Next Article in Journal
Polyoxometalate-Based Photocatalytic New Materials for the Treatment of Water Pollutants: Mechanism, Advances, and Challenges
Previous Article in Journal
Catalysis: Key Technology for the Conversion of CO2 into Fuels and Chemicals
Previous Article in Special Issue
Non-Platinum Group Metal Oxygen Reduction Catalysts for a Hydrogen Fuel Cell Cathode: A Mini-Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxygen Reduction Reactions of Catalysts with Asymmetric Atomic Structures: Mechanisms, Applications, and Challenges

Shandong Provincial Key Laboratory of Chemical Energy Storage and Novel Cell Technology, Advanced Materials Institute, School of Materials Science and Engineering, Qilu University of Technology (Shandong Academy of Sciences), Jinan 250353, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(7), 615; https://doi.org/10.3390/catal15070615
Submission received: 3 June 2025 / Revised: 19 June 2025 / Accepted: 19 June 2025 / Published: 21 June 2025
(This article belongs to the Special Issue Feature Review Papers in Electrocatalysis)

Abstract

Asymmetric-atomic-structure catalysts can modulate the interactions between active sites and intermediates through their unique electronic filling states and asymmetric charge distribution, breaking the linear relationship between adsorption energy and activity, thereby enhancing the catalytic performance of the oxygen reduction reaction (ORR). By introducing heteroelements, vacancies, or clusters into symmetric-atomic-structure catalysts (e.g., M-N4), asymmetric configurations (such as M-Nx, M-Nx-S/B/O, etc.) can be formed. These modifications substantially alter their internal structure, trigger charge redistribution, and create asymmetric sites to reduce reaction energy barriers, effectively regulating the adsorption strength of oxygen intermediates and significantly improving ORR performance. This review systematically summarizes recent advancements in asymmetric-atomic-structure catalysts for ORR, elucidating the intrinsic “structure–performance–application” relationships to provide theoretical guidance for developing high-performance asymmetric atomic catalysts. First, the ORR mechanisms, including the two-electron and four-electron pathways, are introduced. Furthermore, strategies to modulate catalyst selectivity and activity through doping with metallic/nonmetallic elements or introducing defects are discussed. Finally, prospects for asymmetric-atomic-structure catalysts in next-generation energy storage and conversion technologies are outlined, offering novel insights to overcome current ORR performance bottlenecks.

Graphical Abstract

1. Introduction

The depletion of fossil fuels and the associated energy demands and environmental issues have inevitably spurred interest in developing new energy sources and energy storage/conversion technologies, including hydrogen fuel cells, zinc–air batteries, and others [1,2,3,4,5,6,7,8,9]. In these devices, the oxygen reduction reaction (ORR) plays a pivotal role. The ORR, a core reaction in fuel cells and metal–air batteries, can proceed via two distinct pathways in flow cells depending on the number of electrons transferred during electrocatalysis: the two-electron pathway, generating hydrogen peroxide (H2O2) as an intermediate (O2 + 2H+ + 2e → H2O2) [10,11], or the four-electron pathway, producing water (H2O) as the final product (O2 + 4H+ + 4e → 2H2O) [12]. Despite being an exothermic reaction, practical applications of ORR still face significant challenges in energy utilization efficiency [13]. The high bond energy of O-O leads to a high activation energy barrier, resulting in sluggish reaction kinetics [14]; the actual operating potential is lower than the theoretical value, necessitating additional energy input and compromising energy efficiency [15]; the inherent complexity of the two-electron and four-electron transfer pathways severely limits energy conversion efficiency and device longevity [16]. Furthermore, catalyst instability in acidic or alkaline electrolytes, where they are prone to degradation or poisoning [17,18,19], poses additional critical barriers. These challenges collectively hinder the large-scale commercialization of ORR-based technologies. Platinum (Pt) is widely recognized as the most effective and commonly used ORR catalyst [20]. Although nanotechnology has paved the way for advancements in Pt catalysts—such as precise control of size and morphology to create Pt nanowires, exposure of high-index facets in Pt nanoparticles, and alloying strategies to form nanocrystals, nanosheets, and nanocages [21,22]—these approaches have mitigated catalytic activity as a limiting factor for ORR. However, the low atomic utilization efficiency of Pt (where only surface atoms participate in reactions, with less than 20% of internal atoms being utilized) poses significant challenges to the cost and performance of fuel cells [23,24,25]. These issues mean that the high cost and limited durability of Pt catalysts remain major obstacles to their widespread commercial application [26,27]. Consequently, the development of non-precious-metal catalyst materials with cost-effectiveness, abundant raw material availability, high ORR activity, excellent stability, and well-defined selectivity has become an imperative objective. In recent years, non-precious-metal symmetric atomic catalysts (e.g., M-N4 sites with highly symmetric coordination structures) [28,29,30,31] have attracted significant attention due to their cost-effectiveness, high stability, universal synthesis of symmetric configurations, highly ordered active site distribution, and modifiable ligand environments. However, their performance remains constrained by low active site density, insufficient exposure efficiency, imbalanced adsorption strength of intermediates, and metal leaching in acidic media. This has facilitated the emergence of asymmetric atomic engineering strategies as an innovative paradigm for ORR catalyst design [32]. The core principle of asymmetric atomic engineering lies in constructing heterogeneous coordination microenvironments, facilitating precise modulation of the geometric and electronic structures of active centers. This approach breaks the performance limitations of conventional catalysts, fostering synergistic enhancement of activity, selectivity, and stability.
Since the early 21st century, conventional ORR catalysts have predominantly manifested highly symmetric coordination environments [33] (e.g., the planar square symmetry of Fe-N4). The homogeneous electron distribution in symmetric structures facilitates uniform adsorption and activation of O2, while modulation of intermediate adsorption energies helps prevent over-adsorption-induced active site poisoning. However, limitations such as restricted electronic structure tunability, corrosion susceptibility in acidic media, and low conductivity have impeded breakthroughs in catalyst stability, activity, and selectivity. To date, significant theoretical advancements have been achieved in understanding asymmetric-atomic-structure catalysts for the ORR. The core challenge of both 2e and 4eORR pathways lies in balancing the kinetics of intermediate adsorption and desorption [34]. In 2eORR, asymmetric coordination lowers the d-band center position of the active metal atom toward the Fermi level, weakening *OOH adsorption and enabling its direct desorption to produce H2O2. Conversely, for 4eORR, elevating the d-band center strengthens the adsorption of intermediates (*O, *OH), while asymmetric coordination promotes O2 dissociation to achieve complete reduction to H2O. The adsorption configuration of O2 critically dictates pathway selectivity: end-on adsorption preserves the O-O bond (favoring 2eORR), whereas side-on adsorption cleaves it (driving 4eORR) [35]. Asymmetric catalysts optimize O2 adsorption modes by tailoring the electronic structure of metal centers through heterogeneous coordination environments. Synergistic effects between elements further elevate ORR performance [36]. For instance, one element activates O2 via adsorption, while another facilitates the desorption of oxygenated intermediates, ensuring efficient reaction kinetics. Moreover, it has been reported that the choice of carbon supports profoundly impacts catalytic activity [37]. The porous structure of carbon governs reactant/product transport efficiency: hierarchical porous hollow carbon spheres accelerate mass transfer due to their high specific surface area and open channels [38], while graphitic carbon enhances conductivity, reduces charge transfer resistance, and improves ORR kinetics. Additionally, 3D porous carbon supports provide abundant anchoring sites to disperse active centers, maximizing active site exposure [39]. These insights collectively establish a robust framework for designing high-performance asymmetric ORR catalysts.
Although asymmetric atom-structured catalysts have made progress in oxygen reduction reaction, they still have problems like poor structural stability, weak anti-poisoning ability, and high demand for structural control precision. Also, there is a lack of research on the mechanism of asymmetric atom-structured catalysts in two-electron and four-electron pathways (Figure 1) and summaries on activity enhancement. In this review, we outline the mechanism and application of ORR in the two-electron and four-electron pathways of asymmetric-atomic-structure catalysts, and provide feasible strategies for the rational design of asymmetric-atomic-structure catalysts. This paper introduces the oxygen reduction reaction mechanisms of the two-electron pathway and the four-electron pathway. In addition, this review synthesizes the reasons for the improvement of catalyst activity, including reasonably regulating the charge coordination environment, defect initiation, and equilibrium adsorption energy through element doping. Finally, by comprehensively looking forward to the development prospects of asymmetric-atomic-structure catalysts, some design principles of the catalysts are proposed to provide new design paradigms for the subsequent generation of asymmetric-atomic-structure oxygen reduction reaction catalysts.

2. Pathway and Mechanism of Oxygen Reduction Reaction

2.1. Two-Electron Pathway and Mechanism of Oxygen Reduction Reaction

The 2eORR pathway involves the catalytic reduction of molecular oxygen through a 2e transfer process, producing hydrogen peroxide (H2O2) or superoxide radicals (O2) as characteristic products. In this process, the adsorption mode of O2 molecules on the catalyst surface (end-on or side-on) critically determines the reaction pathway. End-on adsorption, where O2 binds to the catalyst surface (e.g., Pt) via a single oxygen atom, may instigate O-O bond cleavage, favoring the 4e pathway [40]. Conversely, side-on adsorption, characterized by parallel alignment of O2 on surfaces (e.g., carbon-based materials), preserves the O-O bond and promotes the 2e pathway [41]. Mechanistically, adsorbed O2 undergoes proton-coupled electron transfer (PCET) to form intermediates such as *OOH (in acidic media) or *O2 (in alkaline media), contingent upon pH. These intermediates subsequently accept a second electron and combine with protons, reducing to H2O2 (acidic) or HO2 (alkaline). Although less energy-efficient than the 4e route (direct H2O formation), the 2e pathway is indispensable in applications like H2O2 synthesis, biomedical systems, and wastewater treatment [42,43,44].
The 2e pathway is fundamentally dictated by the partial reduction of O2 molecules through the acceptance of two electrons, yielding H2O2 as the primary product. Its selectivity hinges critically on the catalyst’s adsorption strength toward the key intermediate (*OOH). Weak adsorption promotes H2O2 desorption, suppressing further progression to the 4e pathway, while strong adsorption drives complete O2 reduction via the 4e route. Thus, refining the adsorption energy of *OOH through heteroatom doping—which modifies the electronic coordination configuration of active sites—is a pivotal strategy to enhance 2e selectivity (Figure 2a–c) [45]. Additionally, introducing heteroatoms into a catalyst’s stable lattice generates electron-rich or electron-deficient regions, creating vacancy defects [46]. In asymmetric atomic structures, such defects alter the local electron density and charge distribution, fine-tuning the catalyst’s O2 adsorption energy and activation capacity. Defects further regulate the adsorption strength of critical intermediates (e.g., *OOH) on the catalyst surface (Figure 2d–g) [47]. An idealized adsorption strength stabilizes intermediates, facilitating subsequent electron transfer steps without premature desorption or over-adsorption that could divert the reaction toward the 4e pathway [48]. When a catalyst contains a single metal element as the active center, incorporating nonmetallic heteroatoms disrupts the symmetric coordination structure. The electronegativity mismatch between the metal and nonmetal induces electronic redistribution or a downshift of electron orbitals at the metal center, enhancing ORR activity [49]. Moreover, nonmetallic doping reduces the metal center’s adsorption strength for oxygen-containing intermediates, promoting their conversion to H2O2 [50].
Introducing nonmetallic elements such as N and O into the catalyst breaks the symmetry of the catalyst structure. This alters orbital hybridization, lowers the d-band center of the central atom, and modifies the electronic structure. Consequently, it weakens the adsorption strength of intermediates along the two-electron pathway, shifting the adsorption strength towards favoring desorption. Furthermore, catalysts featuring bimetallic asymmetric coordination promote the two-electron pathway due to their unique side-on adsorption configuration. Additionally, defect-reconstructed catalysts, such as those with dissolved cation vacancies, alter the oxidation state of the central atom. This optimizes the d-band, lowers the adsorption energy of intermediates, and steers the reaction towards the two-electron pathway.

2.2. Four-Electron Pathway and Mechanism of Oxygen Reduction Reaction

The 4eORR pathway is pivotal for attaining high-efficiency energy conversion, with water (H2O, acidic media) or hydroxide ions (OH, alkaline media) as final products, offering high energy efficiency and environmental compatibility. A critical step in the 4e pathway is the cleavage of the O-O bond in molecular oxygen, which typically serves as the rate-determining step (RDS).
The 4eORR pathway revolves around the complete cleavage of the O-O bond in molecular oxygen (O2), achieving full reduction to H2O [51]. In this process, O2 molecules adsorb onto the catalyst surface via chemisorption, where the O-O bond is stretched as electrons transfer from the catalyst to O2’s antibonding *π orbitals, leading to direct dissociation into two O atoms or stepwise reduction via PCET to form *OOH. These oxygen atoms sequentially accept electrons and protons, ultimately generating H2O (in acidic media) or OH (in alkaline media). The complete reduction of O2 through 4e transfer maximizes energy release, offering superior energy efficiency. The selectivity of the 4eORR pathway is governed by the adsorption strength of the *OOH intermediate, which dictates reaction pathway bifurcation. The electronic structure and charge distribution of the catalyst modulate the active site’s interaction with intermediates, making heteroatom doping a key strategy to tailor the local coordination environment of the central active site and augment 4e selectivity (Figure 3a–e) [52]. Additionally, defect engineering introduces vacancies near active sites, disrupting the periodic atomic arrangement and altering the central atom’s electronic distribution to favor the 4e pathway. Finally, embedding metal clusters into the catalyst structure leverages synergistic effects between clusters and single atoms to adjust the d-band center position of the metal center (Figure 3f,g). This balances the adsorption capacity for oxygen-containing intermediates, facilitating multi-step electron transfer in ORR [53]. Clusters also enhance the structural stability of asymmetric atomic configurations, preventing aggregation or leaching of single atoms during catalysis and prolonging catalyst durability [54].
Introducing nonmetallic elements such as B and S into the catalyst to form an asymmetric structure lowers the oxidation state of the central active atom and alters the electron distribution in its d-orbitals. This simultaneously enhances O2 adsorption while accelerating the cleavage of the O-O bond. The co-coordination of multiple nonmetallic elements modifies the coordination structure of the central metal atom, thereby influencing the hybridization mode of its atomic orbitals. The synergistic effect between the nonmetallic elements and the metal elements promotes the conversion of *O to *OH, thereby biasing the oxygen reduction reaction (ORR) towards the four-electron pathway.

2.3. Research on Reaction Pathways

2.3.1. Rotating Ring–Disk Electrode Technique

The Rotating Ring–Disk Electrode (RRDE) technique is an advanced research method combining hydrodynamic regulation and electrochemical detection, widely used to investigate electrocatalytic mechanisms (e.g., ORR) [55]. The operational mechanism of the RRDE relies on forced convection generated by electrode rotation and selective detection of intermediates at the ring electrode, enabling high-sensitivity real-time dynamic analysis of ORR pathways. In RRDE measurements, the ORR occurs at the disk electrode, while intermediate products are selectively oxidized at the ring electrode. By analyzing the ratio of ring current to disk current, the H2O2 yield and electron transfer number can be calculated. For instance, Cui et al. [56] documented that the prepared BIM-Co2Zn8-500 catalyst achieved a H2O2 selectivity of 92.11% and an electron transfer number of 2.15 in the 0.05–0.65 V vs. RHE range, highlighting its exceptional selectivity for the 2eORR pathway (Figure 4a). Wang et al. [57] demonstrated that the C@PVI-(NCTPP)Fe-800 catalyst exhibited a H2O2 yield below 2% and an electron transfer number greater than 3.95 within the 0.2–0.8 V vs. RHE range (Figure 4b), confirming its predominance in the efficient 4eORR pathway. The Co-N5 catalyst constructed by Zhang et al. [58] exhibited a high H2O2 selectivity of up to 92.5% in 1 M KOH within the potential window of 0.3–0.7 V vs. RHE using the RRDE technique, with the electron transfer number (n) stably maintained between 2.0 and 2.3. This confirms that the dominant reaction pathway is two-electron oxygen reduction. The FeSA-Fe3C/NC catalyst synthesized by Huang et al. [59] consistently showed an H2O2 yield below 5.2% across a wide potential window of 0.2–0.8 V vs. RHE. The electron transfer number calculated via RRDE was n ≈ 4.0, further confirming that the ORR process proceeds efficiently via the four-electron pathway.

2.3.2. In Situ Raman Spectroscopy

In situ Raman spectroscopy is a real-time monitoring method that integrates Raman spectroscopy with in situ reaction conditions, enabling dynamic surveillance of molecular structures, chemical bond changes, and reaction intermediate formation under actual reaction environments, providing direct evidence for elucidating ORR mechanisms [62]. Raman spectroscopy is predicated upon the inelastic scattering effect, detecting frequency shifts caused by interactions between incident light and molecular vibrational modes to reveal molecular vibrations, rotations, and lattice vibrations. In situ Raman spectroscopy couples Raman spectrometers with reaction systems (e.g., electrochemical cells) via specially designed reaction cells or probes, allowing real-time monitoring of structural changes during reactions. The “in situ” feature ensures real-time observation without disrupting the sample’s native state or environment. For example, Pan et al. [60] demonstrated that for the Fe/NSC-vd catalyst in 0.1 M KOH, in situ Raman spectra revealed rapid generation and transformation of *OOH intermediates as the potential decreased, while the protonation process of O2 was slower (Figure 4c). This suggests that Fe-N coordination structures likely serve as the primary active centers for O2 adsorption and activation under alkaline conditions. In contrast, in 0.5 M H2SO4, in situ Raman spectra showed slower conversion of O2 to O2, while *OOH intermediate formation and transformation remained pronounced, indicating that Fe-S coordination structures may dominate as active centers for pivotal ORR steps in acidic media (Figure 4d). These findings highlight dynamic structural changes and intermediate adsorption behaviors during the 4eORR pathway of the Fe/NSC-vd catalyst under divergent electrolytes, confirming distinct reaction mechanisms in alkaline and acidic conditions.

2.3.3. In Situ ATR-SEIRAS

In situ attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) is a highly sensitive surface analysis technique combining attenuated total reflection (ATR) and surface-enhanced infrared absorption spectroscopy (SEIRAS) [63]. It enables real-time monitoring of molecular vibrational information of surface-adsorbed species in practical reaction environments (e.g., electrochemical interfaces, catalytic reactions). The surface enhancement effect grants this technique ultra-high sensitivity for detecting surface species, allowing identification of trace amounts of low-concentration substances and even sub-monolayer-level surface-selective detection, surpassing the limitations of conventional infrared spectroscopy. The in situ monitoring capability permits real-time observation in the actual ORR environment without complex sample pretreatment or labeling, capturing transient intermediates and kinetic information during reactions. With surface specificity, high sensitivity, and environmental compatibility, in situ ATR-SEIRAS has become one of the most effective techniques for elucidating molecular behavior at solid/liquid/gas interfaces. Exemplifying this, Gao et al. [61] analyzed reaction intermediates in the CoMn-NSC catalyst in a 0.1M KOH electrolyte using in situ ATR-SEIRAS: the O-O stretching vibration at 1240 cm−1 indicated the presence of superoxide intermediates (O2), the HOH bending vibration at 1638 cm−1 reflected interactions between water molecules and intermediates, and the broad peak at 3000–3550 cm−1 corresponded to O-H stretching vibrations, confirming hydroxyl adsorption (Figure 4e,f). By tracking potential-dependent spectral changes, the generation of oxygen-containing intermediates and transformation of superoxide intermediates were monitored, validating the CoMn-NSC catalyst’s efficacy in promoting the 4eORR pathway.

2.3.4. Density Functional Theory

Density functional theory (DFT) calculations, a quantum mechanical computational method, have become a core tool in materials simulation by efficiently solving electron density [64]. The Hohenberg–Kohn theorem, one of the cornerstones of DFT, states that the ground-state energy of a system is a unique functional of electron density, meaning all ground-state properties are derivable from a given electron density distribution. The Kohn–Sham equations, another theoretical foundation of DFT, incorporate the exchange–correlation energy functional, which describes the exchange and correlation interactions between electrons. Accurate calculation of this energy is critical to DFT precision. Although approximative treatments of the exchange–correlation functional introduce errors, DFT demonstrates significant value in predicting material properties and guiding experiments. In ORR studies, DFT is applied to analyze the geometric and electronic structures of catalytic sites [65]. For example, Zhao et al. [66] revealed that high-shell heteroatoms in the FeCoN6 catalyst regulate charge distribution through DFT combined with machine learning, significantly reducing overpotential. Jia et al. [67] synthesized the O-PtCo-FeNC catalyst and used DFT to demonstrate that Fe sites lower the d-band center of Pt, weakening intermediate adsorption energy and enhancing ORR activity and stability. DFT calculations also determine reaction mechanisms and pathways by analyzing intermediate adsorption and free energy landscapes. Liu et al. [68] constructed four structural models for the FeSA/N,S-PHLC catalyst via DFT, highlighting the correlation between the FeN3SOH/Pyridine N model and oxygen intermediates. By calculating Gibbs free energy diagrams for different models at U = 1.23 V, the FeN3SOH/Pyridine N model exhibited relatively low overpotential during *OOH transformation, favoring O-OH bond cleavage and promoting the 4eORR pathway.
DFT has become an indispensable tool in ORR research, enabling comprehensive analysis from atomic-level mechanisms to macroscopic performance prediction. It deepens understanding of ORR mechanisms, accelerates the development of low-cost, high-performance catalysts, and drives advancements in technologies such as fuel cells and metal–air batteries.

2.3.5. X-Ray Absorption Near-Edge Structure

X-ray absorption near-edge structure (XANES) is a cutting-edge spectroscopic technique based on synchrotron radiation sources. It analyzes the local electronic structure and geometric environment of materials by detecting changes in the probability of core electron excitation to unoccupied orbitals for specific elements. Its working principle involves measuring the X-ray absorption coefficient of a specific element in the region near its absorption edge (typically from 10–20 eV below the edge to 30–50 eV above the edge). When the energy of the incident X-ray photon reaches a level sufficient to excite a core electron of the specific element into an unoccupied state (the conduction band or molecular orbital), the absorption coefficient exhibits a sharp rise (the absorption edge) [69]. As revealed by Co K-edge XANES spectroscopy analysis of the ZIF-ZC91@Co(OH)2-VCo catalyst synthesized by Zhang et al. [70], the Co K-edge absorption edge shifts towards higher energy. This indicates an increase in the Co oxidation state from +2 to nearly +3. This valence state elevation arises because the cation vacancies (VCo) induce electron loss from surrounding Co atoms, triggering an electronic restructuring. This optimization enhances the catalyst’s adsorption capability for the *OOH intermediate, consequently improving H2O2 selectivity. Analysis of both the Cu K-edge and L-edge XANES for the Cu-S1N3 asymmetric structure catalyst synthesized by Jiang et al. [71] revealed that the Cu L-edge position lies between those of CuPc and CuS. Combined with the K-edge absorption edge position, this indicates an average Cu oxidation state of +1.97. During the oxygen reduction reaction (ORR) process, the absorption edge shifted towards lower energy and the white line intensity decreased, confirming a reduction in the Cu oxidation state from +2 to +1. This demonstrates that the low-valent Cu+ species serves as the ORR active center. The results confirm the existence of C–S and N–Cu bonds, providing evidence for the direct coordination of S to Cu.

2.3.6. Fourier Transform Infrared Spectroscopy

Fourier Transform Infrared Spectroscopy (FTIR) is a powerful technique used to obtain the infrared absorption or emission spectra of solid, liquid, or gas samples. It investigates the molecular structure, chemical bonds, and functional groups of substances based on their selective absorption of infrared radiation. The working principle of FTIR involves using an interferometer to generate an interferogram that contains information across all infrared frequencies. This interferogram passes through the sample. A detector then receives the interferogram signal modified by the sample’s absorption. Subsequently, a computer performs a Fourier transform on this complex interferogram signal, converting it into the conventional infrared absorption spectrum. In situ Fourier Transform Infrared Spectroscopy (In situ FTIR) is an advanced technique that enables real-time monitoring of dynamic structural changes in materials under authentic reaction conditions (such as specific temperature, pressure, atmosphere, or electrochemical environments). This is achieved by integrating a specialized reaction cell (e.g., high-temperature/high-pressure cell, electrochemical cell, catalytic reaction cell) into the FTIR optical path, allowing the infrared beam to pass directly through the sample while it is in its operational state. By rapidly acquiring time-resolved interferograms and performing real-time Fourier transforms, instantaneous spectra revealing molecular bond vibrations/rotations during the reaction are obtained. This allows for the capture of real-time information on intermediates, surface-adsorbed species, and reaction pathways [72]. For the ZnCo-MTF catalyst prepared by Liu et al. [73], Fourier Transform Infrared Spectroscopy (FTIR) revealed that the precursor material ZnCo-ZIF exhibited characteristic peaks of 2-methylimidazole (e.g., C=N stretching vibration, C-H bending vibration, etc.). During the reaction process, the characteristic peaks of 2-methylimidazole gradually diminished, while those of 1,2,3-triazole progressively intensified (e.g., N-H/N-N vibrations, triazole ring skeletal vibrations). The final product, ZnCo-MTF, displayed only the strong characteristic peaks of 1,2,3-triazole, with the 2-methylimidazole peaks completely absent. This demonstrates that the FTIR spectrum of the final product matches the characteristic peaks of the pure 1,2,3-triazole ligand, indicating the complete replacement of 2-methylimidazole by 1,2,3-triazole within the ZnCo-MTF framework, resulting in a new metal–organic framework (MOF) structure. This transformation provides the structural foundation for the subsequently enhanced catalytic performance (high H2O2 production rate/selectivity). During the synthesis of the Fe-N4-SM catalyst constructed by Shen et al. [74], the precursor containing the defect-prone ligand 5-ATZ exhibited a Zn-N vibration peak at 423.0 cm−1. The displacement of this Zn-N vibration peak, coupled with the preservation of the main framework characteristics, confirmed the successful synthesis of the dual-ligand ZTIF precursor. Furthermore, this analysis revealed that the thermal instability of 5-ATZ underpins the subsequent selective C-N bond cleavage and serves as the structural basis for atomic strain modulation.

3. Control of Oxygen Reduction Reaction Pathways

3.1. Control of Two-Electron Pathways

It is a reasonable choice to introduce one or more nonmetallic elements instead of the nonmetallic elements in the original structure to form an asymmetric atomic structure in order to control the tendency of the catalyst to follow the two-electron pathway in the reaction, so as to improve the catalytic efficacy of 2eORR [75].
The incorporation of heteroelements can create catalysts with asymmetric configurations, which induces uneven charge distribution at the central active sites and modifies electron orbital states. As elucidated by Ivan Parkin et al. [76], a surface engineering strategy was employed to synthesize atomically dispersed Co on N-doped carbon black, forming an asymmetric coordination configuration (Co-C/N/O). The asymmetric structure facilitates hybridization between Co 3d orbitals and C/N/O p-orbitals, which lowers the d-band center of Co, modulates the electronic structure, and weakens the adsorption strength of intermediates (e.g., *OOH) (Figure 5a,b), thereby preventing over-strong adsorption that would drive the 4e pathway. Oxygen atoms act as electron-withdrawing sites from Co, while N and C serve as electron-donating components, establishing a charge gradient that stabilizes *OOH adsorption. DFT calculations reveal a positive correlation between the vertical displacement of Co atoms from the graphene plane and the *OOH adsorption energy, highlighting the potential for geometric configuration optimization to fine-tune catalytic activity. Similarly, Zhang et al. [58] synthesized a Co-N5 catalyst by precisely controlling the coordination number of Co atoms through the thermal stability differences of precursors (Figure 5c). The introduction of an additional N atom in the Co-N5 configuration induces uneven charge distribution and electronic rearrangement, breaking the linear scaling relations (LSRs) intrinsic to conventional symmetric Co-N4 catalysts and thereby resolving the activity–selectivity trade-off. DFT calculations demonstrate that the altered d-orbital occupancy of Co optimizes the adsorption energy of *OOH intermediates, preventing both O-O bond cleavage (4e pathway) and insufficient O2 activation (blocked 2e pathway). In situ Raman spectroscopy reveals that Co-N5 sites stabilize *OOH intermediates and slow their consumption rate. KSCN poisoning experiments further confirm that atomically dispersed Co serves as the catalytically active center, with changes in d-orbital states being the key factor for enhanced performance.
Additionally, the incorporation of heteroatoms modulates the adsorption capability of the central active atom toward oxygen-containing intermediates, tailoring the adsorption strength to favor *OOH desorption. Gao et al. [77] synthesized an asymmetrically coordinated Zn-N3O catalyst by adsorbing Zn ions onto polypyrrole (PPy) to achieve an elevated Zn loading of 11.34 wt%, forming Zn-N/O bonds (Figure 5d). The introduction of O-atom coordination withdraws electrons from Zn, increasing its oxidation state and redistributing its electron cloud density. This induces a contraction of Zn’s d-orbital electron density, which strengthens the interaction between Zn’s d-orbitals and the antibonding orbitals of *OOH, thereby diminishing the O-O bond dissociation energy barrier and optimizing the adsorption energy of Zn toward *OOH intermediates to favor the 2eORR pathway. DFT calculations reveal that the density of states (DOS) near the Fermi level in the Zn-N3O structure significantly increases, indicating enhanced hybridization between Zn’s d-orbitals and O’s p-orbitals, which improves intermediate adsorption. Zn K-edge EXAFS spectroscopy further confirms structural modulation by O-coordination, showing altered coordination numbers and bond lengths between Zn and N/O atoms. Additionally, Han et al. [78] developed a high-coordination Co-N5 moiety by precisely regulating the coordination environment of Co atoms, leveraging adjacent electron-withdrawing epoxy groups to construct a Co-N5-O-C catalyst. By adjusting the number of N atoms in the first coordination shell of Co single-atom catalysts (SACs) or partially substituting them with O atoms, Co-NxOy SACs were formed. This modification alters the charge distribution and d-band center of the metal atom, achieving an optimized electronic structure that shifts the reaction pathway toward the 2eORR. The high-coordination N atoms modulate Co’s d-band center, fine-tuning the adsorption energy of the *OOH intermediate. DFT calculations demonstrate that the Co-N5 moiety exhibits a reduced overpotential for the 2eORR compared to the Co-N4 configuration, thereby enhancing pathway selectivity toward H2O2 production.
Bimetallic or multimetallic-doped ORR catalysts typically orchestrate the adsorption strength of intermediates (*OOH) to achieve a balance where oxygen-containing intermediates neither undergo over-adsorption (leading to further reduction via the 4e pathway) nor under-adsorption (causing sluggish kinetics), thereby avoiding O-O bond cleavage [79,80]. In such catalysts, the complementary electronic properties of the two metals synergize: one metal adsorbs and activates O2 to form *OOH, while the other weakens *OOH adsorption, enabling its desorption to generate H2O2. Additionally, differences in bimetallic coordination alter the electronic distribution of active sites, enhancing selectivity for the 2e pathway. The asymmetric coordination environment further induces hybrid orbital energy matching with O2’s *π orbitals, facilitating single-electron transfer to produce superoxide radicals (O2), which subsequently undergo selective reduction to H2O2. This dual-metal synergy not only modulates intermediate adsorption but also tailors electronic interactions to stabilize the 2e pathway. Liu et al. [73] fabricated a hollow nanocubic-structured catalyst (ZnCo-MTF) using ZnCo-ZIF as a precursor and 1,2,3-triazole as a ligand. As shown in Figure 6a,b, the synergistic interaction between Zn and Co induces a unique side-on adsorption configuration for O2: one oxygen atom weakly binds to Zn, while the other strongly interacts with Co. This binding mode effectively polarizes the O2 molecule, lowering the O-O bond dissociation barrier and promoting desorption of the *OOH intermediate, thereby selectively driving the 2e pathway to generate H2O2. The coordination between 1,2,3-triazole ligands and Co forms *d-pπ delocalized orbitals**, increasing the occupation of Co-N antibonding orbitals and weakening Co-N bonds to activate the Co center. Simultaneously, Zn incorporation optimizes the electron density of Co via charge transfer, enhancing O2 adsorption capability. DFT calculations reveal that the DOS of ZnCo-MTF spans the Fermi level, signifying superior conductivity compared to monometallic systems, which accelerates electron transfer kinetics.
Strategies to disrupt structural symmetry in catalysts extend beyond metal/nonmetal doping and include defect engineering to reconstruct the original framework. For instance, Zhang et al. [70] synthesized a ZIF-ZC91@Co(OH)2-VCo catalyst via electrochemical self-optimization reconstruction of ZIF-ZC91, where partial Zn2+ leaching generates cobalt cation vacancies. These vacancies, formed through selective Zn dissolution, modulate the catalyst’s electronic structure and surface morphology. The presence of cation vacancies optimizes the adsorption strength of *OOH intermediates (Figure 6c), suppresses O-O bond cleavage, and promotes the 2eORR pathway. Mechanistically, the vacancies alter the oxidation state of Co and refine its d-band characteristics, thereby reducing the adsorption energy of *OOH. DFT calculations confirm that the energy barrier for the 2e pathway is lower than that of the 4e route, effectively inhibiting O-O bond dissociation.
Notably, incorporating heteroelements into catalyst structures not only induces disparate charge distribution at the central atom but also creates defects near the active sites. These defects generate localized electron-rich regions or charge-asymmetric environments, stabilizing the *OOH intermediate (precursor to H2O2) and preventing its further reduction to *O or *OH, thereby blocking the 4eORR pathway. As demonstrated by Zhang et al. [81], a coaxial Co single-atom catalyst (CoSA-N-C/CNTs) was synthesized via an SCVD strategy. Concentrated nitric acid oxidation was employed to introduce surface defects (e.g., carbon vacancies) and oxygen-containing functional groups on CNTs, forming defective CNTs (d-CNTs). These defects provide abundant anchoring sites, enabling the stable immobilization of Co atoms as single-atom sites on the carbon substrate to form high-density Co-Nx active centers. The defects alter the electronic state of the carbon substrate, modulating the electron density of Co-Nx sites and thereby tuning the adsorption strength of intermediates (*OOH) (Figure 6d,e). DFT calculations reveal that the ΔG*OOH of Co-Nx sites approaches the apex of the volcano plot, indicating that defect-induced electronic optimization achieves near-optimal adsorption energy for enhanced 2eORR selectivity. Additionally, the defect-engineered CoSA-N-C/CNTs exhibit a microporous structure, where the pores and surface roughness (resulting from defects) accelerate the diffusion of reactants (O2) and products (H2O2), while exposing more active sites to enhance catalytic performance.

3.2. Control of Four-Electron Paths

3.2.1. Incorporation Forms of Nonmetallic Elements

To modulate the adsorption energy of central active sites, nonmetallic elements are incorporated into inherently symmetric and stable catalysts, constituting asymmetric atomic structures [82]. Due to the higher electronegativity of nonmetallic elements compared to metallic counterparts, their incorporation alters the electron density distribution of the central active element. The nonmetallic dopant withdraws electrons from neighboring atoms (e.g., carbon or the central metal atom), creating electron-deficient regions. This polarization effect strengthens O2 adsorption and activation, lowers the free energy barrier for oxygen adsorption reactions, and consequently augments ORR activity. For instance, Ding et al. [83] synthesized a single-atom Co catalyst with N/B asymmetric coordination (Co-N3B) via a pyrolysis method and anchored it on boron–nitrogen Co-doped graphene (BNG) to form Co-N3B active sites (Figure 7a). The incorporation of B atoms, owing to their low electronegativity, facilitates electron transfer to the Co center, reducing its oxidation state and increasing its electron density. This electronic modulation enhances Co’s ability to adsorb and activate oxygen intermediates (e.g., *O, *OOH). Projected density of states (PDOS) analysis from DFT reveals stronger d-p orbital hybridization between Co’s d-orbitals and O’s p-orbitals near the Fermi level in Co-N3B, indicating enhanced orbital overlap that promotes the 4eORR pathway. Attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) experiments demonstrate that the Co-BNG catalyst exhibits stronger adsorption capacity for oxygen intermediates (e.g., *O2 and *OOH) and maintains excellent stability during prolonged reaction cycles. Similarly, Zhang et al. [84] fabricated a Co-N3-C/CNT catalyst via a bidentate pre-coordination-assisted pyrolysis strategy. DFT calculations elucidate how the asymmetric low-coordination structure of Co-N3 disrupts the symmetric charge distribution of Co-N4. The reduced coordination (Co-N3) induces asymmetric electron density distribution at the Co atom (Figure 7b–d), enhancing oxygen activation capabilities and optimizing the adsorption/desorption equilibrium of ORR intermediates. The Co-N3 configuration exhibits lower energy barriers during ORR, particularly in the critical *O-to-*OH conversion step, demonstrating superior intrinsic catalytic activity compared to symmetric Co-N4 counterparts. Furthermore, Yin et al. [85] successfully constructed a CoN3S-PC catalyst by anchoring single Co atoms on porous carbon (PC) and subsequently doping S atoms into the first coordination shell of the Co active sites, forming an asymmetric CoN3S coordination structure. The incorporation of S atoms induces electronic rearrangement and geometric distortion at the Co active center, disrupting its charge symmetry and enhancing interactions between Co 3d orbitals and oxygen molecules/intermediates (Figure 7e,f). This strengthens O2 adsorption at the Co sites and accelerates O-O bond cleavage. DFT calculations reveal that S doping modifies the electronic distribution of Co’s d-orbitals, promoting O2 adsorption/activation while suppressing H2O2 generation. Concurrently, the CoN3S sites exhibit a reduced thermodynamic energy barrier and kinetic superiority for the 4eORR pathway, ensuring efficient and selective oxygen reduction.
The coordination of nonmetallic elements directly with the central active site increases the electron density peripherally to the active atom, thereby influencing the adsorption energy of intermediates. Jiang et al. [71] designed an asymmetrically coordinated Cu-S1N3 single-atom active site via an atomic interface engineering strategy, constructing a single-copper-atom catalyst (S-Cu-ISA/SNC) based on MOF-derived hierarchically porous carbon. The introduction of S atoms disrupts the conventional symmetric CuN4 coordination, forming an asymmetric Cu-S1N3 active site that modifies the electronic structure around the central Cu atom. The relatively lower electronegativity of S atoms reduces the loss of valence electrons from Cu atoms, altering the local electron density distribution (Figure 8a) and thereby optimizing the adsorption free energy of ORR intermediates (e.g., *OOH, *O, *OH) (Figure 8b) while diminishing the reaction energy barrier. In situ soft X-ray absorption near-edge structure (XANES) tests and in situ FTIR spectroscopy elucidated the electronic and atomic structural evolution of Cu sites during ORR. DFT calculations, the Koutecky–Levich (K-L) equation, and RRDE tests collectively validated the catalyst’s performance, confirming the presence of low-valent Cu active centers, an electron transfer number (n ≈ 4), and a low H2O2 yield (<4%). Yang et al. [86] synthesized an FeN3S/Cmeso catalyst by embedding single Fe atoms into a mesoporous carbon matrix and introducing S atoms to directly coordinate with the Fe center, forming an FeN3S coordination structure. The direct coordination of S atoms to the Fe center disrupts the symmetry of the FeN4 configuration, increasing the electron density around Fe (Figure 8c). The incorporation of S lowers the d-band center of Fe, weakening the orbital overlap between active sites and oxygenated intermediates (e.g., *OH), thereby optimizing their adsorption energy (Figure 8d). DFT calculations demonstrate that the S-induced coordination enhances the electron cloud density of Fe atoms, which reduces the orbital overlap between active sites and oxygen-containing intermediates. This attenuated adsorption strength facilitates intermediate desorption and lowers the energy barrier during the oxygen reduction process.
The co-coordination of multiple nonmetallic elements modifies the coordination number and coordination structure of the central metal atom, thereby influencing the hybridization mode of metal atomic orbitals [87]. Different coordination structures exhibit distinct geometric configurations and symmetries, which constrain or guide the direction and extent of hybridization in metal atomic orbitals. Following multi-nonmetallic element coordination, charge transfer occurs between the nonmetallic elements and the central metal atom. Nonmetallic elements tend to attract electrons, causing a partial transfer of electron density from the metal atom to the nonmetallic ligands and a reduction in the electron cloud density of the metal atom. This charge redistribution modifies the electronic configuration of the metal atom’s d-orbitals, affecting the formation and characteristics of its hybridized orbitals. Additionally, highly electronegative nonmetallic elements can lower the d-band center of metal sites through electronic modulation, geometric restructuring, and synergistic catalytic effects. This weakens the adsorption strength toward oxygen-containing intermediates (such as *OOH, *O, and *OH), bringing the adsorption energy closer to the peak of the “volcano curve”. Such optimization of reaction kinetics significantly promotes the 4eORR pathway while suppressing H2O2 generation routes, such as reducing the probability of *OOH protonation to form H2O2. As reported by Liu et al. [88], an asymmetric nitrogen–oxygen-coordinated tin single-atom catalyst (Sn-N/O-C) with Sn atoms as active centers was developed via a simple one-step pyrolysis strategy, forming asymmetric SnN2O sites through the co-coordination of N and O atoms. The N and O co-coordination disrupted the symmetry of the Sn sites, optimized the p-orbital electron distribution of Sn, and induced strong hybridization between the Sn 5p orbitals and the O2 2p orbitals. This altered the theoretical d-band center of Sn, promoting electron transfer and accelerating O2 activation and reduction (Figure 8e,f). DFT calculations revealed that the Sn atoms with broken symmetric coordination facilitated charge transfer and O2 activation. RRDE measurements demonstrated a hydrogen peroxide (H2O2) yield of <5% in alkaline media and <3% in acidic media, with an electron transfer number close to 4, indicating the dominance of the 4e pathway. Similarly, Liu et al. [89] introduced O atoms into the first coordination shell of Fe and incorporated adjacent defects, inducing a transition of Fe from a low-spin state to a medium-spin state and forming a Fe-N3O coordination structure, thereby synthesizing a single-atom iron catalyst (Fe-N/O-C). The coordination of N and O atoms in Fe-N/O-C redistributes the charge, increases the electron density at the Fe sites, and significantly weakens the adsorption of oxygen-containing species such as O2, mitigating over-adsorption of O2. Theoretical calculations reveal that the d-band center of Fe-N/O-C is lower than that of Fe-N-C, leading to weakened adsorption of intermediates. The formation of adsorbed *OOH from O2 becomes the new rate-determining step (replacing the original *OH desorption), with a reduced energy barrier. In situ Raman spectroscopy further confirms this shift in the PDS. RRDE tests demonstrate an electron transfer number close to 4 and an extremely low H2O2 yield (<5%), indicating that a highly efficient 4e pathway dominates the reaction.
Figure 8. (a) The molecular orbital of *O adsorbed on Cu-S1N3 is in S-Cu-ISA/SNC. (b) ORR overpotential (ηORR) as a function of the free adsorption energy (∆G*O) of *O on different Cu central groups. (a,b) Reproduced with permission [71]. Copyright 2020, Springer Nature. (c) Charge density difference plot of FeN3S1. (d) Absorption energy of *OH intermediates at different active sites. (c,d) Reproduced with permission [86]. Copyright 2024, Elsevier Ltd. (e) The variation of Gibbs free energy of the ORR pathway at different Sn sites under a potential bias voltage of U = 0 V. (f) The charge density difference plot and relative charge state of Sn. (e,f) Reproduced with permission [88]. Copyright 2024, Wiley-VCH Verlag.
Figure 8. (a) The molecular orbital of *O adsorbed on Cu-S1N3 is in S-Cu-ISA/SNC. (b) ORR overpotential (ηORR) as a function of the free adsorption energy (∆G*O) of *O on different Cu central groups. (a,b) Reproduced with permission [71]. Copyright 2020, Springer Nature. (c) Charge density difference plot of FeN3S1. (d) Absorption energy of *OH intermediates at different active sites. (c,d) Reproduced with permission [86]. Copyright 2024, Elsevier Ltd. (e) The variation of Gibbs free energy of the ORR pathway at different Sn sites under a potential bias voltage of U = 0 V. (f) The charge density difference plot and relative charge state of Sn. (e,f) Reproduced with permission [88]. Copyright 2024, Wiley-VCH Verlag.
Catalysts 15 00615 g008

3.2.2. Co-Incorporation Forms of Nonmetallic and Metallic Elements

The introduction of metal elements beyond the central active site disrupts the local charge symmetry of the original catalyst structure and redistributes the electron cloud density of the active centers [90]. The incorporation of additional metal elements also narrows the energy band of the central active metal sites, enhancing electron mobility. This enables faster electron transfer during catalytic reactions, thereby accelerating the ORR rate. The synergistic interplay of multiple metal elements modulates the adsorption strength of active sites toward oxygen-containing intermediates (e.g., *O, *OOH), ensuring effective adsorption of these intermediates on the catalyst’s surface while preventing overly strong adsorption that impedes desorption. This equilibrium promotes the 4e transfer pathway. Additionally, multimetal systems provide diverse active sites that act synergistically and collectively participate in the ORR process [80]. As reported by Shang et al. [91], a three-step pyrolysis strategy was employed to anchor Se atoms and other metal atoms (e.g., Fe) onto a C2N substrate, forming SeN2-MN2 active sites, thereby synthesizing an asymmetric Se-based diatomic SeM-C2N catalyst (Figure 9a). Among these, SeFe-C2N demonstrated optimal ORR performance. The heteronuclear Se atom induces short-range modulation to polarize the charge distribution of the Fe atom, reducing its oxidation state, enhancing the co-adsorption capacity for intermediates (e.g., *O and *OH), and forming heteronuclear diatomic sites, with diatomic sites accounting for 82% of the active centers. The introduction of Se increases the electron density at the Fe atom, creating a localized polarized electric field that lowers the ΔG value of the rate-determining step. This moderates the adsorption strength of intermediates, accelerates O2 activation and electron transfer, and achieves more efficient conversion of O2 to H2O. DFT calculations integrated with operando XAS techniques confirmed that Se sites adsorb *OOH, while the Fe-Se dual sites co-adsorb *O, synergistically promoting the *O→*OH conversion. The K-L equation, fitted from LSV curves at varying rotation speeds, yielded an electron transfer number of n ≈ 4. RRDE measurements directly determined a H2O2 yield of <4%, and the calculated electron transfer number (based on ring and disk currents) ranged from n = 3.97 to 4.05, conclusively confirming the dominance of the 4e pathway. As reported by Chang et al. [92], a Fe-Co dual-site N-doped porous carbon catalyst (Fe,Co/N-C) was synthesized using a host–guest strategy. By adjusting the precursor ratio of FeCl3 to Co-ZIF (optimal at Fe/Co = 1:1), the Fe,Co/N-C catalyst incorporated Fe-Co dual sites that provided more favorable active sites for O-O bond activation. Compared to single-metal sites, the dual sites significantly reduced the energy barrier for O-O bond cleavage, thereby enhancing catalytic activity. DFT calculations confirmed that upon O2 adsorption on the Fe-Co dual sites, the O-O bond was stretched, and the cleavage energy barrier decreased to 0.25 eV (vs. 0.65 eV for a single Fe site), facilitating the four-electron pathway. K-L equation analysis and RRDE tests indicated an electron transfer number of n ≈ 3.96 and a H2O2 yield below 1.17%.
Incorporating low-electronegativity nonmetallic elements into catalysts possessing dual-metal sites can further modulate the distribution of surface electron cloud density and alter the spin state of metal sites. The interaction between low-electronegativity nonmetallic elements and metal sites enhances the electronic properties of the metal centers. For example, in Fe-based catalysts, the introduction of Si induces a transition of Fe ions from a low-spin state to a medium-spin state. Fe sites in the medium-spin state exhibit higher reactivity, improved oxygen adsorption capacity, and stronger interactions with oxygen-containing intermediates (e.g., *O and *OOH), thereby accelerating the ORR rate [94]. Additionally, low-electronegativity nonmetallic elements can form isolation layers during high-temperature treatments, preventing the aggregation of metal sites into nanoparticles and maintaining the dispersion and high activity of active sites. Finally, these elements tend to donate electrons to the dual-metal sites, causing an upward shift in the d-band center, which fortifies the adsorption of ORR intermediates (*OOH, *OH). As reported by Gao et al. [61], an adsorption–pyrolysis process was utilized to anchor adjacent S,N-coordinated Co atoms and N-coordinated Mn atoms (forming CoN2S-MnN3) onto a nitrogen-doped carbon substrate, constructing CoN2S-MnN3-2OH active sites and synthesizing the CoMn-NSC catalyst. In this structure, S atoms preferentially coordinate with Co to form CoN2S, while Mn retains its N-coordinated MnN3 configuration. The strong interactions between Co and Mn atoms, coupled with multi-electron filling effects induced by asymmetric charge distribution (Figure 9b), enhance orbital hybridization between Co and O2, lower the O-O bond cleavage energy barrier, optimize adsorption energies of reaction intermediates, and improve electron transfer efficiency through dual-atom synergy. Here, Co acts as the primary active site for O2 adsorption, while Mn serves as an auxiliary site that modulates the local electronic environment. This adjustment shifts the d-band center of Co closer to the Fermi level, strengthening its adsorption and activation capabilities for O2, thereby boosting overall catalytic activity (Figure 9c). DFT calculations revealed that the d-band center of Co in the asymmetric structure enhances hybridization with the *π orbitals of O2, improves antibonding electron filling efficiency, and optimizes adsorption strength. K-L equation analysis of polarization curves and RRDE tests confirmed an electron transfer number of n ≈ 4 and a H2O2 yield below 5%. In situ Raman spectroscopy and in situ ATR-SEIRAS further validated the efficient adsorption, activation, and reduction of O2 on the CoMn-NSC catalyst during the reaction. Su et al. [95] employed a heteroatomic pair cooperative modulation strategy to alter the coordination environment of Fe sites by introducing Co and B heteroatomic pairs, thereby synthesizing the Fe-B-Co/NC catalyst. The introduction of Co atoms achieved electronic synergy between Fe and Co, modulating the d-band center position of Fe and balancing the competitive relationships between RDS during the oxygen reduction reaction. Concurrently, asymmetric coordination transformed Fe from an intermediate-spin state (t52ge1g) to a high-spin state (t42ge2g). The high-spin Fe sites enhanced orbital hybridization with O2, promoting O-O bond cleavage to directly generate *O intermediates (rather than accumulating *OOH), thereby enabling rapid four-electron reaction kinetics and improving oxygen reduction efficiency. K-L equation and RRDE tests confirmed that the Fe-B-Co/NC catalyst exhibited an average electron transfer number of 3.87 within the 0.5 V–0.65 V potential range, indicating a 4e pathway during ORR with a H2O2 yield below 6.9%. In situ EIS and SR-FTIR techniques confirmed that Fe sites accelerate O-O bond cleavage in *OOH to rapidly form *O intermediates. In 0.1 M KOH, the Fe-B-Co/NC catalyst demonstrated a half-wave potential (E1/2) of 0.891 V and diffusion-limited current density (Jd) of 5.9 mA cm−2, surpassing commercial Pt/C (0.851 V, 5.19 mA cm−2). During 50 h stability testing, 95% of the initial current density was retained.

3.2.3. Other Incorporation Forms (Functional Groups, Defect Sites, Metallic Clusters, etc.)

Similar to 2eORR, strategies to augment 4eORR selectivity extend beyond doping with solely metallic or nonmetallic elements. Functional groups (e.g., OH) can also be incorporated to coordinate with central active atoms, or combined with elements (e.g., S) for multi-coordination, optimizing the electronic structure of the central atom [96]. Additionally, carbon frameworks with pronounced edge effects (e.g., HsGDY) reconfigure the edge morphology of the catalyst structure, inducing unique electronic states at the edges. This effectively modulates the material’s band structure, enhancing pathway selectivity. Constructing asymmetric anchoring sites at the edges of the carbon substrate strengthens the bonding between metal atoms and the carbon matrix, reducing metal aggregation and leaching during reactions. As documented by Hou et al. [93], iridium single-atom catalysts (Ir-N-HsGDY) were synthesized by anchoring Ir atoms at the sp-N and pyridinic N sites of hydrogen-substituted graphdiyne (HsGDY) and coordinating them with hydroxyl groups (OH), forming IrN2(OH)3 active sites. The introduction of OH ligands broadened the Ir d-band and displaced its center downward, enhancing orbital hybridization between Ir and O/N, optimizing the electronic structure (Figure 9d,e), weakening the adsorption strength of intermediates (e.g., *OH), and reducing reaction energy barriers. The macroporous structure of HsGDY provided space for OH incorporation, while its hierarchical design effectively isolated Ir precursors to prevent aggregation and ensure single-atom dispersion. DFT calculations confirmed that the IrN2(OH)3 active site exhibited the lowest ORR overpotential (0.54 eV), significantly outperforming unmodified IrN2 (2.24 eV). When applied to PEMFCs, the Ir-N-HsGDY catalyst achieved a half-wave potential (E1/2) of 0.82 V vs. RHE in acidic media (0.1 M HClO4), comparable to commercial Pt/C with superior methanol tolerance. In alkaline media (0.1 M KOH), it demonstrated a half-wave potential (E1/2) of 0.89 V vs. RHE within the potential window above 0.78 V, significantly outperforming commercial Pt/C, and exhibited only an 18 mV negative shift in E1/2 after 5000 CV cycles, indicating high durability. Lv et al. [97] synthesized the Mn-N-HsGDY catalyst by asymmetrically anchoring Mn atoms as single atoms at edge sites of the HsGDY carbon substrate, incorporating two nitrogen species (sp-N and pyridinic N) and OH ligands to form asymmetric MnN2(OH)3 active sites, thereby disrupting the electronic state of conventional symmetric M-N4 configurations. The strong coordination between sp-N (high electronegativity) and Mn, combined with the electron-withdrawing effect of OH ligands, synergistically drove the Mn center to a high oxidation state (+3 to +4), altered the electronic structure of Mn atoms, and reduced the desorption energy barrier of the intermediate *OH. DFT calculations elucidated that the asymmetric coordination of the MnN2(OH)3 site weakened *OH adsorption, lowering the RDS energy barrier to 0.502 eV, significantly outperforming symmetric sites (e.g., MnN4). RRDE-derived electron transfer numbers and K-L equation analysis of LSV curve slopes at different rotation rates both yielded an average electron transfer number (n) close to 4, confirming the dominance of the 4eORR pathway.
Additionally, etching defects (e.g., vacancies) adjacent to the central atom that disrupt the original symmetric electronic distribution can also enhance 4eORR performance [98]. Defects alter the electron cloud distribution around the central active atom, modifying electron delocalization, thereby influencing the catalyst’s electronic structure. This creates an unsaturated coordination environment for atoms at defect sites, providing additional active sites for ORR [99]. Zhang et al. [100] prepared an atomically dispersed Fe-N-C catalyst modified with proximal carbon vacancies (Fe-N4-Vc) through pyrolysis combined with urea etching to selectively cleave C-N bonds near Fe-N4 sites, forming carbon vacancies (Vc). In the Fe-N4-Vc structure, the introduced adjacent carbon vacancies induced redistribution of electron density in Fe-N4, lowered the oxidation state of Fe (coexistence of Fe2+ and Fe3+ with an average oxidation state of +2.6), and generated asymmetric electronic states. This caused an upward shift of the Fe d-band center, reduced electron occupancy in antibonding orbitals, optimized adsorption behaviors of O2 and intermediates, lowered reaction free energy barriers, and thereby enhanced ORR activity. DFT calculations demonstrated that Vc significantly reduced the adsorption energy of O2 (from −2.27 eV to −1.66 eV) and optimized the adsorption free energy of intermediates (e.g., *OH). Among ORR free energy diagrams of distinct structures, Fe-N4-Vc exhibited the smallest RDS energy barrier and the highest activity. Shen et al. [74] synthesized the Fe-N4-SM catalyst by selectively cleaving C-N bonds around Fe-N4 sites to construct an asymmetric atomic strain environment, thereby modulating the electronic structure of the Fe center. The selective cleavage of C-N bonds adjusted the asymmetric atomic strain environment at Fe sites, perturbed the symmetric electronic structure of Fe-N4, modulated the d-band broadening and d-orbital electron distribution of the Fe center, and consequently altered the adsorption strength between the catalyst and ORR intermediates, enabling precise regulation of catalytic activity. RRDE tests and K-L equation analysis revealed an average electron transfer number of ~3.98 and H2O2 yield below 1%, confirming the dominance of the four-electron pathway. In situ Raman spectroscopy detected vibration peaks of *O2 and *OOH intermediates, supported by theoretical models, further validating the four-electron mechanism. DFT calculations demonstrated that the RDS of Fe-N4-SM shifted from *OH desorption (energy barrier 0.83 eV) to the first electron transfer step (energy barrier 0.25 eV).
When functional groups (e.g., OH) and defect vacancies are co-incorporated into a catalyst structure, they may exhibit a synergistic effect to jointly enhance the ORR performance. For example, Zhan et al. [101] employed a microenvironment engineering strategy to anchor Fe single atoms on a defective N-doped carbon (DNC) substrate by regulating micromorphology and interfacial microenvironments, synthesizing the Fe/DNC catalyst. Carbon vacancies reshaped the electronic properties of FeN4(OH) sites via inductive effects, optimizing hybridization between Fe 3d and O2 2p orbitals, balancing energy barriers for *OOH formation and *OH reduction, and regulating charge distribution at Fe sites through electron induction to promote O2 adsorption and intermediate desorption (e.g., *OH), thereby accelerating reaction kinetics. The introduction of axial hydroxyl (OH) ligands disrupted the symmetry of FeN4, enhancing adsorption capability for oxygen intermediates and reducing kinetic barriers. RRDE analysis revealed extremely low H2O2 yields (0.35% in alkaline, 6% in acidic, and <5% in neutral conditions) and an electron transfer number close to 4 (≈4.06), confirming the ascendancy of the 4e pathway and validating high ORR efficiency/selectivity. DFT simulations identified FeN4(OH) as the true active center, with the defective model (FeN4(OH)-A) exhibiting the lowest RDS energy barrier (0.43 eV), outperforming symmetric FeN4(OH) (0.64 eV).
Finally, co-incorporating metal active atoms and metal clusters into the catalyst structure can also generate asymmetric atomic configurations [102]. The presence of metal clusters modifies the charge distribution and local coordination environment surrounding metal active sites, facilitating O2 molecule access and adsorption at active sites while stabilizing the adsorption of subsequent reaction intermediates, thereby promoting 4eORR. Additionally, clusters and asymmetric atomic structures collectively create multi-active sites, enabling synergistic catalytic effects. Huang et al. [59] successfully synthesized the FeSA-Fe3C/NC catalyst by encapsulating Fe atoms and Fe3C clusters on nitrogen-doped carbon. The incorporation of Fe3C clusters disrupted the symmetric electronic structure of FeN4, depressed the d-band center of Fe atomic sites, thereby altering the adsorption strength of oxygen intermediates on Fe-N4-C sites, reduced the adsorption energy of *OOH intermediates (0.38 eV), accelerated O-O bond cleavage, and promoted the 4eORR pathway. Additionally, the high specific surface area (443 m2/g) and mesoporous structure (4 nm pore size) of the nitrogen-doped carbon enhanced active site exposure and mass transfer efficiency. The synergistic interaction between atomically dispersed Fe-N4 sites and Fe3C clusters provided an efficient catalytic platform for ORR. DFT calculations demonstrated that the introduction of Fe3C clusters increased the electron cloud density of Fe d-orbitals (Figure 9f), augmenting O2 adsorption and activation while weakening *OH adsorption strength, favoring *OH desorption. This optimized the adsorption/desorption behavior of intermediates, reduced reaction energy barriers, and improved ORR kinetics and activity. In situ infrared spectroscopy confirmed that *OOH intermediate formation was the rate-determining step, and the FeSA-Fe3C/NC catalyst effectively desorbed *OOH intermediates, thereby lowering reaction barriers and facilitating ORR progression.

4. Application of Oxygen Reduction Reaction

4.1. Device for Producing H2O2

Currently, the traditional anthraquinone process remains the dominant technology for H2O2 production due to its maturity and high capacity. This method hinges upon the hydrogenation–oxidation cycle of 2-ethylanthraquinone over palladium catalysts (hydrogenation to form hydroanthraquinone, followed by oxidation to regenerate anthraquinone and release H2O2), integrated with extraction and purification for continuous production. It offers advantages such as mature technology, high productivity (annual production exceeding 100,000 tons per single unit), and high product purity. However, it faces environmental and economic challenges, including dependence on fossil-fuel-derived hydrogen, high energy consumption, solvent leakage, catalyst costs, and byproduct accumulation. With increasing environmental regulations and advancements in renewable energy technologies, emerging methods like electrochemical synthesis are gradually overcoming technical barriers. The electrochemical production of H2O2 via the two-electron oxygen reduction reaction (2eORR) offers a green, low-energy alternative to the anthraquinone process [103,104]. This approach directly utilizes water and oxygen from air as feedstocks, eliminating reliance on fossil-fuel hydrogen or organic solvents, thereby reducing carbon emissions and pollution risks at the source. Reactions proceed under ambient conditions, circumventing the high energy consumption and safety hazards of high-temperature/pressure processes, and generate no anthraquinone degradation byproducts, minimizing wastewater treatment demands. Additionally, catalyst design can selectively suppress the 4eORR pathway, markedly potentiating H2O2 selectivity.
The design of devices for the electrochemical production of H2O2 is critical to augmenting synthesis efficiency and practicality, with current key technological routes including gas diffusion electrode reactors and flow cell systems [105].
A gas diffusion electrode (GDE) reactor is a device that achieves efficient electrochemical reactions through a gas–liquid–solid three-phase interface [106]. Its core structure typically comprises four components: the gas diffusion layer (GDL), microporous layer (MPL), catalyst layer (CL), and current collector. Gas diffusion layer (Figure 10a): Constructed from porous hydrophobic materials (e.g., carbon fiber paper or polytetrafluoroethylene (PTFE) membranes), it ensures efficient gas transport and prevents electrolyte penetration through its pore structure. Microporous layer: A mixture of carbon powder and PTFE forms a hydrophobic and conductive intermediate layer, further refining pore structures, enhancing adhesion between the catalyst layer and GDL, and improving interfacial stability. Catalyst layer: Catalysts are loaded onto carbon supports via spraying, electrodeposition, or electrospinning. Current collector: A conductive metal mesh (e.g., nickel mesh) or carbon cloth provides electrical conduction and mechanical support. The GDE enables high-efficiency reactions by optimizing the three-phase interface (gas–electrolyte–catalyst). In H2O2 electrocatalytic synthesis, O2 is directly conveyed to the catalyst surface, where H2O2 is generated via the 2eORR pathway, significantly reducing mass transfer resistance. Upon reaching the catalyst layer, O2 forms a gas–liquid–solid three-phase interface with the electrolyte. At this interface, oxygen, ions in the electrolyte, and active sites on the electrode surface interact, driving complex electrochemical reactions. Under the influence of the electric field, ions in the electrolyte migrate to the catalyst layer surface and participate in the reaction with O2 to produce H2O2. Zhang et al. [58] constructed the cathode of a three-phase flow cell device using a gas diffusion layer (GDL) made of carbon paper and a microporous layer (MPL) loaded with Co-N5/NC catalyst. With 1.0 M KOH as the electrolyte in O2-saturated alkaline conditions, O2 was reduced to H2O2 via 2eORR, achieving a production rate of 16.1 mol gcat−1h−1 at 150 mA cm−2. Operating at 0.5 V vs. RHE for 20 h, the accumulated H2O2 concentration reached 427.7 mM (14,540 ppm) with a Faradaic efficiency (FE) >90%. Liu et al. [76] coated a Co-C/N/O catalyst onto a gas diffusion electrode and assembled a two-compartment flow cell separated by an NR-212 cation-exchange membrane to separate the anode and cathode chambers. In 1 M phosphate-buffered saline (PBS), galvanostatic electrolysis at 20 mA cm−2 for 5 h yielded an H2O2 production rate of 4.72 mol gcat−1h−1. After 5000 cycles of accelerated durability testing, the activity declined by only 20–39 mV, retaining stable H2O2 selectivity.
A flow cell system is an electrochemical device utilizing gas diffusion electrodes (GDEs) and electrocatalytic materials to enable electrocatalytic O2 reduction at the gas–liquid–solid three-phase interface for H2O2 generation [108]. The core structure of the flow cell comprises the GDE (cathode), anode, flowing electrolyte system, membrane materials, and external circulation and storage systems (Figure 10b). The GDE, serving as the cathode, directly supplies gaseous O2 to the reaction interface, forming a three-phase interface (gas–liquid–solid) and significantly enhancing oxygen mass transfer efficiency. The flowing electrolyte system circulates the electrolyte via pumps to mitigate concentration polarization and rapidly remove generated H2O2, preventing its further reduction to water or decomposition. Commonly used membrane materials include proton-exchange membranes (PEMs) or anion-exchange membranes (AEMs). External circulation and storage systems enable continuous electrolyte flow, regulating reactant concentration and flow rate. To enhance H2O2 production efficiency, designing superior catalysts remains a critical challenge, particularly in acidic environments where ORR tends to favor complete reduction to H2O, necessitating precise structural tuning of catalysts to maintain H2O2 selectivity. Zhang et al. [63] achieved an H2O2 production rate exceeding 2000 mmol gcat−1h−1 with a Faradaic efficiency (FE) >90% in an acidic electrolyte (0.1 M HClO4) using a flow cell assembled with CoSA-N-C/CNTs, maintaining stable performance for 100 h without decay. In alkaline environments, however, challenges such as insufficient catalyst activity or poor membrane stability necessitate improvements in catalyst performance and membrane durability to enhance H2O2 yields. Similarly, Han et al. [55] demonstrated that the Co-N5-O-C catalyst delivered an H2O2 production rate of 11.3 mol g−1h−1 at 200 mA cm−2 in 1M KOH, retaining ~80% FE after 24 h of continuous operation at 100 mA cm−2.
Future research should focus on multiscale collaborative design, spanning from atomic-level modulation of catalyst electronic structures to optimization of reactor architectures, integrated with computational materials science, to achieve low-cost, high-concentration, and fully green process chains for electrochemical H2O2 synthesis, ultimately revolutionizing traditional chemical production paradigms.

4.2. Zinc–Air Battery

Aqueous zinc–air batteries are a novel category of chemical power source that utilize metallic zinc as the anode and oxygen from air as the cathode active material, combining characteristics of both secondary batteries and fuel cells [109]. These batteries typically consist of an air electrode, an alkaline or neutral electrolyte, a separator, and a zinc anode. During discharge, zinc at the anode is oxidized in the alkaline electrolyte to form zincate ions while releasing electrons. At the cathode, oxygen from the air is adsorbed onto the air diffusion electrode and reduced in the electrolyte to generate hydroxide ions (OH), which react with zincate ions to produce zinc oxide (ZnO) and H2O. Electrons flow from the anode to the cathode through an external circuit, generating electrical current. During charging, the anode reaction involves the oxidation of ZnO to Zn2+ and OH by losing electrons, while the cathode reaction reduces water to produce oxygen and OH (Figure 10c). However, aqueous zinc–air batteries suffer from sluggish ORR kinetics, necessitating highly efficient catalysts to reduce overpotentials and enhance reaction rates [110]. When catalysts with metal–nonmetal asymmetric atomic structures are applied, they exhibit superior performance compared to commercial Pt/C catalysts, offering improved stability and catalytic activity for practical applications. Catalysts can be applied in two forms: the first is their direct use as the cathode for catalytic reactions. Zhang et al. [84] employed Co-N3-C/CNT as the air cathode, where the zinc–air battery demonstrated a maximum power density of 159.3 mW cm−2 and a high specific capacity of 892.6 mAh g−1, with no performance decay after 8000 cycles. Yang et al. [86] applied Fe-N3S/Cmeso as the cathode material in a zinc–air battery, attaining an open-circuit voltage of 1.45 V, a peak power density of 178 mW cm−2, and stability exceeding 350 h, with 96.17% capacity retention over 100 h. Ren et al. [111] assembled a zinc–air battery using an asymmetric N,P-coordinated Co single-atom catalyst (d-CoN3P) as the cathode, delivering a specific capacity of 842 mAh g−1 and a power density of 175.7 mW cm−2, exceeding Pt/C (662 mAh g−1 and 123.8 mW cm−2), and retained stable performance after 1163 charge–discharge cycles. The second approach involves loading catalysts onto the cathode for catalytic reactions, where the loaded catalysts appear to outperform those directly used as cathodes. For example, Ding et al. [83] deposited a Co1-BNG catalyst onto the anode to form a composite electrode, achieving a zinc–air battery with a maximum power density of 253 mW cm−2 and durability exceeding 110 h. Jiang et al. [71] applied S-Cu-ISA/SNC as the air cathode catalyst in a zinc–air battery, demonstrating a maximum power density of 225 mW cm−2 and negligible voltage decay during 50 h cycling tests, indicating exceptional discharge stability. Lin et al. [88] employed the Sn-N/O-C catalyst in a ZAB, achieving a maximum power density of 254 mW cm−2, a specific capacity of 781 mAh g−1, and stable performance over 350 h of discharge testing (Figure 10d). Metal–metal asymmetric-atomic-structure catalysts further enhance the performance of aqueous zinc–air batteries, with bimetallic synergistic catalysis exhibiting stronger performance than metal–nonmetal catalysts, yielding higher peak power densities. For instance, Shang et al. [91] applied the SeFe–C2N catalyst as the cathode in a ZAB, achieving a high power density of 287.2 mW cm−2 and a specific capacity of 764.8 mAh g−1, with virtually imperceptible voltage decay over 380 h of charge–discharge cycling. Wang et al. [107] integrated a Co/Cu-N-C catalyst into a ZAB, attaining a peak power density of 256.1 mW cm−2 and cycling stability exceeding 500 h.
Flexible solid-state zinc–air batteries (ZABs) are novel energy storage devices characterized by high theoretical energy density, augmented safety, and flexible bendability. Their working principle aligns with aqueous ZABs, with the core innovation being the replacement of traditional liquid electrolytes with solid-state electrolytes and the achievement of flexibility through material and structural design [112]. The key components of a flexible solid-state ZAB include an air cathode, zinc anode, solid-state electrolyte layer, and flexible encapsulation layer. The air cathode facilitates ORR. The zinc anode combines a metallic zinc layer with a flexible substrate to enable bending. The solid-state electrolyte layer conducts ions, prevents short circuits by isolating electrodes, and provides mechanical flexibility and strength. The flexible encapsulation layer is typically made of polymeric materials to block moisture ingress and maintain mechanical flexibility [113]. Cathode catalyst deactivation significantly degrades ORR kinetics and is a primary cause of capacity decay in flexible solid-state ZABs. For instance, Wang et al. [107] synthesized a Co/Cu-N-C catalyst that demonstrated a high specific capacity of 727.1 mAh g−1 and stable power density of 113.2 mW cm−2 under bending conditions in flexible ZABs (Figure 10e), outperforming Pt/C-based batteries (21.7 mW cm−2). Lv et al. [97] applied the Mn-N-HsGDY catalyst in a 1D solid-state ZAB, achieving an open-circuit voltage of 1.363 V and maintaining stable voltage output during bending to power a timer. In a 2D solid-state ZAB, it delivered a peak power density of 156.6 mW cm−2, surpassing most reported flexible ZABs, and preserved stable voltage differences throughout 1400 min of charge–discharge testing.

4.3. Hydrogen–Oxygen Fuel Cell

Proton-exchange membrane fuel cells (PEMFCs) constitute highly efficient and clean energy conversion devices that directly transform chemical energy into electricity through hydrogen–oxygen electrochemical reactions [114]. Among these, the ORR at the cathode is one of the most critical processes. A PEMFC consists of gas diffusion layers (GDLs), membrane electrode assemblies (MEAs), and bipolar plates. The MEA, which includes catalyst layers (CLs) and a proton-exchange membrane (PEM), serves as the core of the electrochemical conversion process. The performance and efficiency of devices relying on this process depend heavily on the properties of ORR electrocatalysts and PEMs [115]. Thus, developing higher-activity ORR catalysts and cost-effective MEAs with high durability is essential for advancing these energy conversion technologies [116]. Fundamentally, PEMFCs operate as the reverse process of water electrolysis: H2 and O2 react in the presence of catalysts to produce H2O while releasing electrical energy. In the operation of PEMFCs, H2 enters the anode, where it splits into H+ and e under the action of catalysts. The H+ ions migrate through the PEM to the cathode, while the e are forced through an external circuit, generating an electric current to power electronic devices. At the cathode, O2 combines with the migrating H+ and incoming e to drive the ORR, producing H2O. To optimize PEMFC performance, high-performance ORR catalysts are essential. Catalyst particles must be integrated on both sides of the PEM to minimize the transport distances of H+ and e, thereby enhancing reaction kinetics. The chemical stability of the catalysts and their compatibility with the membrane are critical to preventing degradation of both components and maintaining optimal performance. Due to the constrained availability and high cost of noble metal-based catalysts, asymmetric-atomic-structure catalysts have emerged as promising alternatives to Pt-based catalysts. For example, Liu et al. [89] reported a Fe-N/O-C cathode catalyst that achieved a peak power density of 1179 mW cm−2 in PEMFCs (Figure 10f). Yin et al. [85] developed a CoN3S-PC cathode catalyst that delivered a peak power density surpassing 1.32 W cm−2 under O2 at 80 °C. Operating at a constant voltage of 0.6 V, the CoN3S-PC cathode maintained 53.3% of its current density after 37 h (ultimately stabilizing at approximately 300 mA cm−2). Chang et al. [92] synthesized a (Fe,Co)/N-C catalyst that achieved a peak power density of 505 mW cm−2 in PEMFCs (353 K with 0.2 MPa backpressure), reaching 76% of the power density of commercial Pt/C. Additionally, in H2/air fuel cells, it demonstrated negligible voltage decay when operated continuously for 100 h at constant current densities (600 and 1000 mA cm−2). These advancements highlight the potential of metal–nonmetal asymmetric coordination catalysts to outperform Pt/C in PEMFC applications, driving progress toward cost-effective and durable fuel cell technologies.

5. Conclusions and Prospects

In conclusion, asymmetric atomic structures have demonstrated exceptional performance in the ORR, showing practical applications in high-efficiency H2O2 production, metal–air batteries, and hydrogen–oxygen fuel cells, and have materialized as one of the most promising alternatives to traditional noble metal catalysts. Based on recent advancements in theoretical and experimental research, key design principles for developing next-generation high-performance asymmetric-atomic-structure catalysts have been summarized, providing critical guidance for optimizing catalytic activity and selectivity.
(i)
Design Principles for Metal Element Doping: In ORR catalysts, transition metal elements are the predominant choice for constructing asymmetric atomic structures owing to their unique electronic configurations and chemical properties. The partially filled d-orbitals of transition metals can hybridize with the antibonding *π orbitals of O2, modulating the activation energy for O-O bond cleavage. The position of the d-band center critically governs the adsorption strength of oxygen intermediates (e.g., *OOH, *O, *OH). Furthermore, transition metals exhibit multiple oxidation states (e.g., Fe2+/Fe3+, Co2+/Co3+), empowering reversible electron transfer during ORR. This redox flexibility allows their oxidation states to dynamically adapt across reaction steps, thereby facilitating catalytic progression. Additionally, the atomic size and electronegativity of transition metals promote the formation of asymmetric coordination structures with other atoms or ligands. Such configurations induce asymmetric charge distribution and electronic effects, optimizing the adsorption and reactivity toward ORR intermediates. Finally, transition metals possess high chemical and thermal stability, enabling sustained catalytic activity under ORR operating conditions (e.g., acidic/alkaline environments, high potentials) while resisting structural degradation or corrosion.
(ii)
Design Principles for Nonmetal Element Doping: In the fabrication of asymmetric-atomic-structure catalysts, the incorporation of high-electronegativity nonmetal elements induces charge polarization through their strong electron-withdrawing capability, creating localized electron density differences. This effectively modulates the electronic structure of the catalyst, enhancing its adsorption capacity for reactants and catalytic activity. Additionally, synergistic co-doping of multiple high-electronegativity nonmetal elements can be considered to achieve more significant catalytic performance improvements. Interactions between different dopants may collectively modulate the electronic and geometric structures of the catalyst, generating synergistic effects. For low-electronegativity nonmetal elements, their weak electron-withdrawing or electron-donating capabilities can modulate the charge distribution, surface reactive sites, and intermediate adsorption behavior of the catalyst structure. Selecting elements with atomic radii matching the substrate can reduce the risk of lattice distortion and preserves structural stability.
(iii)
Design Principles for Other Dopants: In asymmetric-atomic-structure ORR catalysts, defect engineering amplifies catalytic activity and selectivity by introducing vacancies, edge sites, and other asymmetric configurations. These defects modulate electronic distribution, expose active sites, and optimize reaction pathways. Defects instigate localized states near the Fermi level, reducing the energy gap between O2’s d-orbitals and the catalyst’s electronic states (e.g., carbon vacancies shift the Fermi level downward by 0.2–0.5 eV), thereby enhancing charge transfer efficiency. Finally, defects in porous structures expose additional active sites (e.g., defects in mesoporous carbon walls enhance mass transport and active site density).
(iv)
In summary, for the 2e pathway in the ORR, elements that weakly activate the O-O bond should be selected to reduce O-O bond cleavage capability, stabilize the *OOH intermediate, and retain the O-O bond, thereby enhancing H2O2 selectivity. The asymmetric configurations composed of N and O atoms in structures such as Co-N5, Zn-N3O and other structures promote the two-electron pathway of the oxygen reduction reaction. For the 4e pathway, the O-O bond cleavage capability must be strengthened, and active sites should integrate dual functions of O2 adsorption and O-O bond scission to promote rapid adsorption/desorption equilibrium of oxygen-containing intermediates (*O, *OH). The presence of B and Fe elements in structures such as Co-N3B and Fe-B-Co/NC promotes the four-electron pathway of the oxygen reduction reaction. Additionally, precise regulation of defect type, location, and concentration can optimize the electronic structure and reaction kinetics of active sites, enabling directional selection of ORR pathways (2e/4e). Finally, the choice of carbon matrix must comprehensively consider the synergistic effects of multi-element doping, defect engineering, hierarchical structures, and coordination modulation to achieve electronic structure optimization, rapid mass transport, and efficient exposure of active sites.

Author Contributions

Resources, H.Q.; conceptualization, S.W.; writing—original draft preparation, H.Q.; writing—review and editing, Q.F. and X.Z.; supervision, X.Z.; funding acquisition, X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Colleges and Universities Twenty Terms Foundation of Jinan City (No. 202228053) and the Science, Education, and Industry Integration Innovation Pilot Project of Qilu University of Technology (Shandong Academy of Sciences) (No. 2024RCKY019).

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fu, J.; Cano, Z.P.; Park, M.G.; Yu, A.; Fowler, M.; Chen, Z. Electrically Rechargeable Zinc–Air Batteries: Progress, Challenges, and Perspectives. Adv. Mater. 2016, 29, 1604685. [Google Scholar] [CrossRef] [PubMed]
  2. Fu, J.; Liang, R.; Liu, G.; Yu, A.; Bai, Z.; Yang, L.; Chen, Z. Recent Progress in Electrically Rechargeable Zinc–Air Batteries. Adv. Mater. 2018, 31, 1805230. [Google Scholar] [CrossRef] [PubMed]
  3. Haider, R.; Wen, Y.; Ma, Z.-F.; Wilkinson, D.P.; Zhang, L.; Yuan, X.; Song, S.; Zhang, J. High temperature proton exchange membrane fuel cells: Progress in advanced materials and key technologies. Chem. Soc. Rev. 2021, 50, 1138–1187. [Google Scholar] [CrossRef]
  4. Khan, M.A.; Al-Attas, T.; Roy, S.; Rahman, M.M.; Ghaffour, N.; Thangadurai, V.; Larter, S.; Hu, J.; Ajayan, P.M.; Kibria, M.G. Seawater electrolysis for hydrogen production: A solution looking for a problem? Energy Environ. Sci. 2021, 14, 4831–4839. [Google Scholar] [CrossRef]
  5. Shi, L.; Liu, D.; Lin, X.; Cheng, R.; Liu, F.; Kim, C.; Hu, C.; Qiu, J.; Amal, R.; Dai, L. Stable and High-performance Flow H2-O2 Fuel Cells with Coupled Acidic Oxygen Reduction and Alkaline Hydrogen Oxidation Reactions. Adv. Mater. 2024, 36, 2314077. [Google Scholar] [CrossRef]
  6. Wang, Q.; Kaushik, S.; Xiao, X.; Xu, Q. Sustainable zinc–air battery chemistry: Advances, challenges and prospects. Chem. Soc. Rev. 2023, 52, 6139–6190. [Google Scholar] [CrossRef]
  7. Xiao, F.; Wang, Y.C.; Wu, Z.P.; Chen, G.; Yang, F.; Zhu, S.; Siddharth, K.; Kong, Z.; Lu, A.; Li, J.C.; et al. Recent Advances in Electrocatalysts for Proton Exchange Membrane Fuel Cells and Alkaline Membrane Fuel Cells. Adv. Mater. 2021, 33, 2006292. [Google Scholar] [CrossRef] [PubMed]
  8. Yu, Z.; Liu, L. Recent Advances in Hybrid Seawater Electrolysis for Hydrogen Production. Adv. Mater. 2023, 36, 2308647. [Google Scholar] [CrossRef]
  9. Zhang, W.; Liu, M.; Gu, X.; Shi, Y.; Deng, Z.; Cai, N. Water Electrolysis toward Elevated Temperature: Advances, Challenges and Frontiers. Chem. Rev. 2023, 123, 7119–7192. [Google Scholar] [CrossRef]
  10. 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]
  11. Xu, S.; Lu, R.; Sun, K.; Tang, J.; Cen, Y.; Luo, L.; Wang, Z.; Tian, S.; Sun, X. Synergistic Effects in N,O-Comodified Carbon Nanotubes Boost Highly Selective Electrochemical Oxygen Reduction to H2O2. Adv. Sci. 2022, 9, 2201421. [Google Scholar] [CrossRef] [PubMed]
  12. Zhou, S.; Huang, S.; Wang, X.; Chen, S.; Wang, L.; Zhang, P.; Zhao, C. Activating the PdN4 single-atom sites for 4 electron oxygen reduction reaction via axial oxygen ligand modification. Chem. Eng. J. 2023, 472, 145129. [Google Scholar] [CrossRef]
  13. Tian, X.; Lu, X.F.; Xia, B.Y.; Lou, X.W. Advanced Electrocatalysts for the Oxygen Reduction Reaction in Energy Conversion Technologies. Joule 2020, 4, 45–68. [Google Scholar] [CrossRef]
  14. Zhai, Z.; Wang, Y.-J.; Pan, L.; Huang, F.; Liu, D.; Wang, B. Accelerating O-O bond dissociation in oxygen reduction reaction on the sp3-hybridized carbon. Appl. Surf. Sci. 2025, 691, 162668. [Google Scholar] [CrossRef]
  15. Yan, H.-M.; Wang, G.; Lv, X.-M.; Cao, H.; Qin, G.-Q.; Wang, Y.-G. Revealing the Potential-Dependent Rate-Determining Step of Oxygen Reduction Reaction on Single-Atom Catalysts. J. Am. Chem. Soc. 2025, 147, 3724–3730. [Google Scholar] [CrossRef]
  16. Cheng, J.; Lyu, C.; Li, H.; Wu, J.; Hu, Y.; Han, B.; Wu, K.; Hojamberdiev, M.; Geng, D. Steering the oxygen reduction reaction pathways of N-carbon hollow spheres by heteroatom doping. Appl. Catal. B Environ. 2023, 327, 122470. [Google Scholar] [CrossRef]
  17. Kumar, K.; Dubau, L.; Jaouen, F.; Maillard, F. Review on the Degradation Mechanisms of Metal-N-C Catalysts for the Oxygen Reduction Reaction in Acid Electrolyte: Current Understanding and Mitigation Approaches. Chem. Rev. 2023, 123, 9265–9326. [Google Scholar] [CrossRef]
  18. Wolker, T.; Brunnengräber, K.; Martinaiou, I.; Lorenz, N.; Zhang, G.-R.; Kramm, U.I.; Etzold, B.J.M. The effect of temperature on ionic liquid modified Fe-N-C catalysts for alkaline oxygen reduction reaction. J. Energy Chem. 2022, 68, 324–329. [Google Scholar] [CrossRef]
  19. Du, Y.X.; Yang, Q.; Lu, W.T.; Guan, Q.Y.; Cao, F.F.; Zhang, G. Carbon Black-Supported Single-Atom Co-N-C as an Efficient Oxygen Reduction Electrocatalyst for H2O2 Production in Acidic Media and Microbial Fuel Cell in Neutral Media. Adv. Funct. Mater. 2023, 33, 2300895. [Google Scholar] [CrossRef]
  20. Lim, C.; Fairhurst, A.R.; Ransom, B.J.; Haering, D.; Stamenkovic, V.R. Role of Transition Metals in Pt Alloy Catalysts for the Oxygen Reduction Reaction. ACS Catal. 2023, 13, 14874–14893. [Google Scholar] [CrossRef]
  21. Valério Neto, E.S.; Almeida, C.V.S.; Colmati, F.; Ciapina, E.G.; Salazar-Banda, G.R.; Eguiluz, K.I.B. Pt nanowires as electrocatalysts for proton-exchange membrane fuel cells applications: A review. J. Electroanal. Chem. 2022, 910, 116185. [Google Scholar] [CrossRef]
  22. Yin, S.; Song, Y.; Liu, H.; Cui, J.; Liu, Z.; Li, Y.; Xue, T.; Tang, W.; Zhang, D.; Li, H.; et al. Well-Defined PtCo@Pt Core-Shell Nanodendrite Electrocatalyst for Highly Durable Oxygen Reduction Reaction. Small 2025, 21, 2410080. [Google Scholar] [CrossRef] [PubMed]
  23. Chong, L.; Zhou, H.; Kubal, J.; Tang, Q.; Wen, J.; Yang, Z.; Bloom, I.D.; Abraham, D.; Zhu, H.; Zou, J.; et al. Highly durable fuel cell electrocatalyst with low-loading Pt-Co nanoparticles dispersed over single-atom Pt-Co-N-graphene nanofiber. Chem Catal. 2023, 3, 100541. [Google Scholar] [CrossRef]
  24. Bhuvanendran, N.; Ravichandran, S.; Lee, S.; Sanij, F.D.; Kandasamy, S.; Pandey, P.; Su, H.; Lee, S.Y. Recent progress in Pt-based electrocatalysts: A comprehensive review of supported and support-free systems for oxygen reduction. Coord. Chem. Rev. 2024, 521, 216191. [Google Scholar] [CrossRef]
  25. Liu, H.; Zhao, J.; Li, X. Controlled Synthesis of Carbon-Supported Pt-Based Electrocatalysts for Proton Exchange Membrane Fuel Cells. Electrochem. Energy Rev. 2022, 5, 13. [Google Scholar] [CrossRef]
  26. Sun, Q.; Li, X.-H.; Wang, K.-X.; Ye, T.-N.; Chen, J.-S. Inorganic non-carbon supported Pt catalysts and synergetic effects for oxygen reduction reaction. Energy Environ. Sci. 2023, 16, 1838–1869. [Google Scholar] [CrossRef]
  27. Peng, J.; Tao, P.; Song, C.; Shang, W.; Deng, T.; Wu, J. Structural evolution of Pt-based oxygen reduction reaction electrocatalysts. Chin. J. Catal. 2022, 43, 47–58. [Google Scholar] [CrossRef]
  28. Fu, Y.; Zhao, Q.; Wei, Q.; Bowen, C.R.; Wong, W.-Y.; Yang, W. Non-precious metal-based single-atom catalysts for oxygen reduction reaction: Fundamentals and applications. Mater. Sci. Eng. R Rep. 2024, 160, 100822. [Google Scholar] [CrossRef]
  29. Moghaddam, M.S.; Kafshgari, M.S.; Bahari, A.; Kafshgari, L.A.; Jafari, A. Types, properties, and applications of non-precious oxygen reduction reaction electrocatalyst: A review. J. Energy Chem. 2025, 107, 305–344. [Google Scholar] [CrossRef]
  30. Ao, X.; Wang, H.; Zhang, X.; Wang, C. Atomically Dispersed Metal–Nitrogen–Carbon Catalysts for Acidic Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2025, 17, 2844–2862. [Google Scholar] [CrossRef]
  31. Yang, Z.; Chen, Y.; Zhang, S.; Zhang, J. Identification and Understanding of Active Sites of Non-Noble Iron-Nitrogen-Carbon Catalysts for Oxygen Reduction Electrocatalysis. Adv. Funct. Mater. 2023, 33, 2215185. [Google Scholar] [CrossRef]
  32. Xu, W.; Tang, H.; Gu, H.; Xi, H.; Wu, P.; Liang, B.; Liu, Q.; Chen, W. Research progress of asymmetrically coordinated single-atom catalysts for electrocatalytic reactions. J. Mater. Chem. A 2022, 10, 14732–14746. [Google Scholar] [CrossRef]
  33. Zhao, X.; Hao, Z.; Zhang, X.; Li, L.; Gao, Y.; Liu, L. Alkaline oxygen reduction/evolution reaction electrocatalysis: A critical review focus on orbital structure, non-noble metal catalysts, and descriptors. Chem. Eng. J. 2024, 497, 155005. [Google Scholar] [CrossRef]
  34. Yu, A.; Yang, Y. Atomically Dispersed Metal Catalysts for Oxygen Reduction Reaction: Two-Electron vs. Four-Electron Pathways. Angew. Chem. Int. Ed. 2025, 64, e202424161. [Google Scholar] [CrossRef]
  35. Zhu, P.; Xiong, X.; Wang, X.; Ye, C.; Li, J.; Sun, W.; Sun, X.; Jiang, J.; Zhuang, Z.; Wang, D.; et al. Regulating the FeN4 Moiety by Constructing Fe–Mo Dual-Metal Atom Sites for Efficient Electrochemical Oxygen Reduction. Nano Lett. 2022, 22, 9507–9515. [Google Scholar] [CrossRef]
  36. Buschermöhle, J.G.; Müller-Hülstede, J.; Schmies, H.; Schonvogel, D.; Zierdt, T.; Lucka, R.; Renz, F.; Wagner, P.; Wark, M. Fe–Sn–N–C Catalysts: Advancing Oxygen Reduction Reaction Performance. ACS Catal. 2025, 15, 4477–4488. [Google Scholar] [CrossRef]
  37. Smith, L.A.; Burrow, J.N.; Eichler, J.E.; Tang, F.; Lauro, S.N.; Zhan, X.; Warner, J.H.; Mullins, C.B. A Deep Dive Into the Study of Nitrogen-Doped Carbons as Electrocatalysts for the Oxygen Reduction Reaction via Design of Experiments. Small 2025, 21, 2410010. [Google Scholar] [CrossRef]
  38. Zong, L.; Fan, K.; Cui, L.; Lu, F.; Liu, P.; Li, B.; Feng, S.; Wang, L. Constructing Fe-N4 Sites through Anion Exchange-mediated Transformation of Fe Coordination Environments in Hierarchical Carbon Support for Efficient Oxygen Reduction. Angew. Chem. Int. Ed. 2023, 62, e202309784. [Google Scholar] [CrossRef]
  39. Wang, X.; Yang, C.; Wang, X.; Zhu, H.; Cao, L.; Chen, A.; Gu, L.; Zhang, Q.; Zheng, L.; Liang, H.-P. Green Synthesis of a Highly Efficient and Stable Single-Atom Iron Catalyst Anchored on Nitrogen-Doped Carbon Nanorods for the Oxygen Reduction Reaction. ACS Sustain. Chem. Eng. 2020, 9, 137–146. [Google Scholar] [CrossRef]
  40. Zhu, Y.; Deng, F.; Qiu, S.; Ma, F.; Zheng, Y.; Lian, R. Enhanced electro-Fenton degradation of sulfonamides using the N, S co-doped cathode: Mechanism for H2O2 formation and pollutants decay. J. Hazard. Mater. 2021, 403, 123950. [Google Scholar] [CrossRef]
  41. Wu, L.; Cao, J.; Fu, C.; Chen, Z.; Yin, C.; Lin, Q.; Pan, H.; Wang, K. Lead regulation of electronic properties and local structure of palladium (111) facet for enhanced direct hydrogen peroxide synthesis. J. Colloid Interface Sci. 2025, 682, 1205–1216. [Google Scholar] [CrossRef]
  42. Liu, J.; Zhang, Z.; Han, C.; Xu, Z.; Zhu, Y.; Attfield, J.P.; Zeng, J.; Yang, M. Ligand-modified nickel nitride for natural seawater H2O2 synthesis. Appl. Catal. B Environ. Energy 2025, 373, 125362. [Google Scholar] [CrossRef]
  43. Musaie, K.; Abbaszadeh, S.; Marais, K.; Nosrati-Siahmazgi, V.; Rezaei, S.; Xiao, B.; Dua, K.; Santos, H.A.; Shahbazi, M.A. H2O2-Generating Advanced Nanomaterials for Cancer Treatment. Adv. Funct. Mater. 2025, 2425866. [Google Scholar] [CrossRef]
  44. Yin, J.; Zhang, H.; Luo, M.; Zhao, J.; Zhou, X.; Wang, F.; Zhou, H.; Yuan, Y.; Liu, Y.; Shi, Y.; et al. Electro-Fenton process for wastewater treatment: From selective H2O2 generation to efficient *OH conversion. Chem. Eng. J. 2025, 507, 160709. [Google Scholar] [CrossRef]
  45. Chen, K.-Y.; Huang, Y.-X.; Jin, R.-C.; Huang, B.-C. Single atom catalysts for use in the selective production of hydrogen peroxide via two-electron oxygen reduction reaction: Mechanism, activity, and structure optimization. Appl. Catal. B Environ. 2023, 337, 122987. [Google Scholar] [CrossRef]
  46. Bu, Y.; Wang, Y.; Han, G.F.; Zhao, Y.; Ge, X.; Li, F.; Zhang, Z.; Zhong, Q.; Baek, J.B. Carbon-Based Electrocatalysts for Efficient Hydrogen Peroxide Production. Adv. Mater. 2021, 33, 171264. [Google Scholar] [CrossRef]
  47. 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]
  48. Kong, D.; Wang, Y.; Yan, Y.; Li, Z.; Lu, Y.; Chen, G. Breaking planar constraints: Curvature-tailored Co-N-C electrocatalysts for high-efficiency proton exchange membrane H2O2 production. Appl. Catal. B Environ. Energy 2025, 376, 125454. [Google Scholar] [CrossRef]
  49. Yan, L.; Wang, C.; Wang, Y.; Wang, Y.; Wang, Z.; Zheng, L.; Lu, Y.; Wang, R.; Chen, G. Optimizing the binding of the *OOH intermediate via axially coordinated Co-N5 motif for efficient electrocatalytic H2O2 production. Appl. Catal. B Environ. 2023, 338, 123078. [Google Scholar] [CrossRef]
  50. Yuan, Y.; Ma, J.; Kang, B.; Lee, J.Y. Unraveling the Common Nature of O and S Doping in Improving Electrochemical O2 Reduction Reaction Performance of FeN4C. ACS Catal. 2025, 15, 4039–4050. [Google Scholar] [CrossRef]
  51. Hu, H.; Wang, J.; Liao, K.; Chen, Z.; Zhang, S.; Sun, B.; Wang, X.; Ren, X.; Lin, J.; Han, X. Clarifying the Active Structure and Reaction Mechanism of Atomically Dispersed Metal and Nonmetal Sites with Enhanced Activity for Oxygen Reduction Reaction. Adv. Mater. 2024, 37, 2416126. [Google Scholar] [CrossRef]
  52. Tang, X.; Zhang, Y.; Tang, S.; Lützenkirchen-Hecht, D.; Yuan, K.; Chen, Y. Structural and Chemical Origin of Dual-Atom Sites for Enhanced Oxygen Electroreduction. ACS Catal. 2024, 14, 13065–13080. [Google Scholar] [CrossRef]
  53. Li, Y.; Jiang, H.; Lin, L.; Sun, Z.; Sun, G. Single-atomic iron synergistic atom-cluster induce remote enhancement toward oxygen reduction reaction. J. Energy Chem. 2025, 102, 413–420. [Google Scholar] [CrossRef]
  54. Wu, L.; Wang, Y.; Shao, C.; Wang, L.; Li, B. Copper single atom-modulated functionalization of iron clusters on a porous carbon nanosheet for the oxygen reduction reaction. J. Mater. Chem. A 2025, 13, 5974–5986. [Google Scholar] [CrossRef]
  55. Lazaridis, T.; Stühmeier, B.M.; Gasteiger, H.A.; El-Sayed, H.A. Capabilities and limitations of rotating disk electrodes versus membrane electrode assemblies in the investigation of electrocatalysts. Nat. Catal. 2022, 5, 363–373. [Google Scholar] [CrossRef]
  56. Yu, Y.; Jiao, D.; Yi, L.; Zhao, D.; Zou, J.; Cui, X.; Hu, W. Unique electronic regulation in bimetallic MOF for efficient electrochemical synthesis of H2O2 at 500 mA cm−2. Nano Energy 2025, 141, 111096. [Google Scholar] [CrossRef]
  57. Xu, C.; Li, X.; Guo, P.-P.; Yang, K.-Z.; Zhao, Y.-M.; Chi, H.-M.; Xu, Y.; Wei, P.-J.; Wang, Z.-Q.; Xu, Q.; et al. Creating Asymmetric Fe–N3C–N Sites in Single-Atom Catalysts Boosts Catalytic Performance for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2024, 16, 37927–37937. [Google Scholar] [CrossRef]
  58. Yu, K.; Guan, S.; Zhang, W.; Zhang, W.; Meng, Y.; Lin, H.; Gao, Q. Engineering Asymmetric Electronic Structure of Co-N-C Single-Atomic Sites Toward Excellent Electrochemical H2O2 Production and Biomass Upgrading. Angew. Chem. Int. Ed. 2025, 64, e202502383. [Google Scholar] [CrossRef]
  59. Liu, C.; Yang, R.; Wang, J.; Liu, B.; Chang, X.; Feng, P.; Zhang, X.; Zhong, L.; Zhao, X.; Niu, L.; et al. Synergistic Catalysts with Fe Single Atoms and Fe3C Clusters for Accelerated Oxygen Adsorption Kinetics in Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2025, 64, e202501266. [Google Scholar] [CrossRef]
  60. Zhao, Y.; Chen, H.C.; Ma, X.; Li, J.; Yuan, Q.; Zhang, P.; Wang, M.; Li, J.; Li, M.; Wang, S.; et al. Vacancy Defects Inductive Effect of Asymmetrically Coordinated Single-Atom Fe-N3S1 Active Sites for Robust Electrocatalytic Oxygen Reduction with High Turnover Frequency and Mass Activity. Adv. Mater. 2023, 36, 2308243. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Gao, Z.; Zhang, S.; Guan, X.; Xu, W.; Liang, Y.; Jiang, H.; Li, Z.; Wu, S.; Cui, Z.; et al. Asymmetric-Charge-Distributed Co-Mn Diatomic Catalyst Enables Efficient Oxygen Reduction Reaction. Adv. Funct. Mater. 2025, 518, 2504260. [Google Scholar] [CrossRef]
  62. Zhao, Y.; Adiyeri Saseendran, D.P.; Huang, C.; Triana, C.A.; Marks, W.R.; Chen, H.; Zhao, H.; Patzke, G.R. Oxygen Evolution/Reduction Reaction Catalysts: From In Situ Monitoring and Reaction Mechanisms to Rational Design. Chem. Rev. 2023, 123, 6257–6358. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, Y.; Huang, M.; Kang, Y.; Fang, Y.; Zhao, T.; Wang, H.; Ou, J.; Liu, J.; Zhong, M.; Wang, T.; et al. Ligand effects enhancing low-temperature oxygen reduction kinetics in neutral conditions. Energy Environ. Sci. 2025, 18, 5298–5308. [Google Scholar] [CrossRef]
  64. Liao, X.; Lu, R.; Xia, L.; Liu, Q.; Wang, H.; Zhao, K.; Wang, Z.; Zhao, Y. Density Functional Theory for Electrocatalysis. Energy Environ. Mater. 2021, 5, 157–185. [Google Scholar] [CrossRef]
  65. Xie, L.; Zhou, W.; Huang, Y.; Qu, Z.; Li, L.; Yang, C.; Ding, Y.; Li, J.; Meng, X.; Sun, F.; et al. Elucidating the impact of oxygen functional groups on the catalytic activity of M–N4–C catalysts for the oxygen reduction reaction: A density functional theory and machine learning approach. Mater. Horiz. 2024, 11, 1719–1731. [Google Scholar] [CrossRef]
  66. Li, X.; Jiao, D.; Zhao, X. A Comprehensive Descriptor for Understanding High-Shell Heteroatom-Tuned Oxygen Reduction Reaction Activity on Diatomic FeCoN6 Sites. Renewables 2024, 2, 242–249. [Google Scholar] [CrossRef]
  67. Chen, K.; Huang, J.; Chen, J.; Gao, J.; Lu, Z.; Liu, X.; Lan, S.; Jia, G.; Ci, S.; Wen, Z. Neighboring Iron Single Atomic Sites Boost PtCo Intermetallic for High-Durability ORR Electrocatalysis. Energy Environ. Sci. 2025. [Google Scholar] [CrossRef]
  68. Yuan, M.; Liu, Y.; Du, Y.; Xiao, Z.; Li, H.; Liu, K.; Wang, L. Dual-Shelled Hollow Leafy Carbon Support with Atomically Dispersed (N,S)-Bridged Hydroxy-Coordinated Asymmetric Fe Sites for Oxygen Reduction. Adv. Funct. Mater. 2024, 34, 2401484. [Google Scholar] [CrossRef]
  69. Genz, N.S.; Kallio, A.J.; Oord, R.; Krumeich, F.; Pokle, A.; Prytz, Ø.; Olsbye, U.; Meirer, F.; Huotari, S.; Weckhuysen, B.M. Operando Laboratory-Based Multi-Edge X-Ray Absorption Near-Edge Spectroscopy of Solid Catalysts. Angew. Chem. Int. Ed. 2022, 61, e202209334. [Google Scholar] [CrossRef]
  70. Miao, C.; Xu, S.; An, Z.; Pan, X.; Li, Y.; Hu, N.; Li, L.; Zhou, Y.; Zhao, G. Self-Optimized Reconstruction of Metal–Organic Frameworks Introduces Cation Vacancies for Selective Electrosynthesis of Hydrogen Peroxide. Angew. Chem. Int. Ed. 2025, 64, e202501930. [Google Scholar] [CrossRef]
  71. Shang, H.; Zhou, X.; Dong, J.; Li, A.; Zhao, X.; Liu, Q.; Lin, Y.; Pei, J.; Li, Z.; Jiang, Z.; et al. Engineering unsymmetrically coordinated Cu-S1N3 single atom sites with enhanced oxygen reduction activity. Nat. Commun. 2020, 11, 3049. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, Y.; Rodenbücher, C.; Wippermann, K.; Korte, C. Revealing Interfacial Reactions on Pt Electrodes in Ionic Liquids by In Situ Fourier-Transform Infrared Spectroscopy. Anal. Chem. 2023, 95, 16618–16624. [Google Scholar] [CrossRef]
  73. Li, Z.M.; Zhang, C.Q.; Liu, C.; Zhang, H.W.; Song, H.; Zhang, Z.Q.; Wei, G.F.; Bao, X.J.; Yu, C.Z.; Yuan, P. High-efficiency Electroreduction of O2 into H2O2 over ZnCo Bimetallic Triazole Frameworks Promoted by Ligand Activation. Angew. Chem. Int. Ed. 2023, 63, e202314266. [Google Scholar] [CrossRef]
  74. Liu, B.; Guo, P.; Dai, Y.; Liu, B.; Zhang, Z.; Cai, J.; Xia, Y.; Pan, Q.; Shen, L.; Zhang, Y.; et al. Tailoring asymmetric atomic strain of FeN4 sites for enhanced acidic oxygen reduction reaction. Chem. Eng. J. 2025, 507, 160174. [Google Scholar] [CrossRef]
  75. Du, H.; Li, C.; Liang, Y.; Wu, Z.-S. Key strategies and challenging perspectives of carbon-based electrocatalysts for sustainable H2O2 production. J. Energy Chem. 2025, 106, 864–879. [Google Scholar] [CrossRef]
  76. Liu, L.; Kang, L.; Feng, J.; Hopkinson, D.G.; Allen, C.S.; Tan, Y.; Gu, H.; Mikulska, I.; Celorrio, V.; Gianolio, D.; et al. Atomically dispersed asymmetric cobalt electrocatalyst for efficient hydrogen peroxide production in neutral media. Nat. Commun. 2024, 15, 4079. [Google Scholar] [CrossRef]
  77. Wei, G.; Liu, X.; Zhao, Z.; Men, C.; Ding, Y.; Gao, S. Constructing ultrahigh-loading unsymmetrically coordinated Zn-N3O single-atom sites with efficient oxygen reduction for H2O2 production. Chem. Eng. J. 2023, 455, 140721. [Google Scholar] [CrossRef]
  78. Zhang, W.; Choi, J.W.; Kim, S.; Le, T.T.; Nandy, S.; Hwang, C.-K.; Paek, S.Y.; Byeon, A.; Chae, K.H.; Lee, S.Y.; et al. Penta nitrogen coordinated cobalt single atom catalysts with oxygenated carbon black for electrochemical H2O2 production. Appl. Catal. B Environ. 2023, 331, 122712. [Google Scholar] [CrossRef]
  79. Li, J.; Zheng, W.; Meng, S.; Wang, S.; Zhan, S.; Sun, Y.; Hu, W.; Li, Y. Modulated Fe Central Spin State in FeCu4 Alloy Nanoparticles for Efficient and Selective Activation of H2O2. Adv. Funct. Mater. 2024, 35, 2417538. [Google Scholar] [CrossRef]
  80. Yang, X.; Priest, C.; Hou, Y.; Wu, G. Atomically dispersed dual-metal-site PGM-free electrocatalysts for oxygen reduction reaction: Opportunities and challenges. SusMat 2022, 2, 569–590. [Google Scholar] [CrossRef]
  81. Hu, H.; Zhang, C.; Liu, W.; Qi, H.; Wang, H.; Wang, X.; Zhang, L.; Liu, L.; Bao, L.; Alomar, M.; et al. Coaxial Metal-Nitrogen–Carbon Single-Atom Catalysts Boost Acid Hydrogen Peroxide Production. Adv. Funct. Mater. 2024, 35, 2419220. [Google Scholar] [CrossRef]
  82. Zhang, S.; Sun, B.; Liao, K.; Wang, X.; Chen, Z.; Wang, J.; Hu, W.; Han, X. Boosting Oxygen Reduction Reaction Performance of Fe Single-Atom Catalysts Via Precise Control of the Coordination Environment. Adv. Funct. Mater. 2025, 2425640. [Google Scholar] [CrossRef]
  83. Guan, G.; Liu, Y.; Li, F.; Shi, X.; Liu, L.; Wang, T.; Xu, X.; Zhao, M.; Ding, J.; Yang, H.B. Atomic Cobalt Metal Centers with Asymmetric N/B-Coordination for Promoting Oxygen Reduction Reaction. Adv. Funct. Mater. 2024, 34, 2408111. [Google Scholar] [CrossRef]
  84. Tang, C.; Jin, R.; Xia, Y.; Zhang, Q.; Wang, J.; Wu, M.; Cui, X.; Yang, Y. Asymmetric low-coordination tailoring of single-atom cobalt catalysts enabling efficient oxygen reduction reaction. Nano Energy 2025, 137, 110776. [Google Scholar] [CrossRef]
  85. Chen, L.; Yin, S.; Zeng, H.; Liu, J.; Xiao, X.; Cheng, X.; Huang, H.; Huang, R.; Yang, J.; Lin, W.-F.; et al. Engineering asymmetric electronic structure of cobalt coordination on CoN3S active sites for high performance oxygen reduction reaction. J. Energy Chem. 2024, 98, 494–502. [Google Scholar] [CrossRef]
  86. Liu, C.; Yang, X.; Jia, Z.; Pan, J.; Zhao, F.; Ban, Y.; Cui, X.; Bai, Z.; Yang, L. Boosted electron accumulation around Fe atom by asymmetry coordinated FeN3S1 for efficient oxygen reduction reaction. Int. J. Hydrog. Energy 2024, 90, 1003–1011. [Google Scholar] [CrossRef]
  87. Liu, Y.; Pang, H.; Chen, M.; Ji, X.; Meng, F. Hollow polyhedral Fe single-atom catalysts with S/N dual coordination boosting oxygen and iodide redox for high-efficiency zinc–air/iodide hybrid batteries. J. Mater. Chem. A 2025, 13, 16008–16017. [Google Scholar] [CrossRef]
  88. Lin, X.; Zhang, X.; Liu, D.; Shi, L.; Zhao, L.; Long, Y.; Dai, L. Asymmetric Atomic Tin Catalysts with Tailored p-Orbital Electron Structure for Ultra-Efficient Oxygen Reduction. Adv. Energy Mater. 2024, 14, 2303740. [Google Scholar] [CrossRef]
  89. Li, Y.; Sun, H.; Ren, L.; Sun, K.; Gao, L.; Jin, X.; Xu, Q.; Liu, W.; Sun, X. Asymmetric Coordination Regulating D-Orbital Spin-Electron Filling in Single-Atom Iron Catalyst for Efficient Oxygen Reduction. Angew. Chem. Int. Ed. 2024, 63, e202405334. [Google Scholar] [CrossRef]
  90. Brea, C.; Hu, G. Mechanistic Insight into Dual-Metal-Site Catalysts for the Oxygen Reduction Reaction. ACS Catal. 2023, 13, 4992–4999. [Google Scholar] [CrossRef]
  91. Wang, X.; Zhang, N.; Shang, H.; Duan, H.; Sun, Z.; Zhang, L.; Lei, Y.; Luo, X.; Zhang, L.; Zhang, B.; et al. Precisely designing asymmetrical selenium-based dual-atom sites for efficient oxygen reduction. Nat. Commun. 2025, 16, 470. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, J.; Huang, Z.; Liu, W.; Chang, C.; Tang, H.; Li, Z.; Chen, W.; Jia, C.; Yao, T.; Wei, S.; et al. Design of N-Coordinated Dual-Metal Sites: A Stable and Active Pt-Free Catalyst for Acidic Oxygen Reduction Reaction. J. Am. Chem. Soc. 2017, 139, 17281–17284. [Google Scholar] [CrossRef] [PubMed]
  93. Lv, Q.; Li, M.; Li, X.; Yan, X.; Hou, Z.; Huang, C. Introducing hydroxyl groups to tailor the d-band center of Ir atom through side anchoring for boosted ORR and HER. J. Energy Chem. 2024, 90, 144–151. [Google Scholar] [CrossRef]
  94. Cao, Y.; Zeng, J.; Zheng, X.; Liu, Y.; Lu, J.; Zhang, J.; Wang, Y.; Deng, Y.; Hu, W. Regulating d-orbital spin state of Fe in single-atom electrocatalyst for boosting oxygen reduction activity in neutral electrolyte. J. Mater. Sci. Technol. 2025, 227, 67–75. [Google Scholar] [CrossRef]
  95. An, Q.; Zhang, X.; Yang, C.; Su, H.; Zhou, W.; Liu, M.; Zhang, X.; Sun, X.; Bo, S.; Yu, F.; et al. Engineering Unsymmetrically Coordinated Fe Sites via Heteroatom Pairs Synergetic Contribution for Efficient Oxygen Reduction. Small 2023, 19, 2304303. [Google Scholar] [CrossRef] [PubMed]
  96. Yang, M.; Qin, B.; Si, C.; Sun, X.; Li, B. Electrochemical reactions catalyzed by carbon dots from computational investigations: Functional groups, dopants, and defects. J. Mater. Chem. A 2024, 12, 2520–2560. [Google Scholar] [CrossRef]
  97. Li, M.; Hou, Z.; Li, X.; Gu, X.; Yan, X.; Lv, Q.; Huang, C. Asymmetrical anchor way of manganese atoms on carbon domain edge for enhanced oxygen reduction reaction. Appl. Catal. B Environ. Energy 2024, 356, 124249. [Google Scholar] [CrossRef]
  98. Bhardwaj, S.; Pathak, A.; Das, S.K.; Das, P.; Thapa, R.; Dey, R.S. Decoding Dual-Functionality in N-doped Defective Carbon: Unveiling Active Sites for Bifunctional Oxygen Electrocatalysis. Small 2025, 21, 2411035. [Google Scholar] [CrossRef]
  99. Yan, J.; Li, S.; Liu, M.; Zhou, H.; Su, H. Regulating electron effect by interface-induced dislocation in Fe2P/Fe to accelerate oxygen reduction reactions. Mater. Today Phys. 2025, 54, 101739. [Google Scholar] [CrossRef]
  100. Tu, H.; Zhang, H.; Song, Y.; Liu, P.; Hou, Y.; Xu, B.; Liao, T.; Guo, J.; Sun, Z. Electronic Asymmetry Engineering of Fe–N–C Electrocatalyst via Adjacent Carbon Vacancy for Boosting Oxygen Reduction Reaction. Adv. Sci. 2023, 10, 2305194. [Google Scholar] [CrossRef]
  101. Ji, S.; Wang, Y.; Liu, H.; Lu, X.; Guo, C.; Xin, S.; Horton, J.H.; Zhan, F.; Wang, Y.; Li, Z. Regulating the Electronic Synergy of Asymmetric Atomic Fe Sites with Adjacent Defects for Boosting Activity and Durability toward Oxygen Reduction. Adv. Funct. Mater. 2024, 34, 23146211. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Feng, J.; Shang, Z.; Pang, B.; Zhang, S.; Dong, H.; Yu, L.; Dong, L. Synergistic effects of metal single atoms and clusters on graphene-supported electrocatalysts: Insights into the mechanism of oxygen reduction reaction. Carbon 2025, 233, 119830. [Google Scholar] [CrossRef]
  103. Tian, Q.; Jing, L.; Wang, W.; Ye, X.; Chai, X.; Zhang, X.; Hu, Q.; Yang, H.; He, C. Hydrogen Peroxide Electrosynthesis via Selective Oxygen Reduction Reactions Through Interfacial Reaction Microenvironment Engineering. Adv. Mater. 2024, 37, 2414490. [Google Scholar] [CrossRef] [PubMed]
  104. Ma, R.; Han, G.F.; Li, F.; Bu, Y. Rational Design of Carbon-Based Electrocatalysts for H2O2 Production by Machine Learning and Structural Engineering. Adv. Energy Mater. 2025, 15, 2500953. [Google Scholar] [CrossRef]
  105. Wen, Y.; Zhang, T.; Wang, J.; Pan, Z.; Wang, T.; Yamashita, H.; Qian, X.; Zhao, Y. Electrochemical Reactors for Continuous Decentralized H2O2 Production. Angew. Chem. Int. Ed. 2022, 61, e202205972. [Google Scholar] [CrossRef]
  106. Oliveira Silva, T.; Granados-Fernández, R.; Lobato, J.; Lanza, M.R.V.; Rodrigo, M.A. Boosting New Electrochemical Reactor Designs to Improve the Performance in H2O2 Production Using Gas Diffusion Electrodes. ACS Sustain. Chem. Eng. 2025, 13, 3172–3182. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, L.; Sang, W.; Wang, T.; Wei, X.; Li, Z.; Chen, C.; Kou, Z.; Mu, S. Cu–N–C support confinement stabilizes active Co sites in oxygen reduction reaction. Nano Res. 2025, 18, 94907345. [Google Scholar] [CrossRef]
  108. Bao, Z.; Cao, Q.; Shao, Y.; Zhang, S.; Peng, X.; Li, Y.; Jiang, C.; Zhong, X.; Wang, J. Carbon and Oxygen-Doped Phosphorus Nitride (COPN) for Continuous Selective and Stable H2O2 Production. ACS Catal. 2023, 13, 14492–14502. [Google Scholar] [CrossRef]
  109. Mohd Shumiri, M.A.I.; Mohd Najib, A.S.; Fadil, N.A. Current status and advances in zinc anodes for rechargeable aqueous zinc-air batteries. Sci. Technol. Adv. Mater. 2025, 26, 2448418. [Google Scholar] [CrossRef]
  110. Tu, W.B.; Song, L.N.; Liang, S.; Wang, Y.; Sun, Y.; Wang, H.F.; Xu, J.J. Designing Hydrophobic Micro-Three-Phase Reaction Interfaces to Enhance the ORR Kinetics Toward Zinc-Air Battery. Adv. Energy Mater. 2025, 2404946. [Google Scholar] [CrossRef]
  111. Wang, D.; Li, X.; Ren, P.; Wei, Y.; Li, R.; Chen, Y.; Wang, W.; Yang, P.; Chen, Z. Asymmetric N, P-coordinated Co single-atom catalyst for efficient oxygen reduction reaction in Zn-air batteries. Appl. Catal. B Environ. Energy 2025, 375, 125436. [Google Scholar] [CrossRef]
  112. Zhang, Y.; Deng, Y.-P.; Wang, J.; Jiang, Y.; Cui, G.; Shui, L.; Yu, A.; Wang, X.; Chen, Z. Recent Progress on Flexible Zn-Air Batteries. Energy Storage Mater. 2021, 35, 538–549. [Google Scholar] [CrossRef]
  113. Wu, J.; Wu, W.-Y.; Wang, S.; Kai, D.; Ye, E.; Thitsartarn, W.; Tan, J.B.H.; Xu, J.; Yan, Q.; Zhu, Q.; et al. Polymer electrolytes for flexible zinc-air batteries: Recent progress and future directions. Nano Res. 2024, 17, 6058–6079. [Google Scholar] [CrossRef]
  114. Ahmed, S.; Beauger, C.; Zada, A.; Iqbal, W.; Ahmed, N.; Anwar, M.T.; Hassan, M. Recent advancements in designing high-performance proton exchange membrane fuel cells: A comprehensive review. Appl. Energy 2025, 390, 125753. [Google Scholar] [CrossRef]
  115. Zeng, W.; Guan, B.; Zhuang, Z.; Chen, J.; Zhu, L.; Ma, Z.; Hu, X.; Zhu, C.; Zhao, S.; Shu, K.; et al. Comprehensive review on the advances and comparisons of proton exchange membrane fuel cells (PEMFCs) and anion exchange membrane fuel cells (AFCs): From fundamental principles to key component technologies. Int. J. Hydrogen Energy 2025, 102, 222–246. [Google Scholar] [CrossRef]
  116. Sharma, R.; Morgen, P.; Chiriaev, S.; Lund, P.B.; Larsen, M.J.; Sieborg, B.; Grahl-Madsen, L.; Andersen, S.M. Insights into Degradation of the Membrane–Electrode Assembly Performance in Low-Temperature PEMFC: The Catalyst, the Ionomer, or the Interface? ACS Appl. Mater. Interfaces 2022, 14, 49658–49671. [Google Scholar] [CrossRef]
Figure 1. The two-electron and four-electron pathways of the oxygen reduction reaction under acidic and alkaline conditions.
Figure 1. The two-electron and four-electron pathways of the oxygen reduction reaction under acidic and alkaline conditions.
Catalysts 15 00615 g001
Figure 2. (a,b) Atomic configuration of adsorption of *OOH on Mo-O3S-C and Mo-S4-C. (c) Free energy diagrams of 2eORR on the three studied substrates at the reaction equilibrium potential. (d) Free energy diagrams of different Pt-Nx-C catalysts. (ad) Reproduced with permission [45]. Copyright 2023, Elsevier. (e) Non-doped pentagonal defect (Z5). (f) Doped Gr-N pentagonal defect (Z6). (g) Doped graphene edge (Z9). (h) Doped graphene edge (Z10). (eh) Reproduced with permission [47]. Copyright 2023, American Chemical Society.
Figure 2. (a,b) Atomic configuration of adsorption of *OOH on Mo-O3S-C and Mo-S4-C. (c) Free energy diagrams of 2eORR on the three studied substrates at the reaction equilibrium potential. (d) Free energy diagrams of different Pt-Nx-C catalysts. (ad) Reproduced with permission [45]. Copyright 2023, Elsevier. (e) Non-doped pentagonal defect (Z5). (f) Doped Gr-N pentagonal defect (Z6). (g) Doped graphene edge (Z9). (h) Doped graphene edge (Z10). (eh) Reproduced with permission [47]. Copyright 2023, American Chemical Society.
Catalysts 15 00615 g002
Figure 3. (a) Schematic diagram of the formation of the FeCo-N6-O structure from FeCo-N6 through two consecutive steps of OH electrochemical oxidation and the corresponding change in free energy. (b,c) The charge density difference plots of FeCo-N6 and FeCo-N6-O. (d,e) The PDOS of FeCo−N6 and FeCo−N6−O, respectively. The vertical red and blue lines in the bottom panel define the d-band center of Fe and Co atoms, respectively. (ae) Reproduced with permission [52]. Copyright 2024, American Chemical Society. (f,g) Structural models of FeSA-NC and FeSA/FeAC-NC. (f,g) Reproduced with permission [53]. Copyright 2024, Elsevier BV.
Figure 3. (a) Schematic diagram of the formation of the FeCo-N6-O structure from FeCo-N6 through two consecutive steps of OH electrochemical oxidation and the corresponding change in free energy. (b,c) The charge density difference plots of FeCo-N6 and FeCo-N6-O. (d,e) The PDOS of FeCo−N6 and FeCo−N6−O, respectively. The vertical red and blue lines in the bottom panel define the d-band center of Fe and Co atoms, respectively. (ae) Reproduced with permission [52]. Copyright 2024, American Chemical Society. (f,g) Structural models of FeSA-NC and FeSA/FeAC-NC. (f,g) Reproduced with permission [53]. Copyright 2024, Elsevier BV.
Catalysts 15 00615 g003
Figure 4. Research method of the oxygen reduction reaction pathway. (a) The RRDE technique was used to test the H2O2 yield and electron transfer of the C@PVI-(NCTPP)Fe-800 catalyst. (a) Reproduced with permission [56]. Copyright 2025, Elsevier BV. (b) The RRDE technique was used to test the H2O2 yield and the number of electron transfers of the BIM-Co2Zn8-500 catalyst. (b) Reproduced with permission [57]. Copyright 2024, American Chemical Society. In situ Raman spectra of the Fe/NSC-vd catalyst under the conditions of (c) 0.1M KOH and (d) 0.5M H2SO4. (c,d) Reproduced with permission [60]. Copyright 2023, Wiley-Blackwell. In situ attenuated total reflection surface-enhanced infrared absorption spectra of (e) the CoMn-NSC catalyst and (f) the CoMn-NC catalyst under 0.1 M KOH conditions. (e,f) Reproduced with permission [61]. Copyright 2025, Wiley-VCH Verlag.
Figure 4. Research method of the oxygen reduction reaction pathway. (a) The RRDE technique was used to test the H2O2 yield and electron transfer of the C@PVI-(NCTPP)Fe-800 catalyst. (a) Reproduced with permission [56]. Copyright 2025, Elsevier BV. (b) The RRDE technique was used to test the H2O2 yield and the number of electron transfers of the BIM-Co2Zn8-500 catalyst. (b) Reproduced with permission [57]. Copyright 2024, American Chemical Society. In situ Raman spectra of the Fe/NSC-vd catalyst under the conditions of (c) 0.1M KOH and (d) 0.5M H2SO4. (c,d) Reproduced with permission [60]. Copyright 2023, Wiley-Blackwell. In situ attenuated total reflection surface-enhanced infrared absorption spectra of (e) the CoMn-NSC catalyst and (f) the CoMn-NC catalyst under 0.1 M KOH conditions. (e,f) Reproduced with permission [61]. Copyright 2025, Wiley-VCH Verlag.
Catalysts 15 00615 g004
Figure 5. (a) The calculated ORR activity volcano plot. (b) The relationship between the distance of Co atoms from the center and the adsorption energy of *OOH. (a,b) Reproduced with permission [76]. Copyright 2024, Springer Nature. (c) Schematic diagram of the synthesis of Co-N5/NC and Co-N4/NC catalysts. (c) Reproduced with permission [58]. Copyright 2025, John Wiley and Sons Ltd. (d) Schematic diagram of the synthesis of Zn-N3O-SAC. (d) Reproduced with permission [77]. Copyright 2022, Elsevier.
Figure 5. (a) The calculated ORR activity volcano plot. (b) The relationship between the distance of Co atoms from the center and the adsorption energy of *OOH. (a,b) Reproduced with permission [76]. Copyright 2024, Springer Nature. (c) Schematic diagram of the synthesis of Co-N5/NC and Co-N4/NC catalysts. (c) Reproduced with permission [58]. Copyright 2025, John Wiley and Sons Ltd. (d) Schematic diagram of the synthesis of Zn-N3O-SAC. (d) Reproduced with permission [77]. Copyright 2022, Elsevier.
Catalysts 15 00615 g005
Figure 6. (a,b) Snapshots of the intermediate structures on the surfaces of ZnCo-ZIF (001) and ZnCo-MTF (001). (a,b) Reproduced with permission [73]. Copyright 2023, John Wiley and Sons Ltd. (c) Adsorption configurations of different intermediates in the two-electron and four-electron transfer ORR pathways on medium-cationic-vacancy catalysts. “*” represent the intermediate. (c) Reproduced with permission [70]. Copyright 2025, John Wiley and Sons Ltd. (d,e) The difference in charge density of CoSA-N-C and Co@Co-N-C before and after their interaction with *OOH. (d,e) Reproduced with permission [81]. Copyright 2024, Wiley-VCH Verlag.
Figure 6. (a,b) Snapshots of the intermediate structures on the surfaces of ZnCo-ZIF (001) and ZnCo-MTF (001). (a,b) Reproduced with permission [73]. Copyright 2023, John Wiley and Sons Ltd. (c) Adsorption configurations of different intermediates in the two-electron and four-electron transfer ORR pathways on medium-cationic-vacancy catalysts. “*” represent the intermediate. (c) Reproduced with permission [70]. Copyright 2025, John Wiley and Sons Ltd. (d,e) The difference in charge density of CoSA-N-C and Co@Co-N-C before and after their interaction with *OOH. (d,e) Reproduced with permission [81]. Copyright 2024, Wiley-VCH Verlag.
Catalysts 15 00615 g006
Figure 7. (a) The first shell fitting of the EXAFS spectra of Co1-BNG. (a) Reproduced with permission [83]. Copyright 2024, Wiley-VCH Verlag. (b) FT-EXAFS of Co-N3-C/CNT, Co-N4-C/CNT and reference materials. (c) Optimized structures of *O2, *OOH, *O and *OH on Co-N3. (d) The charge density difference of Co-N3. (bd) Reproduced with permission [84]. Copyright 2025, Elsevier BV. (e) FT-EXAFS fitting curve of the CoN3S-PC-1.4 catalyst. (f) The charge density difference at the CoN3S site. (e,f) Reproduced with permission [85]. Copyright 2024, Elsevier BV.
Figure 7. (a) The first shell fitting of the EXAFS spectra of Co1-BNG. (a) Reproduced with permission [83]. Copyright 2024, Wiley-VCH Verlag. (b) FT-EXAFS of Co-N3-C/CNT, Co-N4-C/CNT and reference materials. (c) Optimized structures of *O2, *OOH, *O and *OH on Co-N3. (d) The charge density difference of Co-N3. (bd) Reproduced with permission [84]. Copyright 2025, Elsevier BV. (e) FT-EXAFS fitting curve of the CoN3S-PC-1.4 catalyst. (f) The charge density difference at the CoN3S site. (e,f) Reproduced with permission [85]. Copyright 2024, Elsevier BV.
Catalysts 15 00615 g007
Figure 9. (a) Atomic configuration of FeSe-C2N. (a) Reproduced with permission [91]. Copyright 2025, Springer Nature. (b) The differential charge density map of CoN2S-MnN3-2OH. (c) Schematic diagram of the ORR process on the CoN2S-MnN3-2OH surface. (b,c) Reproduced with permission [61]. Copyright 2025, Wiley-VCH Verlag. (d) Correlation between Bader charge and the adsorption energy of hydroxyl (OH) groups of Ir atoms in IrO2, IrO2(OH)1, IrO2(OH)2 and IrO2(OH)3. (e) The charge density difference of IrO2(OH)4. (d,e) Reproduced with permission [93]. Copyright 2023, Elsevier BV. (f) The charge density of FeSA-Fe3C/NC. (f) Reproduced with permission [59]. Copyright 2025, John Wiley and Sons Ltd.
Figure 9. (a) Atomic configuration of FeSe-C2N. (a) Reproduced with permission [91]. Copyright 2025, Springer Nature. (b) The differential charge density map of CoN2S-MnN3-2OH. (c) Schematic diagram of the ORR process on the CoN2S-MnN3-2OH surface. (b,c) Reproduced with permission [61]. Copyright 2025, Wiley-VCH Verlag. (d) Correlation between Bader charge and the adsorption energy of hydroxyl (OH) groups of Ir atoms in IrO2, IrO2(OH)1, IrO2(OH)2 and IrO2(OH)3. (e) The charge density difference of IrO2(OH)4. (d,e) Reproduced with permission [93]. Copyright 2023, Elsevier BV. (f) The charge density of FeSA-Fe3C/NC. (f) Reproduced with permission [59]. Copyright 2025, John Wiley and Sons Ltd.
Catalysts 15 00615 g009
Figure 10. Application of oxygen reduction reaction. (a) Schematic diagram of H2O2 gas diffusion electrode reactor. (a) Reproduced with permission [58]. Copyright 2025, John Wiley and Sons Ltd. (b) Schematic diagram of the flow cell for producing H2O2. (b) Reproduced with permission [81]. Copyright 2024, Wiley-VCH Verlag. (c) Schematic diagram of the working principle of zinc–air batteries. (d) Polarization curves and corresponding power density diagrams of ZABs containing Sn-N/O-C and Pt/C. (c,d) Reproduced with permission [88]. Copyright 2024, Wiley-VCH Verlag. (e) Discharge polarization curve and power density of a solid-state flexible ZAB. (e) Reproduced with permission [107]. Copyright 2025, Tsinghua University Press. (f) Polarization curves and corresponding power density graphs of H2-O2 PEMFCs assembled with Fe-N/O-C, Fe-N-C, and Pt/C at 2.0 bar. (f) Reproduced with permission [89]. Copyright 2024, John Wiley and Sons Ltd.
Figure 10. Application of oxygen reduction reaction. (a) Schematic diagram of H2O2 gas diffusion electrode reactor. (a) Reproduced with permission [58]. Copyright 2025, John Wiley and Sons Ltd. (b) Schematic diagram of the flow cell for producing H2O2. (b) Reproduced with permission [81]. Copyright 2024, Wiley-VCH Verlag. (c) Schematic diagram of the working principle of zinc–air batteries. (d) Polarization curves and corresponding power density diagrams of ZABs containing Sn-N/O-C and Pt/C. (c,d) Reproduced with permission [88]. Copyright 2024, Wiley-VCH Verlag. (e) Discharge polarization curve and power density of a solid-state flexible ZAB. (e) Reproduced with permission [107]. Copyright 2025, Tsinghua University Press. (f) Polarization curves and corresponding power density graphs of H2-O2 PEMFCs assembled with Fe-N/O-C, Fe-N-C, and Pt/C at 2.0 bar. (f) Reproduced with permission [89]. Copyright 2024, John Wiley and Sons Ltd.
Catalysts 15 00615 g010
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

Qiu, H.; Wen, S.; Fu, Q.; Zhao, X. Oxygen Reduction Reactions of Catalysts with Asymmetric Atomic Structures: Mechanisms, Applications, and Challenges. Catalysts 2025, 15, 615. https://doi.org/10.3390/catal15070615

AMA Style

Qiu H, Wen S, Fu Q, Zhao X. Oxygen Reduction Reactions of Catalysts with Asymmetric Atomic Structures: Mechanisms, Applications, and Challenges. Catalysts. 2025; 15(7):615. https://doi.org/10.3390/catal15070615

Chicago/Turabian Style

Qiu, Hengxing, Shilong Wen, Qiuju Fu, and Xuebo Zhao. 2025. "Oxygen Reduction Reactions of Catalysts with Asymmetric Atomic Structures: Mechanisms, Applications, and Challenges" Catalysts 15, no. 7: 615. https://doi.org/10.3390/catal15070615

APA Style

Qiu, H., Wen, S., Fu, Q., & Zhao, X. (2025). Oxygen Reduction Reactions of Catalysts with Asymmetric Atomic Structures: Mechanisms, Applications, and Challenges. Catalysts, 15(7), 615. https://doi.org/10.3390/catal15070615

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