Next Article in Journal
Performance Validation of Resonant Wave Power Converter with Variable Moment of Inertia
Previous Article in Journal
Use of Balanced Scorecard (BSC) Performance Indicators for Small-Scale Hydropower Project Attractiveness Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancing Hydrogen Peroxide Synthesis through Coordination Engineering of Single-Atom Catalysts in the Oxygen Reduction Reaction: A Review

Hydrogen Energy Technology Innovation Center of Hubei Province, Faculty of Materials Science and Chemistry, China University of Geosciences, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(18), 6616; https://doi.org/10.3390/en16186616
Submission received: 9 August 2023 / Revised: 29 August 2023 / Accepted: 11 September 2023 / Published: 14 September 2023
(This article belongs to the Section A5: Hydrogen Energy)

Abstract

:
Hydrogen peroxide (H2O2) is an important chemical with a diverse array of applications. However, the existing scenario of centralized high-concentration production is in contrast with the demand for low-concentration decentralized production. In this context, the on-site green and efficient two-electron oxygen reduction reaction (ORR) for H2O2 production has developed into a promising synthetic approach. The development of low-cost, highly active, and durable advanced catalysts is the core requirement for realizing this approach. In recent years, single-atom catalysts (SACs) have become a research hotspot owing to their maximum atom utilization efficiency, tunable electronic structure, and exceptional catalytic performance. The coordination engineering of SACs is one of the key strategies to unlock their full potential for electrocatalytic H2O2 synthesis and holds significant research value. Despite considerable efforts, precisely controlling the electronic structure of active sites in SACs remains challenging. Therefore, this review summarizes the latest progress in coordination engineering strategies for SACs, aiming to elucidate the relevance between structure and performance. Our goal is to provide valuable guidance and insights to aid in the design and development of high-performance SACs for electrocatalytic H2O2 synthesis.

1. Introduction

Hydrogen peroxide (H2O2) is a vital basic chemical widely used in various fields, including chemical synthesis, papermaking, textiles, healthcare, and environmental protection. It also holds potential as an energy carrier [1,2,3,4]. At present, industrial H2O2 production primarily depends on the anthraquinone process [5]. However, this method is not only complex and energy demanding, but also generates a substantial amount of organic waste [6,7,8]. Additionally, to reduce storage and transportation costs, industrially produced H2O2 is concentrated through distillation to high concentrations, posing potential explosion risks. Nevertheless, many real-world applications only require diluted H2O2. The contradiction between the centralized production of high concentrations and the demand for the dispersed production of low concentrations presents a challenging dilemma [9,10,11].
To address the above-mentioned challenges, a cleaner and safer synthesis approach for H2O2 is via the direct electrocatalytic oxygen reduction reaction (2e ORR) using clean electrical energy and O2 (O2 + 2e + 2H+ → H2O2) [12,13,14,15]. This method offers a more environmentally friendly substitute to the conventional anthraquinone process, especially well-suited for on-site applications that require small amounts of low-concentration H2O2, such as water purification and clean disinfection [16,17,18]. However, the 2e ORR process suffers from slow reaction kinetics and competition with the 4e ORR process that leads to the generation of H2O. This competition results in low activity and selectivity, becoming the two main obstacles in electrocatalytic H2O2 production [19,20,21]. Hence, the design and development of advanced catalysts boasting high activity, selectivity, and long-term stability for the 2e ORR have become crucial issues that urgently need to be addressed in electrocatalytic H2O2 synthesis.
Various catalysts have been employed for the electrocatalytic synthesis of H2O2 to date, including noble metals and their alloys, non-noble metals and their oxides, carbides, non-metallic carbon-based catalysts, and single-atom catalysts [22,23,24,25,26,27,28]. Among the numerous catalysts reported, single-atom catalysts (SACs) with isolated metal atoms anchored on the surface of the support have combined the advantages of both homogeneous and heterogeneous catalysts, offering ultra-high atom utilization, high activity, and high selectivity. As a result, they have emerged as the ideal catalysts for the electrocatalytic H2O2 synthesis [29,30,31]. Particularly, carbon-based SACs have attracted significant interest because of their low cost, wide availability, and relatively simple and tunable active site structures (M-Lx-C, metal–ligand–carbon). These properties make them an excellent platform for studying the mechanism and optimizing their performance of electrocatalytic H2O2 synthesis [32,33,34]. Therefore, carbon-based SACs are the primary focus of our subsequent discussion.
The coordination engineering of carbon-based SACs, as a key strategy to regulate the electronic structure of active sites and optimize ORR activity and selectivity, holds great research significance [35]. Despite considerable efforts in recent times, a comprehensive understanding of the structure–performance relationship in carbon-based SACs is still lacking, making their rational design and synthesis challenging [33,35,36,37]. Therefore, this review summarizes the latest achievements in the coordination engineering of SACs for 2e ORR, aiming to expand our understanding of the structure–performance correlation and provide valuable insights into the development of advanced SACs for electrocatalytic H2O2 synthesis.

2. Mechanism of Electrocatalytic Oxygen Reduction for H2O2 Synthesis

The electrocatalytic oxygen reduction reaction (ORR) is an intricate procedure encompassing multiple proton-coupled electron transfer (PCET) steps. In practical reaction systems, both the 4e ORR and 2e ORR pathways coexist and compete with each other (Figure 1a) [38,39]. As an essential reaction in the processes of the storage and conversion of renewable energy, 4e ORR that thermodynamically favors the generation of H2O is mainly used in proton exchange membrane fuel cells (PEMFCs) [40,41] and metal–air batteries [42,43,44,45]. In contrast, the 2e ORR pathway leading to the formation of H2O2 can serve as a hopeful substitute to address the limitations of the anthraquinone process and enable the on-site decentralized production of H2O2 [46,47,48].
The 4e ORR involves three reaction intermediates, *OOH, *O, and *OH (* denotes the catalytic active sites; *OOH represents the state of the OOH intermediate adsorbed at the active sites, they convey the same meaning as those stated here in the following context), and all reaction steps in the acidic media are as follows:
+ O 2 ( g ) + H + + e   * O O H
  * O O H + H + + e   * O + H 2 O ( l )
  * O + H + + e   * O H
  * O H + H + + e + H 2 O ( l )
In contrast, the 2e ORR pathway involves only one reaction intermediate, *OOH, and its reaction steps are as follows:
+ O 2 ( g ) + H + + e   * O O H
  * O O H + H + + e + H 2 O 2 ( l )
The first step of ORR is the adsorption and activation of O2 molecules on the catalyst surface, and its adsorption configuration significantly influences the reaction tendency [49]. There are three typical adsorption configurations for O2 molecules on the catalyst surface (Figure 1b) [50]: Pauling-type adsorption (end-on adsorption on a single atom), Griffith-type adsorption (side-on adsorption on a single atom), and Bridge-type adsorption (side-on adsorption on two atoms). For metal catalysts, O2 molecules tend to adsorb in a side-on manner on surface atoms, causing a weakened O-O bond, which is more prone to breaking, favoring the 4e ORR pathway for generating H2O. On the contrary, SACs have isolated metal atoms as catalytic centers, which preferentially adsorb O2 molecules in an end-on manner, favoring the retention of the O-O bond and generating H2O2 through the 2e ORR pathway [39,51].
In addition to the adsorption configuration, according to the Sabatier principle [52], an ideal 2e ORR catalyst ought to maintain a moderate binding strength with the *OOH intermediate to achieve a balance between adsorption and desorption, resulting in a volcano-shaped relationship (Figure 1c) [50,53]. The optimal catalyst located at the peak of the volcano curve would have both a low-reaction energy barriers, indicating high activity, and the ability to promptly desorb the generated *OOH, ensuring good selectivity. Furthermore, the binding strength of SACs with the intermediate *OOH is largely determined by the electronic structure of the active site. Therefore, it is imperative to apply appropriate strategies for regulating the electronic structure of SACs.
Figure 1. (a) The pathways of ORR illustrated on carbon-based SACs. Reproduced with permission from [39], ACS, 2021. (b) The three typical adsorption structures of O2 molecules on the catalyst surface; (c) volcano plot constructed according to the Sabatier principle. Reproduced with permission from [50], RSC, 2020.
Figure 1. (a) The pathways of ORR illustrated on carbon-based SACs. Reproduced with permission from [39], ACS, 2021. (b) The three typical adsorption structures of O2 molecules on the catalyst surface; (c) volcano plot constructed according to the Sabatier principle. Reproduced with permission from [50], RSC, 2020.
Energies 16 06616 g001

3. Regulation of Coordination Environment

In carbon-based SACs, the isolated metal atoms that act as active centers are dispersed on the carbon support through coordination with non-metallic atoms. Compared to metal surfaces, the spatial separation of metal atoms in SACs on proper supports leads to their unique electronic structure [54]. As the research on carbon-based SACs expands, an increasing number of studies indicate that regulating the electronic distribution of the central metal atom’s frontier d-orbitals is crucial for achieving ideal catalytic activity [53,55,56]. Since p-block elements are commonly used as coordinating atoms in the first coordination layer, the d-p orbital hybridization dominates the electronic structure of most reported active sites in carbon-based SACs [57,58,59,60]. Therefore, changes in the coordinating atoms within the first coordination layer directly impacts the electronic structure of the active sites, thereby affecting the activity and selectivity of ORRs. In addition, when heteroatoms are doped into the carbon support, long-range interactions can also occur in the second or further coordination layers (Figure 2) [35,61,62]. Thus, we first discussed the direct coordination environments in the following sections, including strategies to adjust coordinating atoms, coordination number, and axial coordination. Secondly, we reviewed the modulation of indirect coordination environments. Finally, the development prospects and challenges of carbon-based SACs’ coordination engineering were summarized.

3.1. Adjustment of the First Coordination Layer

In carbon-based SACs, the first-coordination-layer atoms anchor the metal center to the carbon substrate. Consequently, the characteristics of the atoms in the first coordination layer exert a direct impact on the geometric and electronic configurations of the active sites [63,64,65]. Different coordination atoms often exhibit variations in their atomic radius, electronegativity, and electronic structure, among other properties. To achieve a better electrocatalytic performance for H2O2 synthesis, it is a straightforward and sensible strategy to choose appropriate first-coordination-layer atoms for specific metal centers, modify their electronic structure, and optimize their interactions with reaction intermediates.

3.1.1. The Types of Coordination Atoms

The most common coordinating atoms for anchoring metal centers in carbon-based SACs are nitrogen (N) atoms, which have similar atomic radii to carbon, but higher electronegativity. Therefore, the extensive research and documentation have focused on catalysts featuring M-N4-type sites [66,67]. However, even with the same N element, different types of N atoms (pyridinic N/pyrrolic N) can lead to discrepancies in the catalytic performance. Among an array of transition metal-based carbon SACs used for ORR, Co-SACs have been a controversial presence. Numerous experimental and computational evidence has indicated that Co-SACs exhibit the best d-band center and *OOH intermediate adsorption energy. Therefore, it is an excellent catalyst for the electrochemical synthesis of H2O2 [68,69,70]. However, reports of CoN4 sites showing high activity for 4e ORR are not uncommon [71,72,73]. To unravel this mystery, Chen et al. revealed the “structure–performance” relationship of CoN4 catalysts with different N coordinations through theoretical simulations combined with experiments: pyrrolic-type CoN4 centers predominantly drive the 2e ORR, whereas pyridinic-type CoN4 catalyzes 4e ORR (Figure 3a) [74]. The difference in selectivity is attributed to the fully paired d-orbitals in pyridinic-type CoN4, while pyrrolic-type CoN4 exhibits high-spin states with unpaired electrons in the d-orbitals, resulting in weaker electron transfer with *OOH intermediates, making it more inclined towards 2e ORR. Ultimately, Co-N SACDp with all-pyrrolic N coordination demonstrated outstanding H2O2 selectivity of up to 94% and an excellent mass activity of 14.4 A gcatalyst−1 at 0.5 V versus RHE in 0.1 M HClO4.
Nevertheless, stable M-N4-C structures do not always show the best activity and selectivity for the electrochemical synthesis of H2O2 [75]. Selecting suitable coordinating atoms to optimize the electronic structure of the metal center is crucial for breaking the deadlock. Li and colleagues designed and prepared carbon-based SACs with a biomimetic FeN3S1 asymmetric catalytic center (Figure 3b–d) [76]. The newly designed Fe-NS/C catalyst displayed an excellent selectivity of ~90% for synthesizing H2O2. At a high current density of 100 mA cm−2, it achieved an accumulation concentration of 5.8 wt.% H2O2, enabling medical-grade H2O2 disinfectant electro-synthesis. Experiments along with DFT calculations amply confirmed that the remarkable activity was attributed to the introduction of S species in the FeN3S1 site, which broke the symmetry of the FeN4 structure, accelerated proton transfer, promoted the formation of *OOH intermediates, and facilitated product desorption.
Zhao and colleagues investigated an array of carbon-based Pt-SACs featuring different adjacent coordinating atoms [77]. The results showed that, compared to Pt-C4 and Pt-N4 sites, the Pt-S4 site with the S coordination exhibited an excellent performance in reducing O2 to H2O2, owing to the optimized electronic structure that provided proper binding strength to the intermediates.
Figure 3. (a) The differences in ORR selectivity resulting from the distinct 3d electron configurations after the adsorption of *OOH intermediates on pyrrolic and pyridinic CoN4 coordination structures. (b) The variation in electron density of adsorbed *O2 intermediates on FeN4 and FeN3S1 sites; (c) calculated free-energy diagrams for water dissociation process on FeN4 and FeN3S1 sites; (d) mechanism of electrocatalytic H2O2 synthesis on Fe-NS/C. Reproduced with permission from [76], Wiley-VCH, 2023. (e) Calculated DOS and d-band center of ZnO3C and ZnN4; (f) RRDE voltammograms and (g) calculated H2O2 selectivity of ZnO3C and ZnN4. Reproduced with permission from [36], Wiley-VCH, 2021.
Figure 3. (a) The differences in ORR selectivity resulting from the distinct 3d electron configurations after the adsorption of *OOH intermediates on pyrrolic and pyridinic CoN4 coordination structures. (b) The variation in electron density of adsorbed *O2 intermediates on FeN4 and FeN3S1 sites; (c) calculated free-energy diagrams for water dissociation process on FeN4 and FeN3S1 sites; (d) mechanism of electrocatalytic H2O2 synthesis on Fe-NS/C. Reproduced with permission from [76], Wiley-VCH, 2023. (e) Calculated DOS and d-band center of ZnO3C and ZnN4; (f) RRDE voltammograms and (g) calculated H2O2 selectivity of ZnO3C and ZnN4. Reproduced with permission from [36], Wiley-VCH, 2021.
Energies 16 06616 g003
Apart from N atoms, O atoms, which have higher electronegativity than N, can also serve as excellent coordinating atoms [78,79]. Jia et al. tuned the coordinating environment of Zn single atoms through the alteration of functional groups in the MOF precursor, leading to the synthesis of a unique ZnO3C catalyst with O and C coordinations (Figure 3e–g) [36]. Compared to the traditional Zn-N4 structure, this catalyst achieved near-zero overpotential and high selectivity for 2e ORR in 0.1 M KOH. DFT calculations revealed that the outstanding 2e ORR efficacy of ZnO3C was owing to the greater electronegativity of the O atoms, which diluted the electron distribution around the central Zn atom, lowered its d-band center, and thus weakened its adsorption of *OOH intermediates.

3.1.2. The Number of Coordination Atoms

Compared to the investigation of changing coordination atom types, the impact of altering the number of coordination atoms on 2e ORR for H2O2 synthesis has been less studied. This can be attributed to the instability of low or high coordination structures compared to the classical M-L4 configuration, resulting in limited synthetic pathways [80]. From the perspective of coordination chemistry, changes in the number of coordinating atoms significantly affect the oxidation state of the metal center, altering the geometric structure and electronic configuration of the d-orbitals, finally influencing the coupling with the 2p-orbitals of the intermediate O atoms [63]. Therefore, adjusting the number of coordinating atoms around the metal center has become a viable approach for optimizing the 2e ORR performance.
Zhang and co-workers constructed Co-O-C active sites on the surface of graphene oxide (GO) through a two-step acid etching method (Figure 4a,b) [81]. This catalyst displayed an onset potential of 0.91 V versus. RHE and an exceptional H2O2 selectivity of 81.4% in 0.1 M KOH. Through DFT calculations, they confirmed that the low-coordination Co-O3-C was the actual catalytic center. Compared to the Co-O4-C site, where the Co atom had a large bandgap at the Fermi level, resulting in electron transfer inhibition, the Co-O3-C structure had a smaller bandgap, promoting electron transfer between the catalytic center and intermediates, leading to its outstanding 2e ORR efficiency. Gong et al. synthesized a Co-N-C catalyst with low-coordination CoN2 sites and abundant epoxy groups by a one-step microwave thermal shock technique, achieving the modulation of the coordination number of atomically dispersed Co sites and neighboring oxygen functional groups (Figure 4c,d) [82]. This catalyst exhibited high selectivity (91.3%) for H2O2 synthesis, along with a high mass activity (44.4 A g−1) and high kinetic current density (11.3 mA cm−2) at 0.65 V versus RHE. The stability of this low-coordinated catalyst, which required particular attention, was verified through a 10 h chronoamperometric test with no significant decline in the current and selectivity. Subsequent DFT calculations explained that the excellent activity and selectivity of Co-N2-C/HO originated from its distinctive electronic structure, which was synergistically formed by the low-coordination configuration and the existence of oxygen functional groups.

3.1.3. Axial Coordination

As is well-known, the metal centers within M-N4 sites are in an unsaturated coordination state. With the d-orbitals not being fully occupied, there is an opportunity for an interaction with external ligands along the axial direction [35,83,84]. Therefore, in addition to adjusting the in-plane coordination atoms and numbers in the first coordination sphere, controlling axial coordination may become a novel avenue for advancing the future of carbon-based SACs. Therefore, it is necessary to discuss this aspect separately.
Zhao and colleagues reported Co-SAC with an axial N coordination (Figure 5a–c) [85]. SACs with CoN5 sites achieved a substantial H2O2 yield of 3.4 mol gcatalyst−1h−1 in the flow electrolysis cell with an electrolyte of a 0.5 M NaCl solution at 50 mA cm−2 for a long-term stability period of 24 h. Experiments and calculations showed that the Co-N5 structure was a more favorable coordination type compared to Co-N4, as it not only optimized the projected density of Co 3d-orbitals but also exhibited faster reaction kinetics in the 2e ORR pathway.
Similarly, Fan et al. designed a hybrid SAC(Co-N5) by connecting cobalt phthalocyanine molecules with pyridine-functionalized carbon nanotubes through pyridine ligands [86]. The introduction of pyridine ligands disrupted the typical planar symmetry and symmetric charge distribution of Co-N4 sites. The lone-pair electrons in pyridinic N caused a repulsion concerning the electrons in the Co-3dz2 orbital when forming coordination bonds with CoPc. This induced electron accumulation above the Co center, enhanced the interaction between Co-3dz2 and *OOH’s p orbitals, and thus reduced ∆G*OOH, resulting in an onset potential of 0.85 V versus RHE and a high selectivity of 95%. Furthermore, the team extended the study to demonstrate tunable activity and selectivity through the introduction of electron-donating or electron-withdrawing groups on the pyridine ligands, showcasing the versatility of controlling the activity and selectivity through the axial ligand coordination.
Moreover, small-molecule ligands and halide ions exhibit robust coordination capabilities in the M-N4 sites and show a critical impact on enhancing the activity [87]. The preparation of SACs with additional coordination can be achieved through low-temperature thermal treatment. Kim and co-workers reported the reversible control of SACs’ coordination structures through ligand exchange reactions to modify the reaction activity and selectivity of ORR (Figure 5d) [88]. By thermally treating carbon carriers impregnated with RhCl3 in CO or an NH3 gas flow at 200 °C, they obtained atomically dispersed Rh catalysts with CO or NH2 coordinations. The ORR activity of CO-coordinated Rh-SACs is roughly 30-times greater than that of NH2-coordinated Rh-SACs, while the H2O2 selectivity of NH2-coordinated Rh-SACs is about three-times higher than the former. The key innovation of this study lies in the reversible exchange of CO and NH2 ligands and the change in Rh’s oxidation state through different thermal treatment atmospheres. This achieved the reversible adjustment of ORR activity and selectivity, demonstrating the essential function of small-molecule coordination in regulating the ORR performance of SACs.
Cao et al. used DFT calculations with *OOH binding energy as a descriptor to screen transition metal SACs for 2e ORR and guide the catalyst synthesis. They synthesized a Co-phthalocyanine catalyst (CoPc-OCNT) with axial oxygen coordination in oxidized carbon nanotube carriers [89]. The axial oxygen coordination optimized the *OOH binding energy of CoPc, allowing it to be positioned at the top of the volcano curve. Finally, it exhibited a negligible energy barrier, thus demonstrating an excellent 2e ORR performance. Analogously, Xiao and colleagues synthesized SACs with a super-coordinated Ni1N4O2 structure on carboxylated multi-walled carbon nanotubes (OCNTs) through thermal decomposition (Figure 5e) [90]. Comprehensive structural analysis indicated that two oxygen atoms in OCNTs anchored the NiN4 sites, which facilitated the detachment of *OOH and accelerated the reaction kinetics of 2e ORR. This ultimately resulted in a high H2O2 electro-synthesis rate of 5.7 mmol cm−2 h−1 and a high Faradaic efficiency of 96% at an industrial-relevant current density of 200 mA cm−2.
Figure 5. (a) Free energy diagram of 2e ORR at Co-N5 and Co-N4 sites; (b) projected density of state (PDOS) of the Co 3d-orbital in Co-N5 near the Fermi level (Ef); (c) PDOS of the Co 3d-orbital in Co-N4 near the Ef. Reproduced with permission from [85], RSC, 2021. (d) Schematic diagram depicting the ligand exchange reaction at the active site. Reproduced with permission from [88], Wiley-VCH, 2021. (e) Schematic synthesis process of N4Ni1O2/OCNTs and plausible reaction processes of N4Ni1O2, N4Ni1O1, and N4Ni1 for H2O2 electrosynthesis. Reproduced with permission from [90], Wiley-VCH, 2022.
Figure 5. (a) Free energy diagram of 2e ORR at Co-N5 and Co-N4 sites; (b) projected density of state (PDOS) of the Co 3d-orbital in Co-N5 near the Fermi level (Ef); (c) PDOS of the Co 3d-orbital in Co-N4 near the Ef. Reproduced with permission from [85], RSC, 2021. (d) Schematic diagram depicting the ligand exchange reaction at the active site. Reproduced with permission from [88], Wiley-VCH, 2021. (e) Schematic synthesis process of N4Ni1O2/OCNTs and plausible reaction processes of N4Ni1O2, N4Ni1O1, and N4Ni1 for H2O2 electrosynthesis. Reproduced with permission from [90], Wiley-VCH, 2022.
Energies 16 06616 g005

3.2. Adjustment of the Second and Higher Coordination Layers

In contrast to the direct adjustment of atomic type or coordination number in the first coordination layer, doping non-metallic heteroatoms in the second and higher coordination layers may also facilitate electrocatalytic H2O2 synthesis by modulating the charge distribution around the metal sites through long-range delocalization effects [91]. Liu et al. studied the effects of introducing P heteroatoms into the second coordination layer of CoN4 sites in carbon-based SACs using both experimental and DFT calculations (Figure 6a,b) [92]. They found that the introduction of P atoms to the second coordination layer of the Co center in the form of C-P bonds caused tensile strain on the Co-N bonds in the CoN4 site. This induced a reduction in the electron density of the central Co atom, optimizing the adsorption strength of *OOH on the CoN4-PC active site, ultimately enhancing the efficiency of H2O2 synthesis. The exceptional electrochemical performance, including a high onset potential of 0.81 V, over 97% selectivity, and an H2O2 production rate of 11.2 mol g−1catalyst h−1 during 110 h of durability testing, makes it highly promising for large-scale H2O2 production.
Hyeon and colleagues also highlighted that the catalytic performance of CoN4 sites can be modulated by precisely adjusting the atomic arrangement in the second coordination layer [93]. They calculated the effects of adsorbing electron-poor species (4H*/2H*) and electron-rich species (O*/2O*) neighboring the CoN4 sites on the electronic structure and ∆G*OOH (Figure 6c). The final reports revealed that the adsorption of electron-rich species slightly increased ∆G*OOH due to the electron state correction of the Co atom, thereby favoring the 2e ORR pathway. Likewise, Zhang et al. investigated the electrochemical/chemical treatment to adjust the nearby oxygen functional groups surrounding atomically dispersed Co sites for high-activity and selective H2O2 synthesis under acidic conditions (Figure 6d,e) [94]. The incorporation of neighboring C-O-C functional groups adjusted the electronic structure of the CoN4 portion, significantly weakening the *OOH interaction with the Co atom. Finally, it exhibited the exceptional activity and selectivity of 2e ORR. The overpotential was found to be lower than 0.02 V, comparable to or even lower than the state-of-the-art PtHg4 catalysts. This study highlighted the significance of the synergistic effect between the CoN4 portion and adjacent epoxy functional groups.
Furthermore, Liu et al. utilized β-substituted cobalt porphyrin to adsorb on carbon nanotube substrates to construct multiphase Co-SACs (Figure 6f–h) [95]. Through DFT calculations of different substituted SACs’ d-orbital electronic structures and adsorption energies with intermediates, they found that the electron-withdrawing β-substituted group F significantly weakened the Co-O interaction, providing the best adsorption energy for intermediates and higher activity in generating H2O2. As a reward, the CoPorF/CNT catalyst showed over 94% H2O2 selectivity and a turnover frequency (TOF) of 3.51 s−1. In a dual-electrode electrolytic cell, its rate of H2O2 yield reached 10.76 mol gcatalyst−1 h−1.
Additionally, Yan et al. fabricated SACs of antimony (Sb-NSCF) supported on nitrogen and sulfur co-doped nanofibers using nanostructured Sb2S3 as a template [96]. By doping S atoms in the second coordination layer, they adjusted the oxidation state of Sb atoms, impacting their interactions with *OOH intermediates and promoting the enhanced activity (114.9 A gcatalyst−1 at 0.65 V) and selectivity (97.2%) of Sb-N4 portions for 2e ORR.

4. Conclusions and Outlook

Carbon-based SACs with their unique structures and excellent catalytic performance hold great promise as commercial catalysts for electrochemical H2O2 synthesis. The coordination environment of the metal center is a key factor determining its intrinsic catalytic activity and selectivity. By summarizing the recent efforts in regulating the coordination environment for optimizing 2e ORR’s performance, we described the influence of coordination environments on electronic structures, leading to changes in the reaction tendencies. Various relevant descriptors were proposed to qualitatively and quantitatively correlate the differences in electronic structures with catalytic performance, including adsorption energies, metal center oxidation states, spin states, and others. Selecting a universal descriptor from numerous descriptors that can qualitatively determine various performances will save a significant amount of effort and costs in catalyst developments [97]. This review elucidated the core concept that the coordination engineering of SACs alters the d-orbital structure of the metal center through the suitable manipulation of the coordination environment, which in turn affects the binding strength with *OOH, thereby modulating the activity and selectivity of the electrocatalytic ORR. The exploration and discovery of the correlation between structure and performance demonstrated the immense potential of coordination engineering in SACs.
For future development, the research on the coordination engineering of SACs should be pursued from both experimental and theoretical aspects. On the experimental side, special attention should be directed towards hydrogen peroxide synthesis under acidic conditions, given its higher stability and stronger oxidative nature in such environments, which holds significant importance for market applications. In addition to carbon-based supports, exploring other single-atom catalyst substrates, such as metal-based compounds, may present new avenues for the development of SACs. Moreover, developing universal synthesis strategies for SACs with uniformly distributed active sites and the precise control of the coordination environment remains challenging but essential. For example, leveraging the different affinities of metal atoms to various heteroatoms can help achieve specific coordination structures [98,99].
Characterization is also a crucial cornerstone for studying the coordination engineering of SACs. At present, SACs’ structural identification relies mainly on X-ray absorption fine structure (XAFS) combined with high-resolution electron microscopy. However, this method has its limitations and may not provide precise microscopic structural information. Therefore, efforts should be made to develop new structural characterization techniques. Additionally, active catalytic sites in ORR reactions are not static but dynamically evolving during the reaction. Hence, advanced in situ characterization techniques, such as in situ XAFS and in situ Fourier-transform infrared spectroscopy [100], should be actively studied to track the dynamic catalytic behavior of active sites during the catalytic reaction, revealing reaction mechanisms to provide more insightful guidance for the development of improved catalysts.
On the theoretical calculation side, from interpreting the experimental results from the past to producing increasingly instructive experiments at present, the advantages of theoretical calculations in terms of their simplicity and low cost are becoming more evident. Using DFT calculations combined with high-throughput screening to predict catalytic performance can save a significant amount of trial-and-error experimental costs. Thus, computational tools should continue to be fully utilized to rationally design the coordination environment of SACs for excellent 2e ORR performances in upcoming studies. Nevertheless, it is vital to note that theoretical calculations are conducted under simplified and idealized conditions, and they always present differences compared to the actual complex catalytic reactions. Consequently, efforts should be made to develop new theoretical models, considering solvent effects and electrode potential, thus narrowing the gap between theory and experiment to make more accurate predictions or interpretations of the experimental results.
In conclusion, although we achieved certain accomplishments in the coordination engineering of SACs for electrocatalytic H2O2 synthesis, challenges and opportunities coexist. Innovations in both the experimental and computational approaches hold great potential for the further development of SAC coordination engineering and offer promising prospects for the commercial application of electrocatalytic H2O2 production.

Author Contributions

Writing—original draft preparation, H.H. and J.W.; writing—review and editing, J.S., J.L. and W.C. All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China (No. 22179121), Knowledge Innovation Program of Wuhan-Basic Research (2022010801010202), and Research Fund Program of Guangdong Provincial Key Laboratory of Fuel Cell Technology (FC202201).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dong, K.; Liang, J.; Wang, Y.; Zhang, L.; Xu, Z.; Sun, S.; Luo, Y.; Li, T.; Liu, Q.; Li, N.; et al. Conductive Two-Dimensional Magnesium Metal-Organic Frameworks for High-Efficiency O2 Electroreduction to H2O2. ACS Catal. 2022, 12, 6092–6099. [Google Scholar] [CrossRef]
  2. Guo, J.; Liao, L.; Li, Y.; Liang, J.; Wang, Y.; Ying, D.; Jia, J. Enhanced wastewater treatment via direct electrocatalytic activation of hydrogen peroxide in divided cells with flow-through electrode and bipolar membrane. Sep. Purif. Technol. 2022, 292, 121030. [Google Scholar] [CrossRef]
  3. Fukuzumi, S.; Yamada, Y.; Karlin, K.D. Hydrogen Peroxide as a Sustainable Energy Carrier: Electrocatalytic Production of Hydrogen Peroxide and the Fuel Cell. Electrochim. Acta 2012, 82, 493–511. [Google Scholar] [CrossRef] [PubMed]
  4. Kopacz, W.; Okninski, A.; Kasztankiewicz, A.; Nowakowski, P.; Rarata, G.; Maksimowski, P. Hydrogen peroxide–A promising oxidizer for rocket propulsion and its application in solid rocket propellants. FirePhysChem 2022, 2, 56–66. [Google Scholar] [CrossRef]
  5. Wu, Z.; Wang, T.; Zou, J.-J.; Li, Y.; Zhang, C. Amorphous Nickel Oxides Supported on Carbon Nanosheets as High-Performance Catalysts for Electrochemical Synthesis of Hydrogen Peroxide. ACS Catal. 2022, 12, 5911–5920. [Google Scholar] [CrossRef]
  6. Wang, N.; Ma, S.; Zuo, P.; Duan, J.; Hou, B. Recent Progress of Electrochemical Production of Hydrogen Peroxide by Two-Electron Oxygen Reduction Reaction. Adv. Sci. 2021, 8, e2100076. [Google Scholar] [CrossRef]
  7. Xing, Z.H.; Wang, W.L.; Li, X.Z.; Zhang, K.; Gan, L.; Wu, Q.Y. Electrochemical Synthesis of Hydrogen Peroxide Catalyzed by Carbon Nanotubes with Surface Co-NX Sites and Encapsulated Co Nanoparticles. ACS Appl. Mater. Interfaces 2022, 14, 44282–44291. [Google Scholar] [CrossRef]
  8. Rawah, B.S.; Li, W. Electrocatalytic generation of hydrogen peroxide on cobalt nanoparticles embedded in nitrogen-doped carbon. Chin. J. Catal. 2021, 42, 2296–2305. [Google Scholar] [CrossRef]
  9. Yang, S.; Verdaguer-Casadevall, A.; Arnarson, L.; Silvioli, L.; Čolić, V.; Frydendal, R.; Rossmeisl, J.; Chorkendorff, I.; Stephens, I.E.L. Toward the Decentralized Electrochemical Production of H2O2: A Focus on the Catalysis. ACS Catal. 2018, 8, 4064–4081. [Google Scholar] [CrossRef]
  10. Tian, Z.; Zhang, Q.; Thomsen, L.; Gao, N.; Pan, J.; Daiyan, R.; Yun, J.; Brandt, J.; Lopez-Salas, N.; Lai, F.; et al. Constructing Interfacial Boron-Nitrogen Moieties in Turbostratic Carbon for Electrochemical Hydrogen Peroxide Production. Angew. Chem. Int. Ed. 2022, 61, e202206915. [Google Scholar] [CrossRef]
  11. Li, S.; Ma, J.; Xu, F.; Wei, L.; He, D. Fundamental principles and environmental applications of electrochemical hydrogen peroxide production: A review. Chem. Eng. J. 2023, 452, 139371. [Google Scholar] [CrossRef]
  12. Perry, S.C.; Pangotra, D.; Vieira, L.; Csepei, L.-I.; Sieber, V.; Wang, L.; Ponce de León, C.; Walsh, F.C. Electrochemical synthesis of hydrogen peroxide from water and oxygen. Nat. Rev. Chem. 2019, 3, 442–458. [Google Scholar] [CrossRef]
  13. Zhao, X.; Yin, Q.; Mao, X.; Cheng, C.; Zhang, L.; Wang, L.; Liu, T.F.; Li, Y.; Li, Y. Theory-guided design of hydrogen-bonded cobaltoporphyrin frameworks for highly selective electrochemical H2O2 production in acid. Nat. Commun. 2022, 13, 2721. [Google Scholar] [CrossRef] [PubMed]
  14. Guo, Y.; Zhang, R.; Zhang, S.; Hong, H.; Zhao, Y.; Huang, Z.; Han, C.; Li, H.; Zhi, C. Ultrahigh oxygen-doped carbon quantum dots for highly efficient H2O2 production via two-electron electrochemical oxygen reduction. Energy Environ. Sci. 2022, 15, 4167–4174. [Google Scholar] [CrossRef]
  15. Sabri Rawah, B.; Albloushi, M.; Li, W. Electro-synthesis of pure aqueous H2O2 on nitrogen-doped carbon in a solid electrolyte flow cell without using anion exchange membrane. Chem. Eng. J. 2023, 466, 143282. [Google Scholar] [CrossRef]
  16. Zhang, J.; Zhang, H.; Cheng, M.J.; Lu, Q. Tailoring the Electrochemical Production of H2O2: Strategies for the Rational Design of High-Performance Electrocatalysts. Small 2020, 16, e1902845. [Google Scholar] [CrossRef]
  17. Wang, Y.; Waterhouse, G.I.N.; Shang, L.; Zhang, T. Electrocatalytic Oxygen Reduction to Hydrogen Peroxide: From Homogeneous to Heterogeneous Electrocatalysis. Adv. Energy Mater. 2020, 11, 2003323. [Google Scholar] [CrossRef]
  18. Asghar, A.; Abdul Raman, A.A.; Wan Daud, W.M.A. Advanced oxidation processes for in-situ production of hydrogen peroxide/hydroxyl radical for textile wastewater treatment: A review. J. Clean. Prod. 2015, 87, 826–838. [Google Scholar] [CrossRef]
  19. Xia, Y.; Zhao, X.; Xia, C.; Wu, Z.Y.; Zhu, P.; Kim, J.Y.T.; Bai, X.; Gao, G.; Hu, Y.; Zhong, J.; et al. Highly active and selective oxygen reduction to H2O2 on boron-doped carbon for high production rates. Nat. Commun. 2021, 12, 4225. [Google Scholar] [CrossRef]
  20. Xu, H.; Zhang, S.; Geng, J.; Wang, G.; Zhang, H. Cobalt single atom catalysts for the efficient electrosynthesis of hydrogen peroxide. Inorg. Chem. Front. 2021, 8, 2829–2834. [Google Scholar] [CrossRef]
  21. Zhang, C.; Shen, W.; Guo, K.; Xiong, M.; Zhang, J.; Lu, X. A Pentagonal Defect-Rich Metal-Free Carbon Electrocatalyst for Boosting Acidic O2 Reduction to H2O2 Production. J. Am. Chem. Soc. 2023, 145, 11589–11598. [Google Scholar] [CrossRef] [PubMed]
  22. Jiang, Y.; Ni, P.; Chen, C.; Lu, Y.; Yang, P.; Kong, B.; Fisher, A.; Wang, X. Selective Electrochemical H2O2 Production through Two-Electron Oxygen Electrochemistry. Adv. Energy Mater. 2018, 8, 1801909. [Google Scholar] [CrossRef]
  23. Zhang, J.-Y.; Xia, C.; Wang, H.-F.; Tang, C. Recent advances in electrocatalytic oxygen reduction for on-site hydrogen peroxide synthesis in acidic media. J. Energy Chem. 2022, 67, 432–450. [Google Scholar] [CrossRef]
  24. Sheng, H.; Janes, A.N.; Ross, R.D.; Hofstetter, H.; Lee, K.; Schmidt, J.R.; Jin, S. Linear paired electrochemical valorization of glycerol enabled by the electro-Fenton process using a stable NiSe2 cathode. Nat. Catal. 2022, 5, 716–725. [Google Scholar] [CrossRef]
  25. Sang, Z.-y.; Hou, F.; Wang, S.-h.; Liang, J. Research progress on carbon-based non-metallic nanomaterials as catalysts for the two-electron oxygen reduction for hydrogen peroxide production. New Carbon Mater. 2022, 37, 136–151. [Google Scholar] [CrossRef]
  26. Zhang, W.; Li, J.; Wei, Z. Carbon-based catalysts of the oxygen reduction reaction: Mechanistic understanding and porous structures. Chin. J. Catal. 2023, 48, 15–31. [Google Scholar] [CrossRef]
  27. Jin, M.; Liu, W.; Sun, J.; Wang, X.; Zhang, S.; Luo, J.; Liu, X. Highly dispersed Ag clusters for active and stable hydrogen peroxide production. Nano Res. 2022, 15, 5842–5847. [Google Scholar] [CrossRef]
  28. Pizzutilo, E.; Kasian, O.; Choi, C.H.; Cherevko, S.; Hutchings, G.J.; Mayrhofer, K.J.J.; Freakley, S.J. Electrocatalytic synthesis of hydrogen peroxide on Au-Pd nanoparticles: From fundamentals to continuous production. Chem. Phys. Lett. 2017, 683, 436–442. [Google Scholar] [CrossRef]
  29. Li, X.; Huang, Y.; Liu, B. Catalyst: Single-Atom Catalysis: Directing the Way toward the Nature of Catalysis. Chem 2019, 5, 2733–2735. [Google Scholar] [CrossRef]
  30. Chen, Y.; Ji, S.; Chen, C.; Peng, Q.; Wang, D.; Li, Y. Single-Atom Catalysts: Synthetic Strategies and Electrochemical Applications. Joule 2018, 2, 1242–1264. [Google Scholar] [CrossRef]
  31. Wei, Z.; Deng, B.; Chen, P.; Zhao, T.; Zhao, S. Palladium-based single atom catalysts for high-performance electrochemical production of hydrogen peroxide. Chem. Eng. J. 2022, 428, 131112. [Google Scholar] [CrossRef]
  32. Tang, C.; Jiao, Y.; Shi, B.; Liu, J.N.; Xie, Z.; Chen, X.; Zhang, Q.; Qiao, S.Z. Coordination Tunes Selectivity: Two-Electron Oxygen Reduction on High-Loading Molybdenum Single-Atom Catalysts. Angew. Chem. Int. Ed. 2020, 59, 9171–9176. [Google Scholar] [CrossRef]
  33. Gao, J.; Liu, B. Progress of Electrochemical Hydrogen Peroxide Synthesis over Single Atom Catalysts. ACS Mater. Lett. 2020, 2, 1008–1024. [Google Scholar] [CrossRef]
  34. Li, X.; Tang, S.; Dou, S.; Fan, H.J.; Choksi, T.S.; Wang, X. Molecule Confined Isolated Metal Sites Enable the Electrocatalytic Synthesis of Hydrogen Peroxide. Adv. Mater. 2022, 34, 2104891. [Google Scholar] [CrossRef] [PubMed]
  35. Zhu, P.; Song, P.; Feng, W.; Zhao, D.; Liu, T.; Zhang, J.; Chen, C. Tailoring the selectivity and activity of oxygen reduction by regulating the coordination environments of carbon-supported atomically dispersed metal sites. J. Mater. Chem. A 2022, 10, 17948–17967. [Google Scholar] [CrossRef]
  36. Jia, Y.; Xue, Z.; Yang, J.; Liu, Q.; Xian, J.; Zhong, Y.; Sun, Y.; Zhang, X.; Liu, Q.; Yao, D.; et al. Tailoring the Electronic Structure of an Atomically Dispersed Zinc Electrocatalyst: Coordination Environment Regulation for High Selectivity Oxygen Reduction. Angew. Chem. Int. Ed. 2022, 61, 202110838. [Google Scholar] [CrossRef] [PubMed]
  37. Han, G.; Zhang, X.; Liu, W.; Zhang, Q.; Wang, Z.; Cheng, J.; Yao, T.; Gu, L.; Du, C.; Gao, Y.; et al. Substrate strain tunes operando geometric distortion and oxygen reduction activity of CuN2C2 single-atom sites. Nat. Commun. 2021, 12, 6335. [Google Scholar] [CrossRef] [PubMed]
  38. Chen, J.W.; Zhang, Z.; Yan, H.M.; Xia, G.J.; Cao, H.; Wang, Y.G. Pseudo-adsorption and long-range redox coupling during oxygen reduction reaction on single atom electrocatalyst. Nat. Commun. 2022, 13, 1734. [Google Scholar] [CrossRef]
  39. Zhao, X.; Liu, Y. Origin of Selective Production of Hydrogen Peroxide by Electrochemical Oxygen Reduction. J. Am. Chem. Soc. 2021, 143, 9423–9428. [Google Scholar] [CrossRef]
  40. Kodama, K.; Nagai, T.; Kuwaki, A.; Jinnouchi, R.; Morimoto, Y. Challenges in applying highly active Pt-based nanostructured catalysts for oxygen reduction reactions to fuel cell vehicles. Nat. Nanotechnol. 2021, 16, 140–147. [Google Scholar] [CrossRef]
  41. Yin, S.-H.; Yang, S.-L.; Li, G.; Li, G.; Zhang, B.-W.; Wang, C.-T.; Chen, M.-S.; Liao, H.-G.; Yang, J.; Jiang, Y.-X.; et al. Seizing gaseous Fe2+ to densify O2-accessible Fe–N4 sites for high-performance proton exchange membrane fuel cells. Energy Environ. Sci. 2022, 15, 3033–3040. [Google Scholar] [CrossRef]
  42. Zhu, Y.; Wang, X.; Shi, J.; Gan, L.; Huang, B.; Tao, L.; Wang, S. Neuron-inspired design of hierarchically porous carbon networks embedded with single-iron sites for efficient oxygen reduction. Sci. China Chem. 2022, 65, 1445–1452. [Google Scholar] [CrossRef]
  43. Zhu, Y.; Gan, L.; Shi, J.; Huang, G.; Gao, H.; Tao, L.; Wang, S. Co-CoF2 heterojunctions encapsulated in N, F co-doped porous carbon as bifunctional oxygen electrocatalysts for Zn-air batteries. Chem. Eng. J. 2022, 433, 133541. [Google Scholar] [CrossRef]
  44. Li, L.; Liu, X.; Wang, J.; Liu, R.; Liu, Y.; Wang, C.; Yang, W.; Feng, X.; Wang, B. Atomically dispersed Co in a cross-channel hierarchical carbon-based electrocatalyst for high-performance oxygen reduction in Zn–air batteries. J. Mater. Chem. A 2022, 10, 18723–18729. [Google Scholar] [CrossRef]
  45. Shen, Y.; Zhu, Y.; Sunarso, J.; Guan, D.; Liu, B.; Liu, H.; Zhou, W.; Shao, Z. New Phosphorus-Doped Perovskite Oxide as an Oxygen Reduction Reaction Electrocatalyst in an Alkaline Solution. Chem. Eur. J. 2018, 24, 6950–6957. [Google Scholar] [CrossRef]
  46. Zhang, C.; Yuan, L.; Liu, C.; Li, Z.; Zou, Y.; Zhang, X.; Zhang, Y.; Zhang, Z.; Wei, G.; Yu, C. Crystal Engineering Enables Cobalt-Based Metal-Organic Frameworks as High-Performance Electrocatalysts for H2O2 Production. J. Am. Chem. Soc. 2023, 145, 7791–7799. [Google Scholar] [CrossRef]
  47. Lu, X.; Wang, D.; Wu, K.H.; Guo, X.; Qi, W. Oxygen reduction to hydrogen peroxide on oxidized nanocarbon: Identification and quantification of active sites. J. Colloid Interface Sci. 2020, 573, 376–383. [Google Scholar] [CrossRef]
  48. Li, B.Q.; Zhao, C.X.; Liu, J.N.; Zhang, Q. Electrosynthesis of Hydrogen Peroxide Synergistically Catalyzed by Atomic Co-Nx-C Sites and Oxygen Functional Groups in Noble-Metal-Free Electrocatalysts. Adv. Mater. 2019, 31, 1808173. [Google Scholar] [CrossRef]
  49. Zhang, L.; Jiang, S.; Ma, W.; Zhou, Z. Oxygen reduction reaction on Pt-based electrocatalysts: Four-electron vs. two-electron pathway. Chin. J. Catal. 2022, 43, 1433–1443. [Google Scholar] [CrossRef]
  50. Wang, K.; Huang, J.; Chen, H.; Wang, Y.; Song, S. Recent advances in electrochemical 2e oxygen reduction reaction for on-site hydrogen peroxide production and beyond. Chem. Commun. 2020, 56, 12109–12121. [Google Scholar] [CrossRef]
  51. Yan, X.; Shi, W.-W.; Wang, X.-z. Carbon based electrocatalysts for selective hydrogen peroxide conversion. New Carbon Mater. 2022, 37, 223–236. [Google Scholar] [CrossRef]
  52. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier principle to a predictive theory of transition-metal heterogeneous catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  53. Zhang, M.; Zhang, K.; Ai, X.; Liang, X.; Zhang, Q.; Chen, H.; Zou, X. Theory-guided electrocatalyst engineering: From mechanism analysis to structural design. Chin. J. Catal. 2022, 43, 2987–3018. [Google Scholar] [CrossRef]
  54. Li, X.; Liu, L.; Ren, X.; Gao, J.; Huang, Y.; Liu, B. Microenvironment modulation of single-atom catalysts and their roles in electrochemical energy conversion. Sci. Adv. 2020, 6, eabb6833. [Google Scholar] [CrossRef]
  55. Liu, K.; Fu, J.; Lin, Y.; Luo, T.; Ni, G.; Li, H.; Lin, Z.; Liu, M. Insights into the activity of single-atom Fe-N-C catalysts for oxygen reduction reaction. Nat. Commun. 2022, 13, 2075. [Google Scholar] [CrossRef]
  56. Dai, Y.; Liu, B.; Zhang, Z.; Guo, P.; Liu, C.; Zhang, Y.; Zhao, L.; Wang, Z. Tailoring the d-Orbital Splitting Manner of Single Atomic Sites for Enhanced Oxygen Reduction. Adv. Mater. 2023, 35, e2210757. [Google Scholar] [CrossRef]
  57. Cui, X.; Li, W.; Ryabchuk, P.; Junge, K.; Beller, M. Bridging homogeneous and heterogeneous catalysis by heterogeneous single-metal-site catalysts. Nat. Catal. 2018, 1, 385–397. [Google Scholar] [CrossRef]
  58. Wu, X.; Zhang, H.; Zuo, S.; Dong, J.; Li, Y.; Zhang, J.; Han, Y. Engineering the Coordination Sphere of Isolated Active Sites to Explore the Intrinsic Activity in Single-Atom Catalysts. Nanomicro Lett. 2021, 13, 136. [Google Scholar] [CrossRef]
  59. Zheng, R.; Meng, Q.; Zhang, L.; Ge, J.; Liu, C.; Xing, W.; Xiao, M. Co-based Catalysts for Selective H2O2 Electroproduction via 2-electron Oxygen Reduction Reaction. Chem. Eur. J. 2023, 29, e202203180. [Google Scholar] [CrossRef]
  60. Luo, E.; Chu, Y.; Liu, J.; Shi, Z.; Zhu, S.; Gong, L.; Ge, J.; Choi, C.H.; Liu, C.; Xing, W. Pyrolyzed M–Nx catalysts for oxygen reduction reaction: Progress and prospects. Energy Environ. Sci. 2021, 14, 2158–2185. [Google Scholar] [CrossRef]
  61. Zhang, Z.; Li, H.; Wu, D.; Zhang, L.; Li, J.; Xu, J.; Lin, S.; Datye, A.K.; Xiong, H. Coordination structure at work: Atomically dispersed heterogeneous catalysts. Coord Chem. Rev. 2022, 460, 214469. [Google Scholar] [CrossRef]
  62. Fu, C.; Luo, L.; Yang, L.; Shen, S.; Yan, X.; Yin, J.; Wei, G.; Zhang, J. An In-Depth Theoretical Exploration of Influences of Non-Metal-Elements Doping on the ORR Performance of Co−gN4. ChemCatChem 2021, 13, 2303–2310. [Google Scholar] [CrossRef]
  63. Giulimondi, V.; Mitchell, S.; Perez-Ramirez, J. Challenges and Opportunities in Engineering the Electronic Structure of Single-Atom Catalysts. ACS Catal. 2023, 13, 2981–2997. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, Y.; Yang, J.; Ge, R.; Zhang, J.; Cairney, J.M.; Li, Y.; Zhu, M.; Li, S.; Li, W. The effect of coordination environment on the activity and selectivity of single-atom catalysts. Coord Chem. Rev. 2022, 461, 214493. [Google Scholar] [CrossRef]
  65. Han, Y.; Wang, Y.; Xu, R.; Chen, W.; Zheng, L.; Han, A.; Zhu, Y.; Zhang, J.; Zhang, H.; Luo, J.; et al. Electronic structure engineering to boost oxygen reduction activity by controlling the coordination of the central metal. Energy Environ. Sci. 2018, 11, 2348–2352. [Google Scholar] [CrossRef]
  66. Wang, N.; Zhao, X.; Zhang, R.; Yu, S.; Levell, Z.H.; Wang, C.; Ma, S.; Zou, P.; Han, L.; Qin, J.; et al. Highly Selective Oxygen Reduction to Hydrogen Peroxide on a Carbon-Supported Single-Atom Pd Electrocatalyst. ACS Catal. 2022, 12, 4156–4164. [Google Scholar] [CrossRef]
  67. Li, Z.; Li, Y.; Chen, S.; Zha, Q.; Zhu, M. In-situ monitoring of hydrogen peroxide production at nickel single-atom electrocatalyst. Chem. Eng. J. 2023, 460, 141657. [Google Scholar] [CrossRef]
  68. Gao, J.; Yang, H.B.; Huang, X.; Hung, S.-F.; Cai, W.; Jia, C.; Miao, S.; Chen, H.M.; Yang, X.; Huang, Y.; et al. Enabling Direct H2O2 Production in Acidic Media through Rational Design of Transition Metal Single Atom Catalyst. Chem 2020, 6, 658–674. [Google Scholar] [CrossRef]
  69. Zhang, J.; Liu, W.; He, F.; Song, M.; Huang, X.; Shen, T.; Li, J.; Zhang, C.; Zhang, J.; Wang, D. Highly dispersed Co atoms anchored in porous nitrogen-doped carbon for acidic H2O2 electrosynthesis. Chem. Eng. J. 2022, 438, 141657. [Google Scholar] [CrossRef]
  70. Suk, M.; Chung, M.W.; Han, M.H.; Oh, H.-S.; Choi, C.H. Selective H2O2 production on surface-oxidized metal-nitrogen-carbon electrocatalysts. Catal. Today 2021, 359, 99–105. [Google Scholar] [CrossRef]
  71. He, X.; Chang, L.; Han, P.; Li, K.; Wu, H.; Tang, Y.; Gao, F.; Zhang, Y.; Zhou, A. High-performance Co-N-C catalyst derived from PS@ZIF-8 @ZIF-67 for improved oxygen reduction reaction. Colloids Surf. A Physicochem. Eng. Asp. 2023, 663, 130988. [Google Scholar] [CrossRef]
  72. Shi, S.; Wang, B.; Wang, Y.; Yang, Y.; Zhang, Z.; Xu, Y.; Suo, Y. Structure optimization of ZIF-12-derived Co-N-C for efficient oxygen reduction and oxygen evolution. Fuel 2022, 330, 125516. [Google Scholar] [CrossRef]
  73. Gao, J.; Hu, Y.; Wang, Y.; Lin, X.; Hu, K.; Lin, X.; Xie, G.; Liu, X.; Reddy, K.M.; Yuan, Q.; et al. MOF Structure Engineering to Synthesize Co-N-C Catalyst with Richer Accessible Active Sites for Enhanced Oxygen Reduction. Small 2021, 17, e2104684. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, S.; Luo, T.; Li, X.; Chen, K.; Fu, J.; Liu, K.; Cai, C.; Wang, Q.; Li, H.; Chen, Y.; et al. Identification of the Highly Active Co-N4 Coordination Motif for Selective Oxygen Reduction to Hydrogen Peroxide. J. Am. Chem. Soc. 2022, 144, 14505–14516. [Google Scholar] [CrossRef]
  75. Fan, M.; Cui, J.; Zhang, J.; Wu, J.; Chen, S.; Song, L.; Wang, Z.; Wang, A.; Vajtai, R.; Wu, Y.; et al. The modulating effect of N coordination on single-atom catalysts researched by Pt-Nx-C model through both experimental study and DFT simulation. J. Mater. Sci. Technol. 2021, 91, 160–167. [Google Scholar] [CrossRef]
  76. Li, Y.; Chen, J.; Ji, Y.; Zhao, Z.; Cui, W.; Sang, X.; Cheng, Y.; Yang, B.; Li, Z.; Zhang, Q.; et al. Single-atom Iron Catalyst with Biomimetic Active Center to Accelerate Proton Spillover for Medical-level Electrosynthesis of H2O2 Disinfectant. Angew. Chem. Int. Ed. 2023, 62, e202306491. [Google Scholar] [CrossRef]
  77. Zhao, J.; Fu, C.; Ye, K.; Liang, Z.; Jiang, F.; Shen, S.; Zhao, X.; Ma, L.; Shadike, Z.; Wang, X.; et al. Manipulating the oxygen reduction reaction pathway on Pt-coordinated motifs. Nat. Commun. 2022, 13, 685. [Google Scholar] [CrossRef]
  78. Jing, L.; Tian, Q.; Su, P.; Li, H.; Zheng, Y.; Tang, C.; Liu, J. Mesoporous Co–O–C nanosheets for electrochemical production of hydrogen peroxide in acidic medium. J. Mater. Chem. A 2022, 10, 4068–4075. [Google Scholar] [CrossRef]
  79. Jin, H.; Song, T.; Paik, U.; Qiao, S.-Z. Metastable Two-Dimensional Materials for Electrocatalytic Energy Conversions. Acc. Mater. 2021, 2, 559–573. [Google Scholar] [CrossRef]
  80. Rong, X.; Wang, H.J.; Lu, X.L.; Si, R.; Lu, T.B. Controlled Synthesis of a Vacancy-Defect Single-Atom Catalyst for Boosting CO2 Electroreduction. Angew. Chem. Int. Ed. 2020, 59, 1961–1965. [Google Scholar] [CrossRef]
  81. Zhang, B.W.; Zheng, T.; Wang, Y.X.; Du, Y.; Chu, S.Q.; Xia, Z.; Amal, R.; Dou, S.X.; Dai, L. Highly efficient and selective electrocatalytic hydrogen peroxide production on Co-O-C active centers on graphene oxide. Commun. Chem. 2022, 5, 43. [Google Scholar] [CrossRef] [PubMed]
  82. Gong, H.; Wei, Z.; Gong, Z.; Liu, J.; Ye, G.; Yan, M.; Dong, J.; Allen, C.; Liu, J.; Huang, K.; et al. Low-Coordinated Co-N-C on Oxygenated Graphene for Efficient Electrocatalytic H2O2 Production. Adv. Funct. Mater. 2021, 32, 2106886. [Google Scholar] [CrossRef]
  83. Wang, X.; An, Y.; Liu, L.; Fang, L.; Liu, Y.; Zhang, J.; Qi, H.; Heine, T.; Li, T.; Kuc, A.; et al. Atomically Dispersed Pentacoordinated-Zirconium Catalyst with Axial Oxygen Ligand for Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2022, 61, e202209746. [Google Scholar] [CrossRef] [PubMed]
  84. Li, L.; Huang, S.; Cao, R.; Yuan, K.; Lu, C.; Huang, B.; Tang, X.; Hu, T.; Zhuang, X.; Chen, Y. Optimizing Microenvironment of Asymmetric N,S-Coordinated Single-Atom Fe via Axial Fifth Coordination toward Efficient Oxygen Electroreduction. Small 2022, 18, e2105387. [Google Scholar] [CrossRef]
  85. Zhao, Q.; Wang, Y.; Lai, W.-H.; Xiao, F.; Lyu, Y.; Liao, C.; Shao, M. Approaching a high-rate and sustainable production of hydrogen peroxide: Oxygen reduction on Co–N–C single-atom electrocatalysts in simulated seawater. Energy Environ. Sci. 2021, 14, 5444–5456. [Google Scholar] [CrossRef]
  86. Fan, W.; Duan, Z.; Liu, W.; Mehmood, R.; Qu, J.; Cao, Y.; Guo, X.; Zhong, J.; Zhang, F. Rational design of heterogenized molecular phthalocyanine hybrid single-atom electrocatalyst towards two-electron oxygen reduction. Nat. Commun. 2023, 14, 1426. [Google Scholar] [CrossRef]
  87. Liu, X.; Liu, Y.; Yang, W.; Feng, X.; Wang, B. Controlled Modification of Axial Coordination for Transition-Metal Single-Atom Electrocatalyst. Chem. Eur. J. 2022, 28, e202201471. [Google Scholar] [CrossRef]
  88. Kim, J.H.; Shin, D.; Kim, J.; Lim, J.S.; Paidi, V.K.; Shin, T.J.; Jeong, H.Y.; Lee, K.S.; Kim, H.; Joo, S.H. Reversible Ligand Exchange in Atomically Dispersed Catalysts for Modulating the Activity and Selectivity of the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2021, 60, 20528–20534. [Google Scholar] [CrossRef]
  89. Cao, P.; Quan, X.; Nie, X.; Zhao, K.; Liu, Y.; Chen, S.; Yu, H.; Chen, J.G. Metal single-site catalyst design for electrocatalytic production of hydrogen peroxide at industrial-relevant currents. Nat. Commun. 2023, 14, 172. [Google Scholar] [CrossRef]
  90. Xiao, C.; Cheng, L.; Zhu, Y.; Wang, G.; Chen, L.; Wang, Y.; Chen, R.; Li, Y.; Li, C. Super-Coordinated Nickel N4Ni1O2 Site Single-Atom Catalyst for Selective H2O2 Electrosynthesis at High Current Densities. Angew. Chem. Int. Ed. 2022, 61, e202206544. [Google Scholar] [CrossRef]
  91. Tang, C.; Chen, L.; Li, H.; Li, L.; Jiao, Y.; Zheng, Y.; Xu, H.; Davey, K.; Qiao, S.Z. Tailoring Acidic Oxygen Reduction Selectivity on Single-Atom Catalysts via Modification of First and Second Coordination Spheres. J. Am. Chem. Soc. 2021, 143, 7819–7827. [Google Scholar] [CrossRef] [PubMed]
  92. Liu, J.; Wei, Z.; Gong, Z.; Yan, M.; Hu, Y.; Zhao, S.; Ye, G.; Fei, H. Single-atom CoN4 sites with elongated bonding induced by phosphorus doping for efficient H2O2 electrosynthesis. Appl. Catal. B 2023, 324, 122267. [Google Scholar] [CrossRef]
  93. Jung, E.; Shin, H.; Lee, B.H.; Efremov, V.; Lee, S.; Lee, H.S.; Kim, J.; Hooch Antink, W.; Park, S.; Lee, K.S.; et al. Atomic-level tuning of Co-N-C catalyst for high-performance electrochemical H2O2 production. Nat. Mater. 2020, 19, 436–442. [Google Scholar] [CrossRef]
  94. Zhang, Q.; Tan, X.; Bedford, N.M.; Han, Z.; Thomsen, L.; Smith, S.; Amal, R.; Lu, X. Direct insights into the role of epoxy groups on cobalt sites for acidic H2O2 production. Nat. Commun. 2020, 11, 4181. [Google Scholar] [CrossRef]
  95. Liu, C.; Yu, Z.; She, F.; Chen, J.; Liu, F.; Qu, J.; Cairney, J.M.; Wu, C.; Liu, K.; Yang, W.; et al. Heterogeneous molecular Co–N–C catalysts for efficient electrochemical H2O2 synthesis. Energy Environ. Sci. 2023, 16, 446–459. [Google Scholar] [CrossRef]
  96. Yan, M.; Wei, Z.; Gong, Z.; Johannessen, B.; Ye, G.; He, G.; Liu, J.; Zhao, S.; Cui, C.; Fei, H. Sb2S3-templated synthesis of sulfur-doped Sb-N-C with hierarchical architecture and high metal loading for H2O2 electrosynthesis. Nat. Commun. 2023, 14, 368. [Google Scholar] [CrossRef] [PubMed]
  97. Guan, D.; Xu, H.; Zhang, Q.; Huang, Y.C.; Shi, C.; Chang, Y.C.; Xu, X.; Tang, J.; Gu, Y.; Pao, C.W.; et al. Identifying A Universal Activity Descriptor and a Unifying Mechanism Concept on Perovskite Oxides for Green Hydrogen Production. Adv. Mater. 2023, 2305074. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, J.; Zhao, Y.; Chen, C.; Huang, Y.C.; Dong, C.L.; Chen, C.J.; Liu, R.S.; Wang, C.; Yan, K.; Li, Y.; et al. Tuning the Coordination Environment in Single-Atom Catalysts to Achieve Highly Efficient Oxygen Reduction Reactions. J. Am. Chem. Soc. 2019, 141, 20118–20126. [Google Scholar] [CrossRef]
  99. Zhang, E.; Tao, L.; An, J.; Zhang, J.; Meng, L.; Zheng, X.; Wang, Y.; Li, N.; Du, S.; Zhang, J.; et al. Engineering the Local Atomic Environments of Indium Single-Atom Catalysts for Efficient Electrochemical Production of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2022, 61, e202117347. [Google Scholar] [CrossRef]
  100. Hu, J.; Shang, W.; Xin, C.; Guo, J.; Cheng, X.; Zhang, S.; Song, S.; Liu, W.; Ju, F.; Hou, J.; et al. Uncovering Dynamic Edge-Sites in Atomic Co-N-C Electrocatalyst for Selective Hydrogen Peroxide Production. Angew. Chem. Int. Ed. 2023, 62, e202304754. [Google Scholar] [CrossRef]
Figure 2. The coordination environment of atomically dispersed metal centers in carbon-based single-atom catalysts (SACs).
Figure 2. The coordination environment of atomically dispersed metal centers in carbon-based single-atom catalysts (SACs).
Energies 16 06616 g002
Figure 4. (a) The process diagram of synthesizing the Co1@GO catalyst; (b) optimal Co-O-C structure selected via DFT calculations, the gray, red, and blue spheres represent C, O, and Co atoms, respectively. Reproduced with permission from [81], Springer Nature, 2022. (c) Schematic diagram of the synthetic pathways for Co-N4-C/LO and Co-N2-C/HO; (d) optimal Co-N2 structure containing 4 epoxy groups and one O solvent species (CoN2H6-4O-O). Reproduced with permission from [82], Wiley-VCH, 2021.
Figure 4. (a) The process diagram of synthesizing the Co1@GO catalyst; (b) optimal Co-O-C structure selected via DFT calculations, the gray, red, and blue spheres represent C, O, and Co atoms, respectively. Reproduced with permission from [81], Springer Nature, 2022. (c) Schematic diagram of the synthetic pathways for Co-N4-C/LO and Co-N2-C/HO; (d) optimal Co-N2 structure containing 4 epoxy groups and one O solvent species (CoN2H6-4O-O). Reproduced with permission from [82], Wiley-VCH, 2021.
Energies 16 06616 g004
Figure 6. (a) Optimized structure of CoN4-PC with two PC4 sites near the CoN4 site (CoN4-P2); (b) ORR reaction pathways and free energy diagrams for 2e and 4e ORRs at the CoN4-P2 sites, the red and green spheres represent O and H atoms, respectively, while the rest remains the same with figure (a). Reproduced with permission from [92], Elsevier, 2023. (c) Differential charge densities of Co-N4/graphene after 4H*, 2H*, O*, or 2O* was adsorbed near the cobalt atom. Yellow and cyan isosurfaces (±0.003 Bohr−3) show the electron gain and loss, respectively. Reproduced with permission from [93], Springer Nature, 2020. (d) The optimized structures of bare CoN4 and CoN4 sites with different epoxy oxygen coverages, the gray, blue, red and orange spheres represent C, N, O and Co atoms, respectively; (e) calculated Sabatier volcano plot for 2e ORR to H2O2 for bare CoN4 and CoN4 sites featuring various epoxy oxygen coverages obtained from DFT simulations. Reproduced with permission from [94], Springer Nature, 2020. (f) Division of two distinct substitution sites in the cobalt porphyrin structure; (g) schematic illustration showing the preparation of HMC for H2O2 electrosynthesis from ORR; (h) structure of the investigated cobalt porphyrin molecules. Reproduced with permission from [95], RSC, 2023.
Figure 6. (a) Optimized structure of CoN4-PC with two PC4 sites near the CoN4 site (CoN4-P2); (b) ORR reaction pathways and free energy diagrams for 2e and 4e ORRs at the CoN4-P2 sites, the red and green spheres represent O and H atoms, respectively, while the rest remains the same with figure (a). Reproduced with permission from [92], Elsevier, 2023. (c) Differential charge densities of Co-N4/graphene after 4H*, 2H*, O*, or 2O* was adsorbed near the cobalt atom. Yellow and cyan isosurfaces (±0.003 Bohr−3) show the electron gain and loss, respectively. Reproduced with permission from [93], Springer Nature, 2020. (d) The optimized structures of bare CoN4 and CoN4 sites with different epoxy oxygen coverages, the gray, blue, red and orange spheres represent C, N, O and Co atoms, respectively; (e) calculated Sabatier volcano plot for 2e ORR to H2O2 for bare CoN4 and CoN4 sites featuring various epoxy oxygen coverages obtained from DFT simulations. Reproduced with permission from [94], Springer Nature, 2020. (f) Division of two distinct substitution sites in the cobalt porphyrin structure; (g) schematic illustration showing the preparation of HMC for H2O2 electrosynthesis from ORR; (h) structure of the investigated cobalt porphyrin molecules. Reproduced with permission from [95], RSC, 2023.
Energies 16 06616 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

He, H.; Wang, J.; Shi, J.; Li, J.; Cai, W. Enhancing Hydrogen Peroxide Synthesis through Coordination Engineering of Single-Atom Catalysts in the Oxygen Reduction Reaction: A Review. Energies 2023, 16, 6616. https://doi.org/10.3390/en16186616

AMA Style

He H, Wang J, Shi J, Li J, Cai W. Enhancing Hydrogen Peroxide Synthesis through Coordination Engineering of Single-Atom Catalysts in the Oxygen Reduction Reaction: A Review. Energies. 2023; 16(18):6616. https://doi.org/10.3390/en16186616

Chicago/Turabian Style

He, Huawei, Jiatang Wang, Jiawei Shi, Jing Li, and Weiwei Cai. 2023. "Enhancing Hydrogen Peroxide Synthesis through Coordination Engineering of Single-Atom Catalysts in the Oxygen Reduction Reaction: A Review" Energies 16, no. 18: 6616. https://doi.org/10.3390/en16186616

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop