
Catalyst |
Electrolyte |
FE |
NH3 yield rate |
Stability |
Ref. |
Noble metal atom |
Ru@ZrO2/NC |
0.1 M HCl |
21% |
3.665 mg h−1 mgRu−1 |
60 h |
51
|
Au1/C3N4 |
0.005 M H2SO4 |
11.1% |
1.305 mg h−1 mgAu−1 |
∼2 h |
52
|
Ru SAs/N–C |
0.05 M H2SO4 |
29.6% |
120.9 μg h−1 mgcat.−1 |
12 h |
53
|
SA-Ag/NC |
0.1 M HCl |
21.9% |
270.9 μg h−1 mgcat.−1 |
60 h |
54
|
Mo |
SA-Mo/NPC |
0.1 M KOH |
14.6 ± 1.6% |
34.0 ± 3.6 μg h−1 mgcat.−1 |
— |
55
|
Mo-PTA@CNT |
0.1 M K2SO4 |
83 ± 1% |
51 ± 1 μg h−1 mgcat.−1 |
2 h |
56
|
Mo-SAs/AC |
0.1 M Na2SO4 |
57.54 ± 6.98% |
2.55 ± 0.31 mg h−1 mgMo−1 |
10 h |
57
|
MoSAs-Mo2C |
0.005 M H2SO4 and 0.1 M K2SO4 |
7.1% |
16.1 μg h−1 mgcat.−1 |
10 h |
58
|
Fe |
Fe–MoS2 |
0.1 M KCl |
31.6 ± 2% |
97.5 ± 6 μg h−1 cm−2 |
20 h |
59
|
Fe–B/N–C |
0.1 M HCl |
23.0% |
100.1 μg h−1 mgcat.−1 |
30 h |
60
|
Fe–(O–C2)4 |
0.1 M KOH |
51.0% |
307.7 μg h−1 mgcat.−1 |
40 h |
61
|
FeMo/NC |
0.1 M PBS |
11.8 ± 0.8% |
26.5 ± 0.8 μg h−1 mgcat.−1 |
100 000 s |
62
|
Ni |
Ni–Nx–C-700-3 h |
0.1 M KOH |
21 ± 1.9% |
115 μg h−1 cm−2 |
20 h |
63
|
Ni–Mo2C/NC |
0.5 M K2SO4 |
29.05% |
46.49 μg h−1 mg−1 |
12 h |
64
|
Co |
Co/NC_500 |
0.1 M KOH |
10.1% |
5.1 μg h−1 mgcat.−1 |
6 h |
65
|
Zn |
Zn/NC NSs |
0.1 M Na2SO4 |
95.8% |
46.62 μg h−1 mgcat.−1 |
12 h |
66
|
V |
O–V2–NC |
0.1 M HCl |
77% |
9.97 μg h−1 mgcat.−1 |
10 cycles |
67
|
Al |
Al–Co3O4/NF |
0.1 M KOH |
6.25% |
6.48 × 10−11 mol s−1 cm−2 |
24 h |
68
|
4.1.1 Noble metal sites.
Various noble-metal single-atom materials have been initially developed as NRR electrocatalysts at aqueous electrolytes, showing favorable activities. Sun et al. first used Ru single sites supported on N-doped porous carbon for ambient aqueous N2-to-NH3 electrosynthesis.51 They found that the catalyst with a Ru–N–C2 structure provided a high NH3 yield rate of 3.665 mgNH3 h−1 mgRu−1 at −0.21 V vs. RHE. Zeng et al. designed Ru single atoms distributed on N-doped carbon (Ru SAs/N–C) with a Ru–N3/Ru–N4 structure. The composite exhibited a high FE of 29.6% and a high NH3 yield rate of 120.9 μg h−1 mg−1.52 Li et al. prepared atomically dispersed Au on carbon nitride (Au1/C3N4) and investigated its NRR catalytic performance. Owing to the efficient utilization of Au atom, a high NH3 yield rate of Au1/C3N4 was obtained, which was approximately 22.5 times higher than that of Au nanoparticles.53 Luo et al. synthesized a Ag single-atom catalyst (SA-Ag/NC) with the Ag–N4 structure and compared the NRR performance with Ag nanoparticles/NC. The comparative analysis result showed that the SA-Ag/NC exhibited an NH3 yield rate of 270.9 μg h−1 mg−1 and a FE of 21.9%, which was significantly higher than that of Ag nanoparticles/NC. Despite the excellent performance of these noble-metal single-atom catalysts, further advancement and widespread use of various non-noble metal single-atom catalysts continue to be an attractive option.
4.1.2 Fe site.
To construct a bimetal-atom active center to mimic nitrogenase in nature, the Fe atom and its synergistic effect with the Mo atom have been widely investigated for N2-to-NH3 electrosynthesis. Yao et al. prepared Fe2(MoO4)3, and found that it exhibited a high NH3 yield rate of 7.5 μg h−1 mg−1 and an NH3 FE of 1.0% at −0.7 V for 2 h under ambient conditions (0.1 M Na2SO4), which was attributed to the synergistic effect of Fe and Mo elements in the active center.69 Zhao et al. prepared atomically dispersed Fe–(O–C2)4 sites on graphitic carbon. The N-free catalysts with the Fe–O active sites prevented interference caused by the presence of other NRR active coordination structures, especially Fe–Nx active sites.61 The Fe single-atom catalyst exhibited an NH3 yield rate of 32.1 mg h−1 mg−1 and a FE of 29.3%. Zhang et al. found that the interfacial polarization of the Fe atom on MoS2 efficiently enhances the N
N fracture to boost the electrocatalytic NH3 synthesis.59 They prepared the protrusion-like Fe single-atom onto MoS2 nanosheets, which triggered the interfacial polarization (Fig. 5a). The theoretical calculation results showed that the protrusion-shaped Fe single atoms engendered electric fields to polarize N2, and the energy barrier was mitigated in a set for electron flow from the Fe sites to N2 (Fig. 5b). The NH3 FE and NH3 yield rates closely depend on the Fe-atom concentration, confirming Fe atoms as the catalytic centers (Fig. 5c).
 |
| Fig. 5 (a) Side-view and perspective-view crystalline structures of Fe–MoS2. (b) The activation energies and (c) performance comparison of various Fe–MoS2 catalysts with different Fe loading amounts. Reproduced with permission from ref. 59 Copyright 2020, Elsevier. | |
4.1.3 Mo site.
In 2017, Chen et al. systematically investigated the potential of transition metal atoms (Sc to Zn, Mo, Ru, Rh, Pd, and Ag) through DFT simulation to predict their immobilization on defective boron nitride monolayers as the NRR electrocatalysts.70 The result showed that atomically dispersed Mo atoms bonded to N atoms exhibited good performances in activating N2 molecules and stabilizing *N2H, while destabilizing the *NH2 species during the electrochemical NRR process. Furthermore, Xin et al. experimentally verified the catalytic activity of the Mo single atom.55 Mo single atoms anchored to N-doped porous carbon (SA-Mo/NPC) were prepared as a cost-effective catalyst for N2-to-NH3 electrosynthesis. Owing to the optimized abundance of Mo active sites and the three-dimensional hierarchically porous feature, the SA-Mo/NPC catalyst exhibited a remarkable FE of up to 14.6% at −0.3 V vs. RHE in 0.1 M KOH. Jiang et al. prepared a Mo-PTA@CNT catalyst, in which Mo atoms were anchored onto the fourfold hollow sites of phosphotungstic acid and closely embedded in multi-walled carbon nanotubes (CNTs) (Fig. 6a).56 The experimental results showed that Mo(IV) on the 4-H sites served as active sites for N2 adsorption and activation. The optimal Mo-PTA@CNT exhibited a high NH3 yield rate of approximately 51 μg h−1 mgcat.−1 and an excellent FE of approximately 83% at −0.1 V vs. RHE under ambient conditions (Fig. 6b). Wang et al. revealed the role of Mo single atoms in the MoSAs-Mo2C/NCNT electrocatalyst comprising Mo single atoms and Mo2C nanoparticles grown on the N-doped CNTs.58 The electrochemically experimental results showed that MoSAs-Mo2C/NCNTs exhibited a high yield rate of 16.1 μg h−1 cmcat.−2, which was approximately four times that of the Mo2C/NCNTs at the same potential. Moreover, the FE of the MoSAs-Mo2C/NCNTs is twice that of the Mo2C/NCNTs. DFT calculations revealed that the HER-selective MoSAs in the combined entity provided a large *H coverage environment around the Mo2C nanoparticles to activate the N2 molecules, thus improving the overall NRR performance (Fig. 6c and d).
 |
| Fig. 6 (a) Probable crystalline structures of PTA and Mo-PTA. (b) NH3 yield rates and FEs of Mo-PTA@CNT and PTA@CNT. Reproduced with permission from ref. 56 Copyright 2021, Wiley-VCH. The Gibbs free energy diagrams on the MoNC2 SAC and the Mo2C (101) surface along (c) the NRR pathways and (d) the HER. Reproduced with permission from ref. 58 Copyright 2020, Wiley-VCH. | |
4.1.4 Other metal sites.
Research on other single metal atoms has mainly focused on the influence of metal sites on NRR performance under different substrates. Various metal–N–C materials have been investigated as NRR electrocatalysts, in which the metal sites include Ni, Co, Zn, and V. Similar to the Fe and Mo sites, single metal sites on carbon usually reduce HER and promote N2 dissociation.65,67 Wu et al. prepared atomically dispersed Ni sites on the carbon (Ni–Nx–C) catalyst with a NiN3 structure. The result showed that the catalyst exhibited a high NH3 yield rate of 115 μg cm−2 h−1 at −0.8 V vs. RHE and a maximum FE of 20% at −0.6 V vs. RHE.52 By comparison, Ni clusters supported on N-doped carbon had no significant NRR activity, confirming that atomically dispersed unsaturated Ni atoms were the real active sites. Shao et al. prepared atomically dispersed Zn supported on N-doped carbon nanosheets (Zn/NC NSs) with the Zn–N4 structure as the NRR catalyst, achieving excellent NRR performance.66 DFT calculations showed that the introduction of Zn sites effectively reduced the energy barrier for the formation of *NNH. In addition to the carbon substrates, metal compounds have been widely used as substrates to support single atoms. More studies have shown that single metal sites on metal compounds can generate the synergistic effect of active sites for the NRR process. Hu et al. prepared Ni atom-doped Mo2C anchored on porous carbon, and found that the electronic configuration close to the Ni–Mo active sites was optimized. The N2 adsorption was also promoted due to the increased electron transfer.64 Yuan et al. introduced Al atoms into Co3O4 to effectively tune the electronic properties of the catalyst, and the increased surface oxygen vacancies induced by Al doping facilitated the N2 activation.68
4.2 NO-to-NH3 electrosynthesis
NO is more active than N2, thus making electrochemical NORR thermodynamically more feasible than NRR. NO, one of the major pollutants emitted during fossil fuel combustion and other chemical industrial processes, can be converted into valuable products through electrocatalysis, thereby alleviating the environmental load and reversing the anthropogenic global nitrogen-cycle imbalance. As mentioned in the previous section, *NO is a key intermediate that determines the NH3 selectivity in the electroreduction process of these nitrogen oxides.71 Therefore, investigating NORR is crucial for gaining a fundamental understanding of the electroreduction process and achieving efficient NH3 electrosynthesis. Liu et al. reported on single atomic Ce sites anchored on N-doped hollow carbon spheres (Ce-SA/NHCS) that could electrocatalyze NO reduction to NH3 in an acidic solution.72 DFT calculations showed that Ce-SA/NHCS was not beneficial for adsorption of *H compared with *NO. When the SCN− ions were added to the electrolyte due to the blocking effect, the NH3 yield rate of Ce-SA/NHCS rapidly decreased, indicating that the Ce–N4 moieties in Ce-SA/NHCS were active sites. In a typical NORR process, the adsorption configuration of NO was also crucial to modulating the bonding energy of N–O, which could further regulate the NH3 selectivity. The bonding energy of N–O can be modulated through the adsorption configuration of NO on the surface-active sites of electrocatalysts. Our group constructed and compared the Co single-atom catalyst (Co SACs) and the Co nanoparticle (Co NPs) catalyst for NORR.73 Under neutral conditions, Co NPs were inclined to generate NH3, and Co SACs exhibited high NO-to-NH2OH selectivity (Fig. 7a). DFT calculations revealed that the turnover frequencies (TOF) of NH2OH were 50 times higher than the TOF of NH3 on Co SACs (Fig. 7b). The linear adsorption of NO on isolated Co sites might maintain the N–O bond to produce NH2OH through the intermediates of *HNO and *H2NO. Additionally, Co NPs could break the N–O bond to form NH3 through a dissociation mechanism (Fig. 7c).
 |
| Fig. 7 (a) FEs of NH3 and NH2OH on Co SACs and Co NPs at different potentials. (b) TOFs of NH3 and NH2OH over Co SACs obtained by microkinetic modeling. (c) Deduced electrocatalytic NORR pathway on Co SACs and Co NPs for NH2OH and NH3 production. Reproduced with permission from ref. 73 Copyright 2023, Wiley-VCH. | |
Based on this, well-designed single metal atoms anchored on metal-based supports might make it easier to obtain the optimal adsorption configuration of NO, leading to the generation of the target NH3 product.74 Studies have shown that single metal atoms, including Co, Ni, Mn, and Fe, supported on various metal/metal compound substrates are promising NO-to-NH3 electrocatalysts with excellent activity and selectivity.75–77 Zhu et al. successfully combined DFT and machine learning methods to examine Fe@MoS2. They found that the Fe@MoS2 composite was thermodynamically stable, and had strong resistance to electrochemical corrosion under local alloying conditions.77 Chu et al. conducted various studies on NO-to-NH3 electrosynthesis. They reported that atomically dispersed Co anchored on MoS2 (Co1/MoS2) was an efficient NORR catalyst, highlighting Co as a critical element in the Co1/MoS2 catalyst for effective NORR and NOx denitrification.78 Structural characterizations and theoretical calculations unveiled the formation of active Co1–S3 moieties on Co1/MoS2, which could selectively adsorb NO molecules and promote the hydrogenation energetics of the NO-to-NH3 electroreduction process (Fig. 8a, c and d). As a result, Co1/MoS2 exhibited excellent NORR performance with an NH3 yield rate of 217.6 μmol h−1 cm−2 and an NH3 FE of 87.7% at −0.5 V vs. RHE (Fig. 8b). They also designed p-block Sb single atoms confined in amorphous MoO3 (Sb1/a-MoO3) evaluated as the NORR catalyst.79 Compared with the a-MoO3 catalyst, Sb1/a-MoO3 exhibited a high NH3 yield rate of 273.5 μmol h−1 cm−2 and a NH3 FE of 91.7% at −0.6 V vs. RHE. In situ spectroscopic characterizations and theoretical calculations showed that the isolated Sb1 sites optimized the adsorption of *NO/*NHO to reduce the reaction energy barriers, and simultaneously exhibited a higher affinity to NO than to H2O/H species. In another work, they designed single-atom Pd-alloyed Cu (Pd1Cu) as the NORR catalyst, which exhibited excellent long-term durability for 200 h at industrial-level current densities (>0.2 A cm−2).80
 |
| Fig. 8 (a) EXAFS fitting curve of Co1/MoS2. (b) NH3 yield rate and FE of MoS2 and Co1/MoS2. (c) Binding free energies of *NO and *H on Co1–S3. (d) Free energy profiles of the NORR process on MoS2 and Co1/MoS2. Reproduced with permission from ref. 78 Copyright 2023, Elsevier. | |
4.3 NO2−/NO3−-to-NH3 electrosynthesis
Electrochemical NO2−RR and NO3−RR usually exhibit faster reaction rates, owing to the relatively lower dissociation energy of the N
O bond and higher solubility of NO2−/NO3− in the aqueous electrolytes. In addition, converting NO2−/NO3− pollutants into high-value-added products provides a promising approach for both wastewater treatment and NH3 synthesis. In recent years, significant progress has been made on SACs for the NO2−RR and NO3−RR. NO2− is the main byproduct of NO3− electroreduction at low overpotential and has been reported in many cases under a two-electron reduction process. Zhi et al. compared the four mentioned nitrogenous substances. Their results showed that NO2− exhibited the highest calculated reduction potential of 0.897 V vs. RHE at a pH of 0, and KNO2 exhibited the highest solubility (Fig. 9).81 Therefore, using NO2− as the raw material for NH3 electrosynthesis can provide faster reaction thermodynamics and kinetics, and a higher NH3 yield rate. Chu et al. designed atomically dispersed Pd on defective BN nanosheets (Pd1/BN) as a potential NO2−RR catalyst, achieving a high NH3 FE of 91.7% and an NH3 yield rate of 347.1 μmol h−1 cm−2 at −0.6 V vs. RHE.82 The theoretical calculations revealed that isolated Pd atoms served as catalytic centers, selectively adsorbing NO2− and enhancing the hydrogenation process with a minimized reaction barrier. In another work, they revealed that Mo single atoms in the Mo1–ZrO2 electrocatalyst effectively activated NO2− and facilitated the NO2−RR protonation process.83
 |
| Fig. 9 (a) Calculated potentials and chemical reactions of N2, NO, NO2− and NO3− reduction to NH4+. (b) The solubility of N2(g), NO(g), KNO2, and KNO3 in 100 mL of water under one atmosphere and at 298.15 K. Reproduced with permission from ref. 81 Copyright 2022, The Royal Society of Chemistry. | |
Although NO2− has many theoretical advantages, its relatively unstable decomposition and toxicity limit its development in the field of NH3 electrosynthesis. Considerable attention has been paid to the electroreduction of NO3− to NH3. Currently, metal atoms, including Cu, Fe, Ru, and Bi, anchored on a set of supports (including N-doped carbon, metal, and metal compounds) have been extensively investigated in NO3−RR.84
4.3.1 Cu site.
Cu-based electrocatalysts have been extensively reported for NO3−RR, owing to the benefit of NO3− adsorption and matching electronic structure with the molecular orbital of NO3−.85–87 Due to its weak adsorption, the accumulation of NO2− results in sluggish kinetics of the subsequent hydrogenation steps and the formation of N2 and N2O byproducts.87 The isolated Cu single-atom sites can effectively minimize the possibility for the formation of N2 and N2O byproducts during the NO3−RR process. Li et al. compared a series of Cus+n/NC with mixed Cu single atoms (Cu SAs) and nanoparticles incorporated in N-doped carbon, and their counterpart Cu nanoparticles (Cu NPs) for NO3−-to-NH3 electrosynthesis (Fig. 10a–d).88 They systematically elucidated the activity trends of Cu SA/NC, Cus+n/NC, and Cu NPs. The result showed that Cu SA/NC decreased NH3 FE as the Cu single-atom content decreased. However, the production of N2 and N2O gradually increased, which reached the maximum on pure Cu NPs. This result showed that the high and low proportions of Cu SAs in Cus+n/NC facilitates the formation of NH3 and N–N coupled products (N2 and N2O), respectively. Theoretical calculations revealed that the higher NH3 FE of Cu SA/NC was attributed to the lower free energy of the rate-limiting step (*ONH → *N) and effective inhibition for the N–N coupled process.
 |
| Fig. 10 (a) Schematics of the NO3−RR on Cu SAs and Cu nanoparticles. FEs of (b) Cu(0.1 wt%)/NC and (c) Cu(1.6 wt%)/NC at different densities. (d) Free energy diagram of the reaction coordinates of Cu SA/NC and Cu NPs. Reproduced with permission from ref. 88 Copyright 2023, American Chemical Society. (e) Cu K-edge XANES spectra and (f) Cu K-edge FT-EXAFS spectra at different potentials. (g) Proposed mechanisms of aggregation and re-dispersion. Reproduced with permission from ref. 89 Copyright 2022, American Chemical Society. | |
In addition to the symmetrical four N coordination structure, Cu single-atom catalysts with different coordination structures for NO3−RR were investigated to further regulate the NO3−RR performance.89 Zhang et al. found that the coordination environment of the Cu single-atom sites (Cu(I)–N3C1) localized the charge around the central Cu atoms, and adsorbed *NO3 and *H onto neighboring Cu and C sites with balanced adsorption energy.90 Zhu et al. designed and synthesized a Cu single-atom catalyst with a low-coordinated Cu–N3 structure on high-curvature hierarchically porous N-doped CNTs. The catalyst exhibited enhanced NO3−RR activity with an NH3 yield rate of 30.1 mg h−1 mg−1 and an NH3 FE of 89.6%.91 Theoretical calculations showed that low-coordinated Cu–N3 sites and high-curvature carbon support contributed to the fast reaction dynamics and low rate-determining step barrier for NO3−RR. The Gibbs free energies of the intermediates for the Cu–N3-plane were generally lower than those of the Cu–N4-plane, indicating that the Cu–N3 structure had better reaction dynamics than the Cu–N4 structure. Lu et al. designed the symmetry-broken Cu-cis-N2O2 single-atom catalysts in the cis-configuration.92 The low coordination symmetry rendered the active site more polar and accumulated more NO3− near the electrocatalyst surface. Theoretical calculations revealed that the cis-coordination splits the Cu 3d orbitals generated an orbital-symmetry-matched π-complex of the key intermediate *ONH, and reduced the energy barrier when compared with Cu–N4 and Cu-trans-N2O2. In addition, an average NH3 yield rate of 27.84 mg h−1 cm−2 at an industrial level current density of 366 mA cm−2 was retained for more than 2000 h. Wang et al. identified the real active site of the Cu–N4 single-atom catalyst in NO3−RR using operando X-ray absorption spectroscopy (Fig. 10e and f).93 They found that the synthesized Cu–N4 structure successively transformed to Cu–N3, near-free Cu0 single atoms, and eventually to aggregated Cu0 nanoparticles with the applied potential ranging from 0.0 to −1.0 V vs. RHE. After electrolysis, the aggregated Cu0 nanoparticles were disintegrated into single atoms, and then restored to the Cu–N4 structure after being exposed to an ambient atmosphere (Fig. 10g).
4.3.2 Fe site.
Zero-valence Fe has been widely reported as an effective catalyst for removing NO3− in groundwater to manage nitrogenous pollution.94 In addition, Fe-dependent nitrite reductase with Fe–N4 can actively produce NH3 through the photosynthetic nitrate assimilation pathway in nature.95 Jia et al. investigated electrochemical NO3−RR on traditional metal–N3 and –N4 (traditional metal = Ti, V, Cr, Mn, Fe, Co, Ni, and Cu) doped graphene using first-principles calculations.96 The result showed that Fe–N-doped graphene with Fe–N4 structure exhibited excellent NO3−RR performance, which was attributed to the effective adsorption and activation of NO3− through the charge “acceptance-donation” mechanism and its moderate binding due to the occupation of the d–p antibonding orbital. Inspired by the similar coordination structure, atomically dispersed Fe on the carbon substrates with various configurations were investigated as the NO3−RR catalysts.97,98 In 2021, Wang et al. reported on the Fe–N4 structure as an active and selective site for NO3−-to-NH3 electrosynthesis.97 They found that the Fe–N4 single-atom catalyst effectively prevented the N–N coupling step required for N2 due to the insufficient neighboring metal sites. Atanassov et al. investigated the NO3−RR mechanism of Fe–N4 active sites in Fe–Mo dual single-atom catalysts.99 They found that one of the Fe–N4 sites facilitated the transfer of NO2− to the NH3 transfer pathway with approximately 100% NH3 FE and a high yield rate in the NO3−RR process (Fig. 11a–f). The NO3−RR mechanism of the Fe–N4 coordination structure was also identified in another Fe–Mo dual single-atom catalyst system reported by Li et al.100 Yu et al. fabricated atomically dispersed N-coordinated Fe sites on carbon (Fe-PPy SACs) for NO3−-to-NH3 electrosynthesis.39 The Fe-PPy SACs exhibited nearly 100% NH3 selectivity, and a TOF that was over 12 times that of Fe nanoparticles. The mechanistic study revealed the exclusive existence of NO3−-preoccupied Fe(II)–Nx sites, which effectively eliminated the competing H2O adsorption under aqueous conditions; thus, no noticeable H2 byproduct was produced (Fig. 11g–i). Li et al. compared the NO3−RR performance of a zeolitic imidazolate framework-derived Fe single-atom catalyst with Fe–N3S1 structure (Fe–N3S1 SACs) and commercial FeS.101 The Fe–N3S1 SACs exhibited a higher NH3 FE of 93.9% than FeS (67.2%). DFT calculations revealed the easier adsorption of NO3− by forming the *NO3 intermediate on Fe–N3S1. Gu et al. prepared another Fe single-atom catalyst with a Fe–N2O4 coordination structure through pyrolysis of metal–organic frameworks.102 DFT calculations revealed that the Fe–N2O2 structure exhibited higher conductivity and NH3 selectivity than the Fe–N4 structure, which spontaneously triggered the transformation of *NOH to *N. Another metal atom that was close to the Fe atom might produce different NO3−RR properties. Lu et al. synthesized a Fe/Cu diatomic catalyst on holey N-doped graphene, and the original coordination structure of the Fe sites was broken.103 Compared with the Fe/Fe and Cu/Cu configurations, the Fe/Cu diatomic sites provided medium interactions toward most other intermediates and reduced the overall energy barriers for the NO3−RR process.
 |
| Fig. 11 (a) Associative adsorption of NO3− on the Fe–N4 site and (b) dissociative adsorption of NO3− into *O and NO2− on the Mo–N4 site. (c) Proposed catalytic cascade for synergized NO3−RR on the Mo–N4 and Fe–N4 sites. NH3 and NO2− concentrations over a 24 h period of NO3−RR electrolysis catalyzed by (d) Fe–N4, (e) Mo–N4, and (f) FeMo–N–C catalysts. Reproduced with permission from ref. 99 Copyright 2022, American Chemical Society. (g) The proposed preoccupied NO3−RR mechanism for the single-site center and classical competitive mechanism for the bulk surface. (h) TOFs of the Fe-PPy SACs and Fe NPs for NH3 production. (i) Gibbs free-energy diagram of NO3−RR and HER. Reproduced with permission from ref. 39 Copyright 2021, The Royal Society of Chemistry. | |
When the Fe site is anchored on metal-based substrates, it may not serve as the direct active site for the NO3−RR process. As a result, it effectively tunes the d-band center of the metal substrates and modulates the electronic structure of the overall catalysts. Zhi et al. reported a Fe-doped Ni2P nanosheet as the NO3−RR catalyst, which showed larger NH3 yield rates and NH3 FE when compared with Ni2P at each given potential.104 Theoretical calculations revealed that the Fe site caused a downshift of the d-band center of Ni atoms to the Fermi level, optimizing the Gibbs free energies for the NO3−RR intermediates. Chu et al. reported on a Fe-doped V2O5 (Fe-V2O5), in which Fe dopants were in isolated single-atom states, and a high density of Lewis acid Fe–V pairs was readily created.105 Theoretical calculations revealed that Lewis acid Fe–V pairs in Fe-V2O5 synergistically enhanced the NO3− activation and hydrogenation, and concurrently suppressed the HER.
4.3.3 Ru site.
Ru exhibits suitable adsorption/desorption energy toward hydrogen intermediates, and it is considered one of the outstanding active sites for alkaline–water splitting to produce H2.106 Recent studies have shown that Ru-based NO3−RR catalysts exhibit low overpotentials and high performance in the electrochemical reduction of NO3− to NH3.107–110 Wang et al. designed a Ru1Cu catalyst, which showed an industrial-level current density of approximately 1.0 A cm−2 with an NH3 FE of 93%. Theoretical investigations unveiled that the excellent NO3−RR performance of Ru1Cu was attributed to the synergy of Ru and Cu, where Ru sites activated NO3− while the surrounding Cu sites suppressed the HER. Our group incorporated Ru atoms into the octahedral sites of Co3O4 to achieve an NH3 yield rate of 24.6 mg h−1 cm−2.111 Electrochemical in situ spectroscopic analyses and theoretical calculations revealed that Ru sites improved the water molecule coverage and facilitated the *H production, leading to stable and orderly NH3 production.
4.3.4 Other metal sites.
The p-block metals (such as Bi, Sn, In) are poorly bound to H-adatoms.112 Bi, as a typical metal, has relatively low catalytic activity for the HER and strong adsorption/activation ability toward the nitrogen-containing intermediates, making it widely used in small molecule catalytic activation including NO3− electroreduction.113–115 Oschatz et al. reported on a porous Bi single-atom decorated, N-doped carbon with a BiN2C2 coordination structure.116 DFT calculations revealed that the Bi active center exhibited a relatively low charge, resulting in an unfavorable NO3− adsorption that was nonetheless beneficial for stabilizing the hydrogenated intermediates. Recent studies have shown that the incorporation of a p-block metal atom into d-block metal compounds can trigger a fascinating p–d orbital hybridization effect, effectively overcoming the limitations of d-metal compounds in facilitating the HER.117,118 Chu et al. reported that Bi single atoms significantly improved the NO3−RR performance when alloyed with the Pd metal (Fig. 12a and b). DFT calculations revealed that doping Bi single atoms into the Pd metal reduced the energy barrier for the *NO to *NOH hydrogenation process (Fig. 12d).119 This enhancement was attributed to the inhibitory effect of Bi on HER, which prevented *H coupling, resulting in more *H, and the strong interaction between the Bi 6p orbitals and N 2p orbitals, which increased the adsorption capacity of nitrogen-containing reaction intermediates (Fig. 12c). Bader charge analysis revealed the presence of a downshifted d-band center of Pd1 (−1.55 eV) relative to the Pd atom (−1.48 eV) of pristine Pd, effectively destabilizing *NO while retaining *NOH stabilization (Fig. 12e). In another work, Chu et al. designed Bi-doped FeS2, triggering a fascinating p–d orbital hybridization effect.120 The electronic structure of the active centers and the adsorption/desorption behaviors of the N-containing intermediates were optimized, thus enhancing the NO3−RR catalytic properties.
 |
| Fig. 12 (a) NH3 yield rate and (b) NH3 FE of Pd and Bi1Pd at various potentials. (c) Fitted lines of the EPR intensity/reaction time versus NO3− concentration. (d) Free energy diagram of NO3−RR pathways on Pd and Bi1Pd; inset: atomic configurations of *NO on Pd and Bi1Pd. (e) PDOS of Pd atom on Pd and Pd1, Pd2 and Bi1 atoms on Bi1Pd. Reproduced with permission from ref. 119 Copyright 2023, Wiley-VCH. | |
In addition to these single metal atoms, some other metal atom sites have similar advantages, in which one of the typical metal sites is Co.121 Zhu et al. synthesized Co single atoms anchored on carbon nanofibers (Co SAs/CNFs) with a Co–C3O1 coordination structure.122 The experimental and theoretical results revealed that the synthesized Co SAs/CNF electrocatalyst exhibited excellent electrocatalytic NO3−RR activity, a smaller energy barrier for the rate-determining step of *NO to *NOH, and a stronger capability for inhibiting HER than Co nanoparticles. This regulation mechanism was also confirmed in another study which compared the Co–P1N3 and Co–N4 reported by Liu et al.123 Zhao et al. also developed the N, O trans-coordination Ag single atoms on carbon nanotubes, which inhibited the HER, optimized the adsorption of intermediates, and facilitated the potential-determining step of *NO → *NOH in the NO3−RR process.124
5. Conclusion and outlook
Electroreduction of small nitrogenous substances (N2, NO, NO2−, and NO3−) for NH3 electrosynthesis has recently attracted considerable attention. Metal single-atom catalysts (SACs) have made significant advancements in the field of NH3 electrosynthesis, particularly in terms of enhancing catalytic performance and elucidating fundamental mechanisms. This review concisely compares the features of diverse nitrogenous substances and provides a synopsis of the impact of these properties on the NH3 electrosynthesis. This review also summarizes the frequently used metal single-atom sites in SACs and explores their role in various reaction processes involved in the NH3 electrosynthesis. The reported computational predictions and experimental findings on various single metal sites for electrochemical NH3 synthesis are summarized. Regarding the N2-to-NH3 electroreduction synthesis, Mo, Fe, and V sites, which are the metal centers in nitrogenase enzymes, have been widely studied and exhibit excellent performance. Electroreduction of NO/NO2−/NO3− to NH3 has a similar reaction path, where NO, as the reaction center of divergence, is often used to study the mechanism. Experimental reports on the electroreduction of NO and NO2− to NH3 are currently limited. In the electroreduction of NO3−-to-NH3 reaction, Cu, Fe, Ru, Bi, and other single-atom sites have been extensively studied. The coordination structure around them and supports around the metal sites play a regulating role in the rate-determining step, water splitting, and NH3 selectivity in this reaction.
Although SACs have successfully improved NH3 electrosynthesis, several challenges still need to be addressed. Several perspectives have been outlined to advance nitrogen-containing substances to NH3 electrosynthesis, focusing on enhancing the catalytic activity and providing fundamental insights to guide future research. (1) The preparation process of SACs is relatively complex, requiring precise control of reaction conditions and the dispersion and anchoring of single metal atoms on the catalytic supports. This requirement increases the difficulty and cost of preparation, limiting the possibility of large-scale application. Therefore, universal and simple synthesis methods should be encouraged in the future. (2) The long-term stability of SACs may be compromised by environmental factors and reaction conditions, such as humidity, gas concentration, applied voltage, and pH. These factors can affect the stability of single metal atoms, reducing or deactivating catalyst activity. Notably, the stability of the previous single-atom structure under highly reducing conditions remains an open question, as changes in the metal valence may alter the catalyst structure. Additionally, compared with metal and metal compound catalysts, SACs have a more complex and diverse structure, making the structure–activity mechanisms and relationships in NH3 electrosynthesis more complex. Unraveling the dynamic evolution of the SACs during electrolysis is crucial to identifying the actual active site. Therefore, further extensive research and development efforts are needed to fully harness the potential of SACs, especially using advanced technical methods (such as in situ spectroscopic techniques and theoretical calculations). (3) Although SACs have broad application prospects, there are still some challenges in complex continuous electrocatalytic hydrogenation reactions. Therefore, the design and development of new types of SACs are necessary. For instance, bi-/multi-metal-atom catalysts with multiple active centers may potentially provide higher activity than SACs through the synergistic action of two or more atoms.
Author contributions
Zhaole Lu: investigation, writing – original draft, and writing – review & editing. Jijie Zhang: conceptualization, writing – review & editing, and supervision. Yuting Wang: writing – review & editing. Yifu Yu: supervision. Lingjun Kong: supervision.
Data availability
No primary research results, software or code have been included, and no new data were generated or analysed as part of this review.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
We acknowledge the National Natural Science Foundation of China (no. 22205156 and 22475146).
Notes and references
- J. W. Erisman, M. A. Sutton, J. Galloway, Z. Klimont and W. Winiwarter, Nat. Geosci., 2008, 1, 636–639 CrossRef CAS.
- R. Service, Ammonia, a renewable fuel made from sun, air, and water, could power the globe without carbon, Science [News], 2018 DOI:10.1126/science.aau7489.
- L. Wang, M. Xia, H. Wang, K. Huang, C. Qian, C. T. Maravelias and G. A. Ozin, Joule, 2018, 2, 1055–1074 CrossRef CAS.
- R. F. Service, Science, 2014, 345, 610 CrossRef CAS PubMed.
- V. Rosca, M. Duca, M. T. de Groot and M. T. M. Koper, Chem. Rev., 2009, 109, 2209–2244 CrossRef CAS PubMed.
- Y. Wang, W. Zhou, R. Jia, Y. Yu and B. Zhang, Angew. Chem., Int. Ed., 2020, 59, 5350–5354 CrossRef CAS PubMed.
- J. Meng, C. Cheng, Y. Wang, Y. Yu and B. Zhang, J. Am. Chem. Soc., 2024, 146, 10044–10051 CrossRef CAS PubMed.
- Y. Xu, C. Cheng, J. Zhu, B. Zhang, Y. Wang and Y. Yu, Angew. Chem., Int. Ed., 2024, 63, e202400289 CrossRef CAS PubMed.
- Z. Zhang, H. Zhang, H. Jiang and L. Li, J. Mater. Chem. A, 2024, 12, 33334–33361 RSC.
- M. A. Shipman and M. D. Symes, Catal. Today, 2017, 286, 57–68 CrossRef CAS.
- S. Li, J. Liang, P. Wei, Q. Liu, L. Xie, Y. Luo and X. Sun, eScience, 2022, 2, 382–388 CrossRef.
- Y. Wang, C. Wang, M. Li, Y. Yu and B. Zhang, Chem. Soc. Rev., 2021, 50, 6720–6733 RSC.
- X. Zhang, Y. Wang, C. Liu, Y. Yu, S. Lu and B. Zhang, Chem. Eng. J., 2021, 403, 126269 CrossRef CAS.
- L. Kong, M. Zhong, W. Shuang, Y. Xu and X.-H. Bu, Chem. Soc. Rev., 2020, 49, 2378–2407 RSC.
- Y. Huang, C. Tang, Q. Li and J. Gong, Appl. Surf. Sci., 2023, 616, 156440 CrossRef CAS.
- C. Choi, S. Back, N.-Y. Kim, J. Lim, Y.-H. Kim and Y. Jung, ACS Catal., 2018, 8, 7517–7525 CrossRef CAS.
- Y. Wang, H. Su, Y. He, L. Li, S. Zhu, H. Shen, P. Xie, X. Fu, G. Zhou, C. Feng, D. Zhao, F. Xiao, X. Zhu, Y. Zeng, M. Shao, S. Chen, G. Wu, J. Zeng and C. Wang, Chem. Rev., 2020, 120, 12217–12314 CrossRef CAS PubMed.
- Y. Jiao, Y. Zheng, M. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2015, 44, 2060–2086 RSC.
- F. Rehman, S. Kwon, C. B. Musgrave, M. Tamtaji, W. A. Goddard and Z. Luo, Nano Energy, 2022, 103, 107866 CrossRef CAS.
- S. Wang, H. Gao, L. Li, K. S. Hui, D. A. Dinh, S. Wu, S. Kumar, F. Chen, Z. Shao and K. N. Hui, Nano Energy, 2022, 100, 107517 CrossRef CAS.
- C. Han, S. Zhang, H. Zhang, Y. Dong, P. Yao, Y. Du, P. Song, X. Gong and W. Xu, eScience, 2024, 4, 100269 CrossRef.
- L. Zhang, M. Zhou, A. Wang and T. Zhang, Chem. Rev., 2020, 683–733 CrossRef PubMed.
- H. Niu, Z. Zhang, X. Wang, X. Wan, C. Shao and Y. Guo, Adv. Funct. Mater., 2021, 31, 2008533 CrossRef CAS.
- X. Guo, J. Gu, S. Lin, S. Zhang, Z. Chen and S. Huang, J. Am. Ceram. Soc., 2020, 142, 5709–5721 CrossRef CAS PubMed.
- N. Gruber and J. N. Galloway, Nature, 2008, 451, 293–296 CrossRef CAS PubMed.
- M. Paidar, I. Roušar and K. Bouzek, J. Appl. Electrochem., 1999, 29, 611–617 CrossRef CAS.
- P. H. van Langevelde, I. Katsounaros and M. T. M. Koper, Joule, 2021, 5, 290–294 CrossRef.
- K. Li, D. J. Jacob, H. Liao, J. Zhu, V. Shah, L. Shen, K. H. Bates, Q. Zhang and S. Zhai, Nat. Geosci., 2019, 12, 906–910 CrossRef CAS.
- J. Liu, Z. Li, C. Lv, X.-Y. Tan, C. Lee, X. J. Loh, M. H. Chua, Z. Li, H. Pan, J. Chen, Q. Zhu, J. Xu and Q. Yan, Mater. Today, 2024, 73, 208–259 CrossRef CAS.
- P. Li, Z. Fang, Z. Jin and G. Yu, Chem. Phys. Rev., 2021, 2, 041305 CrossRef CAS.
- J. Choi, B. H. R. Suryanto, D. Wang, H.-L. Du, R. Y. Hodgetts, F. M. Ferrero Vallana, D. R. MacFarlane and A. N. Simonov, Nat. Commun., 2020, 11, 5546 CrossRef CAS PubMed.
- H. Wan, A. Bagger and J. Rossmeisl, Angew. Chem., Int. Ed., 2021, 60, 21966–21972 CrossRef CAS PubMed.
- Y. Wang and M. Shao, ACS Catal., 2022, 12, 5407–5415 Search PubMed.
- S. B. Patil and D. Y. Wang, Small, 2020, 16, 2002885 CrossRef CAS PubMed.
- C. J. M. Van Der Ham, M. T. M. Koper and D. G. H. Hetterscheid, Chem. Soc. Rev., 2014, 43, 5183–5191 RSC.
- Y. Yu, X. Wei, W. Chen, G. Qian, C. Chen, S. Wang and D. Min, ChemSusChem, 2024, 17, e202301105 CrossRef CAS PubMed.
- Y. Xiong, Y. Wang, J. Zhou, F. Liu, F. Hao and Z. Fan, Adv. Mater., 2024, 36, 2304021 CrossRef CAS PubMed.
- J. Long, S. Chen, Y. Zhang, C. Guo, X. Fu, D. Deng and J. Xiao, Angew. Chem., Int. Ed., 2020, 59, 9711–9718 CrossRef CAS PubMed.
- P. Li, Z. Jin, Z. Fang and G. Yu, Energy Environ. Sci., 2021, 14, 3522–3531 RSC.
- D. H. Kim, S. Ringe, H. Kim, S. Kim, B. Kim, G. Bae, H.-S. Oh, F. Jaouen, W. Kim, H. Kim and C. H. Choi, Nat. Commun., 2021, 12, 1856 CrossRef CAS.
- D. Wu, P. Lv, J. Wu, B. He, X. Li, K. Chu, Y. Jia and D. Ma, J. Mater. Chem. A, 2023, 11, 1817–1828 RSC.
- M.-A. Légaré, G. Bélanger-Chabot, R. D. Dewhurst, E. Welz, I. Krummenacher, B. Engels and H. Braunschweig, Science, 2018, 359, 896–900 Search PubMed.
- H. K. Lim, H. Shin, W. A. Goddard, Y. J. Hwang, B. K. Min and H. Kim, J. Am. Chem. Soc., 2014, 136, 11355–11361 CrossRef CAS PubMed.
- Z. Zhang, J. Xiao, X. Chen, S. Yu, L. Yu, R. Si, Y. Wang, X. Meng, Y. Wang, Z. Tian and D. Deng, Angew. Chem., Int. Ed., 2018, 57, 16339–16342 Search PubMed.
- L. Wu, S. Sun, L. Zhang, R. Wang, J. Feng, X. Sun and B. Han, Sci. China: Chem., 2024, 67, 1969–1975 CrossRef CAS.
- 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.
- B. K. Burgess and D. J. Lowe, Chem. Rev., 1996, 96, 2983–3012 CrossRef CAS PubMed.
- Y. Pang, C. Su, G. Jia, L. Xu and Z. Shao, Chem. Soc. Rev., 2021, 50, 12744–12787 RSC.
- J. Yang, Y. Guo, R. Jiang, F. Qin, H. Zhang, W. Lu, J. Wang and J. C. Yu, J. Am. Chem. Soc., 2018, 140, 8497–8508 CrossRef CAS PubMed.
- H. Du, C. Yang, W. Pu, L. Zeng and J. Gong, ACS Sustainable Chem. Eng., 2020, 8, 10572–10580 CrossRef CAS.
- 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.
- Z. Geng, Y. Liu, X. Kong, P. Li, K. Li, Z. Liu, J. Du, M. Shu, R. Si and J. Zeng, Adv. Mater., 2018, 30, 1803498 CrossRef.
- X. Wang, W. Wang, M. Qiao, G. Wu, W. Chen, T. Yuan, Q. Xu, M. Chen, Y. Zhang, X. Wang, J. Wang, J. Ge, X. Hong, Y. Li, Y. Wu and Y. Li, Sci. Bull., 2018, 63, 1246–1253 CrossRef CAS PubMed.
- Y. Chen, R. Guo, X. Peng, X. Wang, X. Liu, J. Ren, J. He, L. Zhuo, J. Sun, Y. Liu, Y. Wu and J. Luo, ACS Nano, 2020, 14, 6938–6946 CrossRef CAS.
- L. Han, X. Liu, J. Chen, R. Lin, H. Liu, F. Lue, S. Bak, Z. Liang, S. Zhao, E. Stavitski, J. Luo, R. R. Adzic and H. L. Xin, Angew. Chem., Int. Ed., 2019, 58, 2321–2325 CrossRef CAS PubMed.
- W. Liao, L. Qi, Y. Wang, J. Qin, G. Liu, S. Liang, H. He and L. Jiang, Adv. Funct. Mater., 2021, 31, 2009151 CrossRef CAS.
- J. Geng, S. Zhang, H. Xu, G. Wang and H. Zhang, Chem. Commun., 2021, 57, 5410–5413 RSC.
- Y. Ma, T. Yang, H. Zou, W. Zang, Z. Kou, L. Mao, Y. Feng, L. Shen, S. J. Pennycook, L. Duan, X. Li and J. Wang, Adv. Mater., 2020, 32, 2002177 CrossRef CAS PubMed.
- J. Li, S. Chen, F. Quan, G. Zhan, F. Jia, Z. Ai and L. Zhang, Chem, 2020, 6, 885–901 CAS.
- Y. Zhao, S. Zhang, C. Han, Q. Lu, Q. Fu, H. Jiang, L. Yang, Y. Xing, Q. Zheng, J. Shen, L. Yan and X. Zhao, Chem. Eng. J., 2023, 468, 143517 Search PubMed.
- S. Zhang, M. Jin, T. Shi, M. Han, Q. Sun, Y. Lin, Z. Ding, L. R. Zheng, G. Wang, Y. Zhang, H. Zhang and H. Zhao, Angew. Chem., Int. Ed., 2020, 59, 13423–13429 CrossRef CAS.
- W. Liu, L. Han, H.-T. Wang, X. Zhao, J. A. Boscoboinik, X. Liu, C.-W. Pao, J. Sun, L. Zhuo, J. Luo, J. Ren, W.-F. Pong and H. L. Xin, Nano Energy, 2020, 77, 104990 Search PubMed.
- S. Mukherjee, X. Yang, W. Shan, W. Samarakoon, S. Karakalos, D. A. Cullen, K. More, M. Wang, Z. Feng, G. Wang and G. Wu, Small, Methods, 2020, 4, 1900821 CAS.
- Y. Song, H. Wang, Z. Song, X. Zheng, B. Fan, X. Han, Y. Deng and W. Hu, ACS Appl. Mater. Interfaces, 2022, 14, 17273–17281 CrossRef CAS PubMed.
- Y. Gao, Z. Han, S. Hong, T. Wu, X. Li, J. Qiu and Z. Sun, ACS Appl. Energy Mater., 2019, 2, 6071–6077 CrossRef CAS.
- Y. Wei, X. Wang, M. Sun, M. Ma, J. Tian and M. Shao, Energy Environ. Mater., 2024, 7, e12630 CrossRef CAS.
- L. Wang, Y. Liu, H. Wang, T. Yang, Y. Luo, S. Lee, M. G. Kim, T. T. T. Nga, C.-L. Dong and H. Lee, ACS Nano, 2023, 17, 7406–7416 CrossRef CAS.
- X.-W. Lv, Y. Liu, R. Hao, W. Tian and Z.-Y. Yuan, ACS Appl. Mater. Interfaces, 2020, 12, 17502–17508 Search PubMed.
- C. Chen, Y. Liu and Y. Yao, Eur. J. Inorg. Chem., 2020, 3236–3241 Search PubMed.
- J. Zhao and Z. Chen, J. Am. Chem. Soc., 2017, 139, 12480–12487 Search PubMed.
- R. Yang, Y. Wang, H. Li, J. Zhou, Z. Gao, C. Liu and B. Zhang, Angew. Chem., Int. Ed., 2024, 63, e202317167 CrossRef CAS PubMed.
- W. Zhang, X. Qin, T. Wei, Q. Liu, J. Luo and X. Liu, J. Colloid Interface Sci., 2023, 638, 650–657 Search PubMed.
- J. Zhou, S. Han, R. Yang, T. Li, W. Li, Y. Wang, Y. Yu and B. Zhang, Angew. Chem., Int. Ed., 2023, 62, e202305184 CrossRef CAS PubMed.
- K. Chen, N. Zhang, F. Wang, J. Kang and K. Chu, J. Mater. Chem. A, 2023, 11, 6814–6819 Search PubMed.
- P. Lv, D. Wu, B. He, X. Li, R. Zhu, G. Tang, Z. Lu, D. Ma and Y. Jia, J. Mater. Chem. A, 2022, 10, 9707–9716 RSC.
- L. Yang, J. Fan, B. Xiao and W. Zhu, Chem. Eng. J., 2023, 468, 143823 CrossRef CAS.
- L. Yang, J. Fan and W. Zhu, J. Mater. Chem. A, 2023, 11, 7159–7169 RSC.
- X. Li, K. Chen, X. Lu, D. Ma and K. Chu, Chem. Eng. J., 2023, 454, 140333 Search PubMed.
- K. Chen, Y. Zhang, J. Xiang, X. Zhao, X. Li and K. Chu, ACS Energy Lett., 2023, 8, 1281–1288 CrossRef CAS.
- K. Chen, J. Xiang, Y. Guo, X. Liu, X. Li and K. Chu, Nano Lett., 2024, 24, 541–548 CrossRef CAS.
- R. Zhang, S. Zhang, Y. Guo, C. Li, J. Liu, Z. Huang, Y. Zhao, Y. Li and C. Zhi, Energy Environ. Sci., 2022, 15, 3024–3032 RSC.
- J. Xiang, H. Zhao, K. Chen, X. Li, X. Li and K. Chu, J. Colloid Interface Sci., 2024, 653, 390–395 CrossRef CAS PubMed.
- W. Du, Y. Zhang, K. Chen, G. Zhang and K. Chu, J. Cleaner Prod., 2023, 424, 138875 CrossRef CAS.
- Y. Zhang, H. Zheng, K. Zhou, J. Ye, K. Chu, Z. Zhou, L. Zhang and T. Liu, Adv. Mater., 2023, 35, 2209855 CrossRef CAS PubMed.
- P. Liu, J. Yan, H. Huang and W. Song, Chem. Eng. J., 2023, 466, 143134 CrossRef CAS.
- G.-F. Chen, Y. Yuan, H. Jiang, S.-Y. Ren, L.-X. Ding, L. Ma, T. Wu, J. Lu and H. Wang, Nat. Energy, 2020, 5, 605–613 CrossRef CAS.
- M. Xu, Q. Xie, D. Duan, Y. Zhang, Y. Zhou, H. Zhou, X. Li, Y. Wang, P. Gao and W. Ye, ChemSusChem, 2022, 15, e202200231 CrossRef CAS PubMed.
- H. Yin, F. Dong, Y. Wang, H. Su, X. Li, Y. Peng, H. Duan and J. Li, Nano Lett., 2023, 23, 11899–11906 CrossRef PubMed.
- Y. Wang, H. Yin, F. Dong, X. Zhao, Y. Qu, L. Wang, Y. Peng, D. Wang, W. Fang and J. Li, Small, 2023, 19, 2207695 CrossRef CAS PubMed.
- Y. Xue, Q. Yu, Q. Ma, Y. Chen, C. Zhang, W. Teng, J. Fan and W. X. Zhang, Environ. Sci. Technol., 2022, 56, 14797–14807 CrossRef CAS PubMed.
- Y. Wang, W. Zhang, W. Wen, X. Yu, Y. Du, K. Ni, Y. Zhu and M. Zhu, Adv. Funct. Mater., 2023, 33, 2302651 CrossRef CAS.
- X.-F. Cheng, J.-H. He, H.-Q. Ji, H.-Y. Zhang, Q. Cao, W.-J. Sun, C.-L. Yan and J.-M. Lu, Adv. Mater., 2022, 34, 2205767 CrossRef CAS PubMed.
- J. Yang, H. Qi, A. Li, X. Liu, X. Yang, S. Zhang, Q. Zhao, Q. Jiang, Y. Su, L. Zhang, J.-F. Li, Z.-Q. Tian, W. Liu, A. Wang and T. Zhang, J. Am. Chem. Soc., 2022, 144, 12062–12071 CrossRef CAS PubMed.
- J. Zhang, Z. Hao, Z. Zhang, Y. Yang and X. Xu, Process Saf. Environ. Prot., 2010, 88, 439–445 CrossRef CAS.
- H. Zheng, G. Wisedchaisri and T. Gonen, Nature, 2013, 497, 647–651 CrossRef CAS PubMed.
- Y. Wang, D. Wu, P. Lv, B. He, X. Li, D. Ma and Y. Jia, Nanoscale, 2022, 14, 10862–10872 RSC.
- Z.-Y. Wu, M. Karamad, X. Yong, Q. Huang, D. A. Cullen, P. Zhu, C. Xia, Q. Xiao, M. Shakouri, F.-Y. Chen, J. Y. Kim, Y. Xia, K. Heck, Y. Hu, M. S. Wong, Q. Li, I. Gates, S. Siahrostami and H. Wang, Nat. Commun., 2021, 12, 2870 CrossRef CAS PubMed.
- H. Wang, S. Man, H. Wang, V. Presser, Q. Yan and Y. Zhang, Appl. Catal., B, 2023, 332, 122778 CrossRef CAS.
- E. Murphy, Y. Liu, I. Matanovic, S. Guo, P. Tieu, Y. Huang, A. Ly, S. Das, I. Zenyuk, X. Pan, E. Spoerke and P. Atanassov, ACS Catal., 2022, 12, 6651–6662 CrossRef CAS.
- J. Wan, H. Zhang, J. Yang, J. Zheng, Z. Han, W. Yuan, B. Lan and X. Li, Appl. Catal., B, 2024, 347, 123816 CrossRef CAS.
- F. Liu, J. Li, N. An, J. Huang, X. Liu and M. Li, J. Hazard. Mater., 2024, 465, 133484 CrossRef CAS PubMed.
- W. D. Zhang, H. Dong, L. Zhou, H. Xu, H.-R. Wang, X. Yan, Y. Jiang, J. Zhang and Z.-G. Gu, Appl. Catal., B, 2022, 317, 121750 CrossRef CAS.
- S. Zhang, J. Wu, M. Zheng, X. Jin, Z. Shen, Z. Li, Y. Wang, Q. Wang, X. Wang, H. Wei, J. Zhang, P. Wang, S. Zhang, L. Yu, L. Dong, Q. Zhu, H. Zhang and J. Lu, Nat. Commun., 2023, 14, 3634 CrossRef CAS PubMed.
- R. Zhang, Y. Guo, S. Zhang, D. Chen, Y. Zhao, Z. Huang, L. Ma, P. Li, Q. Yang, G. Liang and C. Zhi, Adv. Energy Mater., 2022, 12, 2103872 CrossRef CAS.
- N. Zhang, G. Zhang, P. Shen, H. Zhang, D. Ma and K. Chu, Adv. Funct. Mater., 2023, 33, 2211537 CrossRef CAS.
- Z. Liu, L. Zeng, J. Yu, L. Yang, J. Zhang, X. Zhang, F. Han, L. Zhao, X. Li, H. Liu and W. Zhou, Nano Energy, 2021, 85, 105940 CrossRef CAS.
- X. Bai, K. Zhou, L. Lin, H. Li, Q. Li and G. Fan, Fuel, 2024, 376, 132746 CrossRef CAS.
- Y. Wang, H. Li, W. Zhou, X. Zhang, B. Zhang and Y. Yu, Angew. Chem., Int. Ed., 2022, 61, e202202604 CrossRef CAS PubMed.
- Y. Yao, L. Zhao, J. Dai, J. Wang, C. Fang, G. Zhan, Q. Zheng, W. Hou and L. Zhang, Angew. Chem., Int. Ed., 2022, 61, e202208215 CrossRef CAS PubMed.
- L. Lv, Y. Shen, J. Liu, X. Meng, X. Gao, M. Zhou, Y. Zhang, D. Gong, Y. Zheng and Z. Zhou, J. Phys. Chem. Lett., 2021, 12, 11143–11150 Search PubMed.
- Z. Lu, R. Yang, Y. Yu, Y. Wang, B. Zhang and L. Kong, Chem. Catal., 2025, 5, 101152 CrossRef.
- G. Zhang, F. Wang, K. Chen, J. Kang and K. Chu, Adv. Funct. Mater., 2024, 34, 2305372 CrossRef CAS.
- Y.-C. Hao, Y. Guo, L.-W. Chen, M. Shu, X.-Y. Wang, T.-A. Bu, W.-Y. Gao, N. Zhang, X. Su, X. Feng, J.-W. Zhou, B. Wang, C.-W. Hu, A.-X. Yin, R. Si, Y.-W. Zhang and C.-H. Yan, Nat. Catal., 2019, 2, 448–456 Search PubMed.
- R. Zhao, Q. Yan, L. Yu, T. Yan, X. Zhu, Z. Zhao, L. Liu and J. Xi, Adv. Mater., 2023, 35, 2306633 CrossRef CAS PubMed.
- H. Lin, J. Wei, Y. Guo, Y. Li, X. Lu, C. Zhou, S. Liu and Y.-Y. Li, Adv. Funct. Mater., 2024, 34, 2409696 CrossRef CAS.
- W. Zhang, S. Zhan, J. Xiao, T. Petit, C. Schlesiger, M. Mellin, J. P. Hofmann, T. Heil, R. Mueller, K. Leopold and M. Oschatz, Adv. Sci., 2023, 10, 2302623 CrossRef CAS PubMed.
- L. Li, C. Tang, H. Jin, K. Davey and S.-Z. Qiao, Chem, 2021, 7, 3232–3255 CAS.
- Y. Sun, Y. Wang, H. Li, W. Zhang, X.-M. Song, D.-M. Feng, X. Sun, B. Jia, H. Mao and T. Ma, J. Electrochem. Chem., 2021, 62, 51–70 CAS.
- K. Chen, Z. Ma, X. Li, J. Kang, D. Ma and K. Chu, Adv. Funct. Mater., 2023, 33, 2209890 CrossRef CAS.
- G. Zhang, G. Wang, Y. Wan, X. Liu and K. Chu, ACS Nano, 2023, 17, 21328–21336 CrossRef PubMed.
- J. Ni, J. Yan, F. Li, H. Qi, Q. Xu, C. Su, L. Sun, H. Sun, J. Ding and B. Liu, Adv. Energy Mater., 2024, 14, 2400065 CrossRef CAS.
- X. Zheng, J. Hao, Z. Zhuang, Q. Kang, X. Wang, S. Lu, F. Duan, M. Du and H. Zhu, Nanoscale, 2024, 16, 4047–4055 RSC.
- J. Li, M. Li, N. An, S. Zhang, Q. Song, Y. Yang, J. Li and X. Liu, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2123450119 CrossRef CAS PubMed.
- Z. Shen, Y. Yu, Z. Zhao, M. A. Mushtaq, Q. Ji, G. Yasin, L. N. U. Rehman, X. Liu, X. Cai, P. Tsiakaras and J. Zhao, Appl. Catal., B, 2023, 331, 122687 Search PubMed.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.