Next Article in Journal
The Development of Hexagonal Boron Nitride Crystal Growth Technologies and Their Applications in Neutron Detection
Previous Article in Journal
Interface Charge Transfer Engineering in NiFe Layered Double Hydroxide-Cs0.32WO3 Heterostructures for Enhanced Oxygen Evolution Reaction
Previous Article in Special Issue
Analysis of Transition from Compact to Mossy Structures During Galvanostatic Zinc Electrodeposition and Its Implications for CO2 Electroreduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Improving the Alkaline Oxygen Reduction Performance of Atomically Dispersed Metal–Nitrogen–Carbon Catalysts

1
School of Metallurgy and Environment, National Energy Metal Resources and New Materials Key Laboratory, Hunan Provincial Key Laboratory of Nonferrous Value-Added Metallurgy, Central South University, Changsha 410083, China
2
School of Materials Science and Engineering, Hunan University of Science and Technology, Xiangtan 411201, China
3
Research Institute of Renewable Energy and Advanced Materials, Zijin Mining Group Co., Ltd., Changsha 410013, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(16), 1257; https://doi.org/10.3390/nano15161257
Submission received: 26 July 2025 / Revised: 8 August 2025 / Accepted: 12 August 2025 / Published: 15 August 2025

Abstract

Atomically dispersed metal–nitrogen–carbon (M-N-C) catalysts are regarded as ideal catalytic materials for the oxygen reduction reaction (ORR) under alkaline conditions. Compared with other ORR catalysts, M-N-C catalysts exhibit notable advantages, including low cost, high atomic utilization efficiency, and considerable catalytic potential. We provide a systematic review of recent research advances in enhancing the ORR performance of M-N-C catalysts, focusing on catalytic activity and stability. First, the reaction mechanism of the ORR on the surfaces of the M-N-C catalysts is elucidated. Second, the primary strategies employed in recent years to improve their catalytic activity and stability are summarized. Finally, critical research directions that should be prioritized to expedite the commercialization of M-N-C catalysts are outlined.

1. Introduction

To achieve the “dual carbon” goals, the development and application of clean and efficient green new energy sources have become focal areas of research and practice. As a promising energy conversion technology, metal–air batteries are considered a leading solution to address the growing global energy crisis and environmental issues. This is attributed to their high theoretical energy density, straightforward design, and benefits in terms of safety and eco-friendliness [1,2]. Compared with acidic and neutral electrolytes, which are associated with various limitations, alkaline electrolytes represent a more favorable option for metal–air batteries. However, the efficiency of these batteries is hindered by the slow oxygen reduction reaction (ORR) occurring at the cathode, which has greatly delayed their widespread commercial adoption. Thus, the exploration and development of highly effective alkaline oxygen reduction catalysts are essential to promote the practical application and commercial viability of metal–air batteries.
ORR catalysts can be categorized based on their composition into Pt-based catalysts, transition metal compound catalysts, and carbon-based catalysts. Pt-based materials are considered optimal catalysts for the ORR because they exhibit favorable adsorption energy toward oxygen-containing intermediate species [3]. However, the scarcity and high cost of Pt constrain the practical application of Pt-based catalysts. To balance high catalytic activity with reduced Pt consumption, two primary strategies are currently employed to enhance their catalytic performance. The first strategy involves alloying control [4,5]. Wang et al. aimed to reduce Pt consumption through a synergistic design of composition and structure [6]. The as-synthesized Pd@Pt2.7L/C catalyst, obtained by precisely tuning the Pd core structure and the Pt/Pd mass ratio, demonstrates a half-wave potential (E1/2) of 0.9 V, accompanied by a kinetic current density that is 7.8-fold greater than that of commercial Pt/C. However, the increasing cost of Pd constrains its application potential. The second strategy involves structural optimization. Chen et al. synthesized hollow Pt3Ni nanocages through the etching of PtNi3 crystals [7]. The porous structure exposes a greater number of active sites, while the incorporation of Ni optimizes the electronic structure of Pt, leading to an E1/2 of 0.91 V and a Tafel slope as low as 46 mV dec−1. Wang et al. synthesized an ultra-thin Pt shell with tungsten carbide as the core [8], whereas Tian et al. constructed a Pt shell–titanium nickel nitride core structure (with a Pt content of 5 wt%) [9]. Both strategies significantly improved the ORR catalytic activity through enhanced Pt dispersion and d-band center modulation. Although the aforementioned strategies reduce Pt consumption while maintaining high catalytic activity, Pt is prone to oxidation into low-conductivity oxides during the long-term operation of metal–air batteries. Furthermore, its relatively high cost continues to hinder large-scale applications.
Catalysts based on transition metals are considered viable alternatives to Pt-based catalysts, as the eg orbitals of the metal atoms show significant overlap with the 2p orbitals of O2, which facilitates stronger O-O bond coupling. These catalysts are primarily classified into four categories: oxides, carbides, sulfides, and phosphides. Despite the natural abundance of transition metal oxides, including spinel and perovskite structures, their catalytic performance is frequently constrained by poor electrical conductivity and strong surface hydrophilic characteristics. Through the metal atom transposition effect, the charge transfer efficiency of this type of catalyst can be markedly enhanced, leading to improved ORR activity [10]. Transition metal carbides such as tungsten carbide (WC), formed by the embedding of carbon atoms into metal lattices, possess a Pt-like electronic structure and can demonstrate excellent ORR performance. Guo et al. prepared a tungsten carbide (WC) nanoparticle (NP) catalyst with an average particle diameter of 1.9 nanometers [11]. This structural design enhances electrical conductivity while effectively preventing the aggregation of tungsten carbide nanoparticles, leading to excellent ORR performance and a favorable four-electron reaction pathway. The multivalent characteristics of sulfur and phosphorus elements endow transition metal sulfides and transition metal phosphides with dynamic redox capabilities. Dou et al. synthesized a plasma-treated Co9S8/GO catalyst exhibiting a half-wave potential of 0.84 V, which is comparable to that of commercial Pt/C [12]. Although considerable advancements have been achieved in enhancing the catalytic performance of transition metal-based catalysts, their catalytic activity toward the ORR remains insufficient to meet the demands of practical applications. Additionally, their poor electrical conductivity and restricted surface area impede both electron transfer and mass transport during the ORR. In addition, such catalysts exhibit weak binding affinities toward O2 and oxygen-containing intermediates, along with pronounced surface electric field effects, which increase the difficulty of O-O bond cleavage. Consequently, the ORR catalytic activity remains suboptimal.
Carbon-based catalysts have gained significant attention as potential materials for ORR catalysis owing to their excellent electrical conductivity, extensive surface area, and economic viability. Their catalytic activities originate from the asymmetric charge distribution induced by the incorporation of defects or impurity atoms into the sp2-hybridized carbon skeleton, which enhances their capacities to adsorb and chemically interact with O2 [13]. Presently, carbon-based catalysts can be broadly divided into three main groups. The first group includes pristine carbon materials such as graphene, carbon nanotubes, and carbon derivatives obtained from metal–organic frameworks (MOFs). The catalytic activity of pure carbon materials can typically be enhanced through defect engineering. Highly defective carbon quantum dots synthesized via an interfacial self-corrosion strategy achieve an E1/2 of 0.90 V under alkaline conditions [14]. However, their complex synthesis process hinders practical applications. The second group consists of non-metal-doped carbon materials, with nitrogen-doped carbon materials attracting significant interest owing to their advantageous electronegativity balance and superior electron-donating properties. Nitrogen-doped carbon materials can facilitate O2 mass transfer by constructing mesoporous structures, and their catalytic activity can even approach or rival that of Pt-based catalysts [15]. In contrast to the first two categories of carbon materials, transition metal-doped carbon materials possess a greater abundance of active sites for the ORR. As far back as 1964, Jasinski and colleagues identified the catalytic reduction activity of M-N4 macrocyclic complexes, where M stands for a transition metal and N represents a nitrogen atom coordinated to the metal [16]. In 2010, studies explored affordable nitrogen-rich precursors, inorganic metal sources, and carbon supports with high surface areas as economical substitutes for costly metal-macrocyclic compounds in the synthesis of M-N-C catalysts, achieving outstanding electrocatalytic activity [17,18,19]. At present, transition metal-doped carbon catalysts, especially atomically dispersed transition metal–nitrogen–carbon (M-N-C) catalysts, demonstrate numerous highly active sites and are considered the most viable contenders for ORR catalytic applications (Table 1).
Catalytic activity and stability are the two main criteria for evaluating the ORR performance of M-N-C catalysts. Here, we provide a systematic and comprehensive overview of recent research advances aimed at enhancing the catalytic activity and stability of M-N-C catalysts (Figure 1). Based on current research progress, this paper outlines several strategically forward-looking development directions aimed at accelerating the commercialization of M-N-C catalysts.

2. Advances in Research on M-N-C Catalysts

M-N-C catalysts feature metal active sites (M-Nx) that are atomically dispersed and anchored onto a carbon-based framework (Figure 2a). However, due to carrier interactions and reaction conditions, the charge state of metal sites is predominantly cationic rather than neutral. The enhanced electronic structure of M-N-C catalysts arises from the coordination interaction between the transition metal, M, and nitrogen atoms as well as their synergistic interaction with the carbon matrix [50]. As illustrated in Figure 2b, the d-band of the M atom in the M-N-C catalyst is broader compared with that of pure metal, and the d-band center shifts negatively away from the Fermi level. These phenomena, resulting from the hybridization of different p orbitals of nitrogen atoms, facilitate the adsorption and desorption of oxygen-containing intermediates. Secondly, the electronegativities of M and N in M-N-C differ, leading to charge transfer from M to N (Figure 2c). As the d-band becomes increasingly electron-rich, its hybridization with the oxygen p orbitals is enhanced, which facilitates the cleavage of the O-O bond. Ultimately, integrating M-Nx sites within the carbon matrix promotes the generation of additional pyrrolic nitrogen rings, thereby enhancing the catalytic reaction rate [51,52,53,54,55].
To confirm the atomic-level dispersion of M-N-C catalysts, a combination of multiple characterization techniques is typically required, such as aberration-corrected transmission electron microscopy (AC-TEM) and X-ray absorption spectroscopy (XAFs). AC-TEM can directly visualize the dispersion of metal atoms, thereby confirming their isolated existence. XAFs can detect the presence of metal–metal bonds in materials, thereby indicating whether the metals are atomically dispersed. Additionally, Mössbauer spectroscopy, X-ray photoelectron spectroscopy (XPS), and electron paramagnetic resonance spectroscopy (EPR) can serve as complementary techniques to further confirm the atomic-level dispersion of metals on supports.

2.1. Oxygen Reduction Mechanism and Current Challenges of M-N-C Catalysts

The ORR reaction pathway of M-N-C catalysts is predominantly determined by the adsorption and subsequent condensation behavior of O2 on the catalyst surface [38,56,57]. There are three main types: Griffiths, Pauling, and Bridge (Figure 3a). In the Griffiths type, the strong coupling between the d orbitals of the M atom and oxygen molecules leads to a weakening of the O-O bond, thus guiding the ORR to follow a four-electron reaction mechanism. In the Pauling type, only one oxygen atom within the O2 molecule interacts with the M atom, which hinders O-O bond cleavage and results in the ORR proceeding via a two-electron reaction pathway, thereby negatively impacting catalytic activity. The Bridge configuration is applicable to M-N-C catalysts containing two or more adjacent M atoms, where oxygen molecules are simultaneously adsorbed on two metal sites. This structural feature facilitates O-O bond cleavage and allows the ORR to progress through a four-electron reaction route [58].
Additionally, based on whether the O-O bond cleavage occurs before proton–electron transfer, the ORR reaction routes on M-N-C catalysts can be categorized into dissociative and associative pathways (Figure 3b). Both pathways involve key oxygen-containing intermediates (*O and *OH), whereas *OOH is exclusively present in the association pathway. The binding energies of these oxygen-containing intermediate species on the surface of M-N-C catalysts are strongly related to the location of the metal d-band center [59]. Typically, the binding strength of oxygen-containing intermediates on catalyst surfaces should not be too strong or too weak. If the binding strength is too high, reaction intermediates may continuously occupy the catalyst’s active surface, thereby impeding the ongoing progress of the catalytic process. Conversely, if the adsorption strength is too weak, it can impede reactant adsorption and subsequent desorption processes [60]. In addition, an unconventional 2*OH mechanism has been proposed: for traditional catalysts featuring similar, uniformly distributed active sites (M-Nx), 2ΔG(*OH) = ΔG(*O). However, for M-N-C catalysts, ΔG(2*OH) = ΔG(*O) + 1.5 eV; this indicates that the 2*OH reaction pathway should not be overlooked [61]. At the same time, it was observed that the Co-N-C catalyst prefers the *O mechanism, whereas the Fe-N-C catalyst favors the 2*OH reaction pathway [62].
Despite the distinct (Pt-like) electronic characteristics displayed by M-N-C catalysts, significant challenges remain in replacing Pt-based catalysts for large-scale ORR applications. The catalytic performance of M-N-C catalysts is mainly influenced by three key factors: (1) Thermal migration and aggregation of metal atoms lead to the formation of low-activity nanocrystalline materials. (2) The strong binding between oxygen-containing intermediate species and metal active sites leads to a decreased ORR reaction rate, as illustrated in Figure 4. M-N-C is primarily situated on the left region of the volcano plot, suggesting that the binding interaction between M-N-C and *OH is overly strong, thereby impeding desorption and ultimately influencing catalytic performance [63]. (3) The uneven arrangement of pore structures across the catalyst surface restricts the efficient transport of oxygen-containing intermediate species [64,65]. In addition, corrosion of the carbon framework caused by oxidation results in the leaching of metal active sites, thereby compromising catalytic stability [66]. Thus, additional refinement of the microstructural and electronic configurations of M-N-C catalysts is necessary to improve their catalytic performance and durability in ORR.

2.2. Strategies for Enhancing the Catalytic Activity of M-N-C Catalysts

Currently, researchers are primarily exploring two strategies to boost the catalytic efficiency of M-N-C catalysts: (1) enhancing the concentration of M-Nx coordination centers to expand the interaction area between catalytic sites and oxygen-containing intermediate species, thereby accelerating the ORR rate; (2) enhancing the intrinsic catalytic performance by tuning the electronic configuration and coordination surroundings of the metal center, thereby promoting oxygen adsorption and the conversion of reaction intermediates.

2.2.1. Increasing the Density of M-Nx Sites

Owing to the elevated surface energy of transition metal single atoms, they tend to cluster together and undergo oxidation throughout the synthesis and catalytic reaction stages, presenting a significant challenge in the preparation of highly atomically dispersed M-N-C catalysts [66,67]. Therefore, it is essential to restrict the free migration of M atoms to ensure their stabilization through M-N bonding within the carbon support [68]. It is worth noting that the strength of the M-N bond directly influences the loading capacity of the M atoms. Nitrogen-containing organic compounds can donate lone pairs of electrons to N atoms, which exhibit strong coordination affinity toward M atoms, thereby promoting the formation of stable M-N bonds. Porphyrins [69] and phthalocyanines [70] offer a well-defined non-coordinating environment that enables stable binding with M atoms, resulting in the generation of well-dispersed M-N-C active sites. In addition, 2D sheet-like nanographene with a unique structure is considered a promising candidate for hosting M-Nx sites owing to its large specific surface area, superior electrical conductivity, and unique graphite support [65,71]. Graphene inherently lacks active sites that are favorable for doping owing to its chemical inertness. Conversely, graphene with nitrogen doping or structural defects can promote robust π–π stacking interactions, thereby enabling enhanced coordination for anchoring metal atoms, which promotes the generation of isolated single-metal-atom catalytic centers [72,73]. Chen et al. constructed Fe/N-doped electrocatalysts by assembling nitrogen-doped graphene nanosheets with single iron atoms (Figure 5a) [74]. The Fe atoms anchored on the catalyst coordinate with nitrogen atoms in graphene to form stable Fe-N-NG structures. High-temperature pyrolysis and acid leaching generate single-atom dispersed Fe-Nx sites, while the short-range order and large interlayer spacing of the Fe-N-NG structure facilitate O2 diffusion, thereby contributing to enhanced ORR catalytic activity. In addition, Chen et al. successfully developed a high-performance and stable Mn-N-C oxygen reduction catalyst via a three-step synthetic approach integrating chemical doping, mechanochemical doping, and pyrolysis. This synthetic strategy not only significantly enhanced the doping content and density of manganese but also established a structural framework conducive to the synergistic enrichment of manganese and nitrogen species [75].
In addition to stabilizing individual metal atoms via strong interactions between M atoms and carbon supports, nitrogen-rich precursors with cross-linking capabilities can also suppress metal atom migration during high-temperature pyrolysis and facilitate their uniform distribution on the support surface, ultimately contributing to the formation of densely populated M-N-C active sites. MOFs composed of metal nodes and nitrogen-enriched ligands not only provide nitrogen-based coordination environments for metal atoms, but also yield carbon matrices with large surface areas and well-organized porous architectures following high-temperature calcination, which promotes the accessibility of more active centers [21,59,66,76,77]. Fe3+ ions were incorporated into a Zn-based MOF through chemical doping to form a metal complex, which was subsequently utilized for the preparation of an Fe-N-C catalyst featuring Fe-N4 active sites via pyrolysis [27]. The Fe doping content not only has a direct impact on the ORR activity, but also exerts an indirect influence on the catalytic performance by altering key factors including particle dimensions, pore architecture, and nitrogen concentration. Although the encapsulation of metal atoms through enhanced metal–substrate interactions or the construction of nanocage structures represents an effective strategy for stabilizing metal atoms, this approach depends on specific support materials and synthesis conditions, thus limiting its broad applicability to some extent.
Apart from the suppression effect induced by the template, combining multiple stabilization approaches synergistically can more effectively prevent the formation of metal clusters, thus enabling the realization of high metal-atom loading. The stepwise co-anchoring method introduced by Li and colleagues offers a practical solution for the scalable production of M-N-C catalysts featuring high metal-atom loading (Figure 5b,c) [62]. Initially, metal ions form complexes with chelating agents (such as melamine and dicyandiamide) and are immobilized on a high-surface-area porous carbon substrate, offering initial stabilization for the metal atoms. Following this, the surplus chelating agents present on the support surface create a physical barrier around the metal complexes, providing an additional layer of stabilization. The residual products formed during the thermal decomposition of the chelating agent-bound metal complexes help further disperse the metal atoms, acting as a third level of protective stabilization. Meanwhile, N-containing carbon frameworks are able to bond with metal atoms during the thermal decomposition process, resulting in the formation of M-Nx active sites, which effectively inhibit the clustering of metal atoms.
In recent years, M-N-C composites have garnered considerable interest owing to their exceptional atomic efficiency and superior catalytic capabilities. However, traditional synthesis methods face numerous challenges, including high energy consumption, complicated procedures, and significant waste of metal materials, all of which severely hinder their practical application [70,78,79]. Therefore, the scalable and versatile synthesis of highly efficient M-N-C catalysts represents a critical step toward reducing production costs and accelerating industrialization. Mechanochemistry is an effective synthesis technique that utilizes mechanical energy to induce chemical reactions among different solid materials. Ball milling is one of the key techniques for realizing mechanochemical synthesis. It involves repeatedly compressing materials using hard steel or oxide balls to trigger chemical reactions at the interfaces of reactants under relatively mild temperature conditions. Compared with other kilogram-scale methods for preparing single-atom catalysts (SACs), the mechanical ball milling approach for synthesizing M-N-C materials offers several notable advantages [80]: (1) Metal atoms can be evenly distributed across the support, attaining a single-atom dispersion state. (2) High-quality and high-density M-N-C sites with defective scaffolds can be prepared for multifunctional catalytic applications. (3) M-N-C materials can be synthesized without the need for solvents or post-synthesis heat treatment, offering enhanced environmental compatibility. (4) This approach is straightforward and effective, enabling the large-scale production of catalysts at the kilogram level without compromising their structural integrity.

2.2.2. Enhancing Intrinsic Activity

Improving the inherent activity of the active centers is essential for further enhancing the catalytic efficiency of M-N-C materials. A typical M-N-C catalyst features a central metal atom coordinated by adjacent nitrogen atoms within the carbon matrix. Additionally, heteroatoms such as sulfur (S), oxygen (O), and phosphorus (P) can be incorporated into the carbon skeleton to create diverse coordination environments. In M-N-C catalysts, M atoms serve as active centers and interact directly with oxygen intermediates. First-order coordinating atoms bond to M atoms and participate in electron sharing, whereas second-order or higher-order coordinating atoms (collectively referred to as the coordination environment) can adjust the electronic configuration of M-N-C sites through extended electron delocalization. Additional functional groups can also affect the electronic environment of M-N-C active sites via coordination linkages or non-covalent intermolecular forces. Thus, the intrinsic activity of M-N-C sites for the oxygen reduction reaction can be substantially improved by carefully choosing the central metal atom and coordinating elements along with fine-tuning the coordination environment and incorporating suitable auxiliary groups.
Choosing a suitable metal center:
The active sites in M-N-C catalysts are typically metal atoms embedded within the carbon matrix. The electronic structure of these metal atoms plays a crucial role in the ORR, as it determines the adsorption strength of oxygen-containing intermediates on the metal surface, thereby influencing the intrinsic catalytic activity. For transition metals, the filling state of their d orbitals, as well as the interaction between these d orbitals and the p orbitals of oxygen atoms in oxygen-containing intermediates, directly determines the adsorption behavior of the intermediates and the rate of electron transfer during the ORR process. Zheng and colleagues explored the oxygen reduction reaction (ORR) performance of graphene-supported M-N-C catalysts by integrating computational modeling with experimental analysis. Their findings revealed that Fe-N-C catalysts displayed the most superior ORR performance compared with Co-N-C, Mn-N-C, and Ni-N-C catalysts (Figure 6a–c) [81]. Peng and colleagues further studied the impact of Fe(acac)3, FeCl3, and Fe(NO3)3 on Fe-N-C catalyst performance (Figure 6d–f) [82]. Due to their stronger hydrolysis capabilities, Fe(acac)3 and FeCl3 metal salts enable the synthesis of highly dispersed FeNx sites, whereas Fe(NO3)3, with weaker hydrolysis ability, leads to the formation of inactive Fe3C after thermal reduction. The results demonstrated that the catalysts synthesized from three distinct metal salts displayed varying catalytic performances with the following activity sequence: Fe(acac)3 > FeCl3 > Fe(NO3)3 [83]. Thus, Fe is considered the most favorable central metal in M-N-C catalysts for the oxygen reduction reaction, and Fe(acac)3 exhibits the best performance among metal salt precursors. Zhang et al. revealed that distinct metal active centers lead to diverse adsorption configurations of oxygen-containing intermediates, thereby exerting a significant influence on the ORR activity of M-N-C catalysts [84]. In the case of Ni-N-C catalysts with lower binding energy, *O preferentially adsorbs at the M-N bridge site, and this preferential adsorption plays a crucial role in enhancing the ORR activity. In addition, Ruan et al. reported a highly efficient binary FeV-DAS-NC catalyst, wherein the synergistic interaction between Fe and V metal atoms markedly enhances the ORR activity [85].
Regulation of metal coordination atoms:
Crystal field theory suggests that coordination atoms interact with central metals to create coordination field effects, which adjust metal bonding energy, impact oxygen intermediate adsorption at metal sites, and thus affect ORR catalytic performance. Typically, the nitrogen coordination structures present at M-N-C active sites are mainly composed of pyridinic, pyrrolic, graphitic, and oxidized nitrogen forms. These distinct configurations demonstrate varying adsorption behaviors toward oxygen intermediates [86,87,88]. Among these configurations, pyridinic nitrogen not only coordinates with metals to form active sites for ORR, but also activates adjacent carbon atoms, converting them into additional active sites.; pyrrolic nitrogen enhances electron transfer around metal atoms, thereby balancing the adsorption and desorption behaviors of oxygen intermediates [89]; graphitic nitrogen adjusts the electrical conductivity of the catalyst and facilitates the mass transfer of reaction products, thereby favoring the four-electron ORR mechanism and enhancing the catalyst’s intrinsic ORR activity [90]. Zhang et al. successfully designed and synthesized a novel Co/Cu-N-C catalyst via a complex-assisted pyrolysis strategy [91]. The precise modulation of the coordination environment around the Co sites by the Cu-N carrier resulted in a significant enhancement of catalytic activity.
In addition to the influence of the N configuration, the number of nitrogen atoms also plays a role in modulating the charge distribution of metal sites, thereby affecting the adsorption strength between metal active sites and oxygen-containing species [92]. For example, theoretical calculations were employed to predict the optimal adsorption configuration at Fe-Nx sites (x = 1–5) (Figure 6g–i) [93]. Among all Fe-Nx structures, the Fe-N4 configuration exhibits the lowest adsorption barrier for the key oxygen-containing intermediate (i.e., *OH). With the coordination number rising from 1 to 4, the ORR overpotential gradually diminishes. However, when the nitrogen coordination number further increases to 5, the ORR overpotential actually rises. Therefore, the ORR activity and nitrogen coordination number in Fe-Nx catalysts exhibit a “volcanic” relationship.
Figure 6. (a) U1 and U2 values on different square-planar active centers are plotted as a function of the oxygen binding energy; (b) DFT results; (c) LSV curves. (ac) Reproduced with permission [81]. Copyright 2016, Elsevier. (d) Nitrogen adsorption–desorption curve; (e) CV curves and (f) LSV curves in 0.1 M KOH. (df) Reproduced with permission [82]. Copyright 2014, American Chemical Society. (g) Formation energies; ORR free energy curves of FeNx at (h) 0 V and (i) 1.23 V. (gi) Reproduced with permission [93]. Copyright 2013, Royal Society of Chemistry.
Figure 6. (a) U1 and U2 values on different square-planar active centers are plotted as a function of the oxygen binding energy; (b) DFT results; (c) LSV curves. (ac) Reproduced with permission [81]. Copyright 2016, Elsevier. (d) Nitrogen adsorption–desorption curve; (e) CV curves and (f) LSV curves in 0.1 M KOH. (df) Reproduced with permission [82]. Copyright 2014, American Chemical Society. (g) Formation energies; ORR free energy curves of FeNx at (h) 0 V and (i) 1.23 V. (gi) Reproduced with permission [93]. Copyright 2013, Royal Society of Chemistry.
Nanomaterials 15 01257 g006
Regulation of the remote coordination environment:
M-N-C catalysts enable remote atoms and functional groups within the carbon matrix to interact via long-range electron delocalization effects. Therefore, modulating the electronic properties of the carbon matrix can influence the electronic structure of the metal center, thereby enhancing the intrinsic activity of the M-N-C catalyst. The formation of M-N-C catalysts is primarily determined by metal precursors, organic carbon sources, and pyrolysis conditions. Pyrolysis serves as a critical process for constructing the complex architecture of M-N-C materials [94,95,96]. Liu et al. identified three unique FeN4 active sites, each exhibiting different local carbon structural coordination and formed at varying pyrolysis temperatures, denoted as Fe-N4-C10 (D1), Fe-N4-C12 (D2), and Fe-N4-C8 (D3) (Figure 7a) [95]. Generally, at low temperatures, D2 is the predominant active site in Fe-N-C catalysts. As the temperature increases, the relative abundance of D1 rises compared with D2. Upon heating to a certain temperature, D1 dissociates from D2 and is subsequently recaptured at vacancy defects or edges, leading to the formation of new D1 sites. Electrochemical results demonstrate that D1 exhibits higher ORR activity compared with D2. This phenomenon can be attributed to the Fe2+/Fe3+ redox pair formed during high-temperature pyrolysis, which possesses a high redox potential and thereby contributes to the stability of the catalytically active Fe2+ under elevated potential conditions. D3 originates from microporous carbon materials and displays behavior opposite to D1, exhibiting higher intrinsic activity compared with both D1 and D2 [95]. Meanwhile, D3, characterized by abundant edge microporous structures, helps further boost the intrinsic ORR activity of the Fe-N-C catalysts. Based on this analysis, studies of the catalytic mechanisms of D1, D2, and D3 (Figure 7b) reveal that D3 located at micropore edges exhibits the lowest O-O bond cleavage activation energy (0.20 eV), considerably smaller compared with D1 (0.50 eV) and D2 (0.80 eV). Moreover, D3 follows an ideal four-electron reaction pathway.
Analysis of the catalytic reaction behavior of D1, D2, and D3 reveals that the *OOH dissociation process on D1 and D2 involves the movement of the *OH product to the apex of a neighboring carbon atom. Conversely, the edge microporous architecture of D3 promotes the breaking of the O-O bond and the dissociation of *OOH at the metal center. This behavior causes D3 to have a smaller reaction barrier for converting *OOH to *O and *OH (Figure 7c). Based on these findings, Chen et al. further explored the influence of the chemical properties of adjacent nitrogen or carbon atoms surrounding M-Nx sites on the catalytic activity of M-N-C catalysts. The results demonstrate that, within the same D2 configuration, the ORR activity of active sites formed under NH3 pyrolysis conditions is significantly higher than that of active sites generated under Ar pyrolysis conditions. The reason for this difference lies in the fact that pyrolysis under NH3 atmosphere enhances the formation of nitrogen species with high alkalinity on the carbon support (Figure 7d), which is associated with the high degree of electron delocalization on the carbon-based surface and its strong electron-donating capability. This enhances the adsorption between the central metal atom and oxygen-containing ORR intermediate, boosting D2’s intrinsic ORR activity. Based on the aforementioned research findings, alterations in the coordination environment of the carbon skeleton within M-N-C catalysts can modulate their dissociation behavior toward reaction intermediates, thereby improving their intrinsic catalytic activity [94]. Therefore, the synergistic effects of selecting optimal metal active centers and optimizing their coordination environment (including the type and number of first-shell coordination sites and remote regulation by carbon supports) can significantly improve the catalytic performance of M-N-C catalysts.
In M-N-C catalysts, the simultaneous achievement of high active-site density and high intrinsic activity may involve addressing the following key challenges: (1) preventing metal atoms from aggregating under high metal-loading conditions while maintaining their atomic dispersion state; (2) precisely controlling the coordination environment to ensure uniform electronic structures across all active sites; (3) achieving a balance between constructing a porous structure favorable for mass transfer and maintaining adequate electrical conductivity.

2.3. Strategies for Enhancing the Stability of M-N-C Catalysts

In practical applications, the stability of M-N-C catalysts remains an area with substantial room for improvement. Studies show that Fe-N-C catalysts demonstrate superior catalytic performance compared with other transition metal M-N-C catalysts in alkaline ORR [26,97,98,99]. Therefore, enhancing the stability of Fe-N-C catalysts plays a critical role in the development of high-performance M-N-C catalysts suitable for industrial applications. The catalytic performance of Fe-N-C catalysts is challenging to maintain efficient, and the associated specific activity decay can be categorized into two distinct stages: an initial rapid decline in catalytic activity followed through a gradual decline over extended electrocatalytic processes [100]. To preserve the catalytic performance of the catalyst, it is essential to prevent or mitigate the aforementioned degradation process. As illustrated in Figure 8, the primary factors contributing to the decline in stability can be summarized as follows: (1) micropore flooding, (2) Fe leaching, (3) N-group protonation, (4) carbon corrosion, (5) resistance loss, and (6) micropore blockage [50]. To date, strategies to enhance catalyst stability have been advanced through two primary approaches: enhancing the electronic characteristics of Fe-N-C active centers while strengthening the oxidative resistance of the carbon support.

2.3.1. Enhancing Electronic Effect

The deactivation of M-N-C catalysts is primarily attributed to the dissolution or agglomeration of metal atoms. Enhancing the electronic interaction between the carbon substrate and metal sites facilitates the formation of a more stable coordination structure, thereby effectively suppressing metal atom dissolution and agglomeration. Additionally, modulating the electronic interaction between the carbon substrate and metal sites can optimize the electronic structure of metal atoms, thereby preventing active site poisoning caused by excessive adsorption of oxygen-containing intermediates. Some metal nanocrystals on the catalyst surface (such as metal atoms, metal oxides, metal nitrides, and metal carbides) do not directly function as active sites; instead, they influence the charge transfer capability of Fe-Nx sites, thereby enhancing catalytic performance. These nanoparticles serve as supplementary elements that can influence the electronic configuration of Fe-N-C catalysts. Fe/Fe3C nanocrystals alone do not confer high catalytic activity; however, catalysts incorporating both Fe/Fe3C nanocrystals and Fe-N4 sites demonstrate superior activity and stability [107]. After the Fe/Fe3C nanocrystals are removed, the catalytic activity and stability exhibit significant degradation, indicating that Fe/Fe3C nanocrystals contribute to enhancing the ORR performance of Fe-N4 sites (Figure 9a,b). Additionally, the interaction between Fe/Fe3C and Fe-N4 sites enhances oxygen adsorption and mitigates the detachment of Fe atoms during the complex ORR process [108].
The Fe atoms in the diatomic catalyst interact with other transition metal atoms, thereby enhancing the structural stability of the catalyst and mitigating Fe leaching during the reaction process. Additionally, diatomic catalysts can reduce the catalyst’s surface energy and improve its catalytic efficiency by adjusting atomic-level interactions. Zhang et al. synthesized ten bimetallic catalysts using macrocyclic precursors comprising six homonuclear species (Fe2, Co2, Ni2, Cu2, Mn2, and Pd2) and four heteronuclear counterparts (Fe-Cu, Fe-Ni, Cu-Mn, and Cu-Co). The Fe-Cu bimetallic catalyst demonstrated exceptional stability, attributed to its suppression of the Fenton process (with only a 12 mV E1/2 loss after 30,000 cycles, Figure 9c) [109]. Qiu et al. investigated charge transfer between two heteronuclear metal atoms using density functional theory (DFT) [110]. By comparing the number of electrons transferred across different d orbitals, they successfully identified the CoFe-N-C bimetallic atomic site catalyst with optimal performance and elucidated the cooperative electronic interaction between Fe and Co during the ORR catalytic process. Charge transfer readily occurs between adjacent metal atoms, efficiently adjusting the electronic configuration of the metallic active site Fe, optimizing the adsorption energy of reaction intermediates, and thereby significantly enhancing ORR performance. Li et al. proposed a unique semi-metallic electronic structure based on theoretical screening and rational design of Fe/Zn-N-C double-atom catalysts (Figure 9d) [111]. This structure, formed through the synergistic interaction of Fe and Zn, introduces spontaneous spin polarization at the Fermi level, thereby enhancing electron conductivity and facilitating the capture and bonding of free oxygen. Meanwhile, during the ORR process, the FeN4 active center remains intact as a result of the sacrificial ZnN4 site. The 2*H model of the FeZnN6 site in the Fe/Zn-N-C diatomic catalyst confirmed that the N3 and N4 positions selectively coordinate with protons, resulting in the cleavage of the Zn-N bond. After protonation, slight increases in the bond lengths of Fe-N3 and Fe-N4 were observed (from 1.79 Å to 1.91 Å), which may help preserve the FeN4 site from degradation, thereby supporting a continuous and sustainable ORR process. Additionally, the protonation step leaves the electronic configuration of the FeN4 reaction site unchanged (but instead converts ZnN4 to ZnN2), thereby preserving the fully spin-polarized charge carriers present at the Fermi energy and maintaining its catalytic reactivity.
Figure 9. (a) Schematic illustrations of Fe(0)@FeNC; (b) catalyst activity and stability of Fe(0)@FeNC. (a,b) Reproduced with permission [108]. Copyright 2016, American Chemical Society. (c) TEM and EELS images of Fe/Cu-N-C. Reproduced with permission [109]. Copyright 2023, American Chemical Society. (d) ORR process of Fe/Zn-N-C. Reproduced with permission [111]. Copyright 2008, Royal Society of Chemistry. (e) Calculated free energy evolution diagrams for S-Fe-N4. Reproduced with permission [112]. Copyright 2022, Wiley-VCH. (f) ELF images of FeAB-O. Reproduced with permission [113]. Copyright 2020, Springer Nature. (g) ORR process of OH-Fe-N4 catalyst. Reproduced with permission [114]. Copyright 2019, American Chemical Society.
Figure 9. (a) Schematic illustrations of Fe(0)@FeNC; (b) catalyst activity and stability of Fe(0)@FeNC. (a,b) Reproduced with permission [108]. Copyright 2016, American Chemical Society. (c) TEM and EELS images of Fe/Cu-N-C. Reproduced with permission [109]. Copyright 2023, American Chemical Society. (d) ORR process of Fe/Zn-N-C. Reproduced with permission [111]. Copyright 2008, Royal Society of Chemistry. (e) Calculated free energy evolution diagrams for S-Fe-N4. Reproduced with permission [112]. Copyright 2022, Wiley-VCH. (f) ELF images of FeAB-O. Reproduced with permission [113]. Copyright 2020, Springer Nature. (g) ORR process of OH-Fe-N4 catalyst. Reproduced with permission [114]. Copyright 2019, American Chemical Society.
Nanomaterials 15 01257 g009
Introducing axial ligands into the metal center to modulate the electronic structure of the catalytic site effectively suppresses the detachment of metal sites. Typically, the d orbitals associated with the central metal atoms that are perpendicular to the Fe-N-C plane remain partially unoccupied and are capable of interacting with external ligands. In Fe-N-C catalytic systems, axial coordination between metal atoms and ligands including P, O, S, and OH groups can occur. These ligands are capable of altering the electronic configuration of metal centers, which in turn helps improve the stability of the catalyst. For example, in the S-Fe-N4 catalyst, a single Fe atom is coordinated with four pyridinic nitrogen atoms and forms an axial bond with sulfur. Axial coordination with sulfur alters the electronic configuration of the iron center, leading to improved structural stability of the catalytically active regions (Figure 9e) [112]. Moreover, axially coordinated oxygen atoms can break the electronic symmetry of iron, leading to localized electron accumulation and thus facilitating O2 adsorption as well as enhancing charge transfer between the active centers and oxygen-containing molecules (Figure 9f) [113]. The O-Fe-N4 catalyst featuring axial oxygen coordination exhibits remarkably enhanced catalytic performance and remarkable stability, significantly outperforming commercial Pt/C.
Apart from external ligands, ORR intermediates can also naturally coordinate with the central metal atoms, generating internal auxiliary groups. By modulating the planar or axial coordination of oxygen intermediates at the central metal site, the ORR performance can be significantly enhanced. Within the potential range of 0.28 to 1.00 V, Fe atoms spontaneously coordinate with *OH intermediates. The OH-Fe-N4 structure formed by the dynamic coverage of *OH intermediates acts as the true ORR active site, demonstrating superior performance compared with the original Fe-N4 site (Figure 9g) [114]. *OH can modulate the electronic configuration of Fe and modulate the binding and release of reaction intermediates. However, this electrochemical in situ condition strategy is not without limitations, as *OH axial coordination enhances the selectivity toward H2O2, thereby compromising the stability of the catalyst. Han et al. synthesized Cl-Fe-N4-CS catalysts that exhibited enhanced ORR catalytic performance [115]. The Cl-Fe-N4-CS catalyst exhibited an E1/2 value of 0.921 V, with no degradation in E1/2 observed after 10,000 cycles of CV testing, significantly surpassing both the E1/2 of Fe-N4-C and the performance of commercial Pt/C. Density of states analysis indicates that axial coordination of Cl and S doping in the carbon skeleton enhances hybridization between the Fe 3d orbitals at the active center and the O 2p orbitals of the catalyst, which facilitates acceleration of the ORR. Among them, Cl coordination predominantly contributes to high catalytic activity, whereas S doping primarily accounts for the enhanced stability of Fe-N-C catalysts. In summary, axial coordination involving metal atoms represents a viable strategy for enhancing the stability of Fe-N-C catalysts.

2.3.2. Enhancing the Antioxidant Capacity of Carbon-Based Materials

Carbon substrate corrosion represents one of the primary factors contributing to the degradation of carbon-based catalyst performance, with its prevalence predominantly attributed to oxidative processes. Given the detrimental impact of carbon substrate corrosion on the stability of Fe-N-C catalysts, mitigating carbon substrate oxidation is essential. The primary strategies for enhancing the oxidation resistance of carbon substrates include carbon graphitization and edge functionalization. To date, the majority of Fe-N-C catalysts exhibit insufficient graphitization of the carbon matrix, leading to poor corrosion resistance and consequently resulting in active site degradation and a decline in turnover frequency (TOF) [116,117]. Additionally, carbon substrates with low graphitization exhibit poor electron transfer and conductivity [118]. Generally, porous carbon carriers exhibit a high specific surface area, yet their electrochemical oxidation resistance remains limited [119]. Graphitized carbon skeletons exhibit high resistance to electrochemical oxidation under high electrode potentials. They also effectively suppress the oxidation and waterlogging of catalytic active sites, thereby minimizing resistance losses during mass transfer. However, metal atoms tend to be poorly dispersed within these structures [120,121]. Therefore, the development of Fe-N-C catalysts featuring porous graphitic carbon frameworks enriched with Fe-Nx active sites represents a key challenge in achieving efficient and stable ORR. As shown in Figure 10a, Chen et al. developed a Fe-N-C catalyst (PANI-Fe-MCS) composed of thin carbon sheets and porous carbon spheres, exhibiting excellent catalytic performance [122]. A larger electrochemical active area, a reasonable pore size distribution, and a controllable chemical structure contribute to enhanced catalytic performance. PANI-Fe-MCS was employed as a cathode catalyst in an aluminum–air battery and subjected to long-term stability testing, thereby demonstrating its high efficiency and stability. As shown in Figure 10b, Kang et al. reported that carbon nanotubes incorporated dense Fe-Nx centers and were encapsulated by graphitic nitrogen, demonstrating high catalytic activity and stability [123]. The adjacent graphitic nitrogen facilitates improved d-band filling, thereby reducing the partial magnetic moment (spin density), while the incorporation of iron atoms preserves high catalytic activity. Meanwhile, mesoporous/macroporous carbon nanotubes are effective in suppressing water flooding behavior. Compared with the Fe-N-C catalyst derived from ZIF-8 with low graphitization, the carbon nanocomposite Fe-N/CNT-2 demonstrates reduced performance degradation.
Carbon materials typically exhibit edge defects, which are broadly classified into armchair and zigzag configurations. Edge sites generally display different thermodynamic and electrochemical characteristics compared with the basal plane [125,126]. Generally, carbon oxidation predominantly occurs at defect-rich edges. Electrocatalysts with exposed inner surfaces exhibit greater resistance to oxidation compared with those with exposed edge structures [126]. Therefore, controlling the edge structures of the carbon substrate represents an effective strategy for mitigating carbon corrosion. Baek et al. employed O2, H2O, CO2, and H2 to selectively trim the free edges of graphene nanosheets through in situ or ex situ approaches, resulting in catalysts functionalized with edge oxidation groups at distinct positions, designated as EOG-O2, EOG-H2O, EOG-CO2, and EOG-H2 (Figure 10c) [124]. Accelerated chemical corrosion tests and electrochemical stability assessments demonstrated that the EOG-O2 catalyst featuring an ether ring structure ((G)C-OC-C(G)) exhibited superior electrochemical and chemical oxidation stability. Samples containing abundant hydrogen-containing functional groups are prone to further oxidation and exhibit inferior electrochemical stability. Among these groups, the (G) C–COOH functional group is highly susceptible to oxidative destruction, thereby exhibiting the lowest energy barrier. Therefore, current reports indicate that controlling the distribution of defects represents an effective strategy for preventing corrosion of the carbon matrix, thereby enhancing its electrochemical and chemical stability [49,127].

3. Summary and Outlook

Despite being regarded as some of the alkaline ORR catalysts with the most potential, M-N-C materials still fall short in terms of catalytic activity and stability for widespread commercial use. This paper provides a systematic overview of the general strategies employed in recent years to enhance the catalytic activity and stability of M-N-C catalysts. In terms of enhancing catalytic activity, the primary strategies include increasing the number of active sites in the catalyst (i.e., enhancing the M-Nx site density) and improving the intrinsic activity of these active sites, such as by selecting appropriate metal centers, adjusting the coordination atoms around the metal, and optimizing the remote coordination environment. The crucial factor in developing high-efficiency M-N-C catalysts lies in maximizing the number of active sites while maintaining their highest possible intrinsic activity. Additionally, the stability of M-N-C catalysts represents another critical factor influencing their commercialization potential. Currently, strategies to enhance the stability of M-N-C catalysts primarily focus on strengthening the electronic interactions between the carbon substrate and active sites as well as improving the antioxidant properties of the carbon substrate. In theory, further enhancement of the catalytic activity of M-N-C catalysts is achievable alongside improvements in their stability. This could also represent one of the key design principles for future M-N-C catalysts. Additionally, to accelerate the large-scale commercialization of M-N-C catalysts, further research efforts are likely needed in the following areas in the coming phase:
(1)
Performance testing of M-N-C catalysts under actual operating conditions deserves greater emphasis. At present, the testing conditions for most M-N-C catalysts deviate substantially from their real-world operating environments. Increased focus on evaluating catalyst performance under actual operating conditions can significantly expedite their commercialization. To ensure the reliability of M-N-C catalyst performance during the transition from laboratory research to practical applications, the following key parameters should be prioritized in evaluation to bridge the gap between experimental conditions and real-world operating environments: (a) catalytic stability under high current densities; (b) mechanical stability of the catalyst structure; (c) the catalyst’s adaptability to dynamic operating conditions.
(2)
Achieving mass production of M-N-C catalytic materials remains a key challenge. Most of the literature reports on the synthesis of M-N-C catalysts are still limited to the laboratory scale. Scaling up M-N-C catalysts from laboratory-scale synthesis to industrial applications presents the following technical challenges: (a) Dispersibility and agglomeration of metal nanoparticles. Under laboratory conditions, uniform dispersion of single metal atoms can be achieved by precisely controlling the precursor ratio and reaction parameters. However, during the scaling-up process, the uniformity of the precursor mixture decreases, which can easily cause metal atom agglomeration and reduce the density of active sites. (b) Controlled synthesis of porous structures. The performance of M-N-C catalysts relies on a high specific surface area and well-defined hierarchical pore structures (micropores, mesopores, macropores). However, during large-scale synthesis, the uniformity and reproducibility of these pore structures become challenging to maintain, which compromises mass transfer efficiency and limits the exposure of active sites. (c) Temperature gradient challenges in industrial high-temperature heat treatment processes. These thermal gradients can result in localized over-burning or insufficient carbonization, which disrupts the metal–nitrogen coordination structure and consequently diminishes catalytic activity. The template method effectively alleviates the problems faced by M-N-C catalysts in industrialization, such as poor metal dispersion, uncontrollable pore structure, and high cost, through structural pre-design, confinement effects, and in-situ doping strategies.
(3)
Research on the in situ characterization of M-N-C catalysts under operational conditions should be prioritized. Techniques such as in situ Raman spectroscopy and in situ infrared spectroscopy are capable of directly elucidating the reaction pathways of M-N-C catalysts during the ORR, facilitating the rapid identification of key kinetic-determining steps and thereby accelerating the resolution of technical bottlenecks. In situ Raman and infrared spectroscopies enable real-time monitoring of oxygen-containing intermediates adsorbed on the surface of M-N-C catalysts during the ORR. Analyzing variations in the adsorption concentration of oxygen-containing intermediates on the catalyst surface provides insights into identifying the rate-determining steps during the ORR.

Author Contributions

Conceptualization, Y.Z., Q.W. and J.C.; writing—original draft preparation, J.C.; writing—review and editing, Y.Z., M.W., S.L., Q.W., X.D., Z.L. and Y.L.; supervision, Y.Z. and Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support from the National Natural Science Foundation of China (52374314, 52204326, 52204327), the Science and Technology Innovation Program of Hunan Province (2023RC3063), and the Innovation Mover Program of Central South University (2023CXQD028).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Qiyu Wang was employed by the company Zijin Mining Group Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Kang, J.-H.; Lee, J.; Jung, J.-W.; Park, J.; Jang, T.; Kim, H.-S.; Nam, J.-S.; Lim, H.; Yoon, K.R.; Ryu, W.-H.; et al. Lithium–Air Batteries: Air-Breathing Challenges and Perspective. ACS Nano 2020, 14, 14549–14578. [Google Scholar] [CrossRef]
  2. Liang, Z.; Wang, W.; Lu, Y.-C. The path toward practical Li-air batteries. Joule 2022, 6, 2458–2473. [Google Scholar] [CrossRef]
  3. Nan, H.; Dang, D.; Tian, X.L. Structural engineering of robust titanium nitride as effective platinum support for the oxygen reduction reaction. J. Mater. Chem. A 2018, 6, 6065–6073. [Google Scholar] [CrossRef]
  4. Zhao, M.; Wang, X.; Yang, X.; Gilroy, K.D.; Qin, D.; Xia, Y. Hollow Metal Nanocrystals with Ultrathin, Porous Walls and Well-Controlled Surface Structures. Adv. Mater. 2018, 30, 1801956. [Google Scholar] [CrossRef]
  5. Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.; Liu, Z.; Kaya, S.; Nordlund, D.; Ogasawara, H.; et al. Lattice-strain control of the activity in dealloyed core–shell fuel cell catalysts. Nat. Chem. 2010, 2, 454–460. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, M.; Wang, H.; Elnabawy, A.O.; Hood, Z.D.; Chi, M.; Xiao, P.; Zhang, Y.; Mavrikakis, M.; Xia, Y. Facile One-Pot Synthesis of Pd@Pt1L Octahedra with Enhanced Activity and Durability toward Oxygen Reduction. Chem. Mater. 2019, 31, 1370–1380. [Google Scholar] [CrossRef]
  7. Chen, C.; Kang, Y.; Huo, Z.; Zhu, Z.; Huang, W.; Xin, H.L.; Snyder, J.D.; Li, D.; Herron, J.A.; Mavrikakis, M.; et al. Highly Crystalline Multimetallic Nanoframes with Three-Dimensional Electrocatalytic Surfaces. Science 2014, 343, 1339–1343. [Google Scholar] [CrossRef]
  8. Hunt, S.T.; Milina, M.; Alba-Rubio, A.C.; Hendon, C.H.; Dumesic, J.A.; Román-Leshkov, Y. Self-assembly of noble metal monolayers on transition metal carbide nanoparticle catalysts. Science 2016, 352, 974–978. [Google Scholar] [CrossRef]
  9. Park, J.; Zhang, L.; Choi, S.-I.; Roling, L.T.; Lu, N.; Herron, J.A.; Xie, S.; Wang, J.; Kim, M.J.; Mavrikakis, M.; et al. Atomic Layer-by-Layer Deposition of Platinum on Palladium Octahedra for Enhanced Catalysts toward the Oxygen Reduction Reaction. ACS Nano 2015, 9, 2635–2647. [Google Scholar] [CrossRef]
  10. Wu, G.; Wang, J.; Ding, W.; Nie, Y.; Li, L.; Qi, X.; Chen, S.; Wei, Z. A Strategy to Promote the Electrocatalytic Activity of Spinels for Oxygen Reduction by Structure Reversal. Angew. Chem. Int. Ed. 2016, 55, 1340–1344. [Google Scholar] [CrossRef]
  11. Guo, J.; Mao, Z.; Yan, X.; Su, R.; Guan, P.; Xu, B.; Zhang, X.; Qin, G.; Pennycook, S.J. Ultrasmall tungsten carbide catalysts stabilized in graphitic layers for high-performance oxygen reduction reaction. Nano Energy 2016, 28, 261–268. [Google Scholar] [CrossRef]
  12. Dou, S.; Tao, L.; Huo, J.; Wang, S.; Dai, L. Etched and doped Co9S8/graphene hybrid for oxygen electrocatalysis. Energy Environ. Sci. 2016, 9, 1320–1326. [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. Tang, C.; Wang, H.-F.; Chen, X.; Li, B.-Q.; Hou, T.-Z.; Zhang, B.; Zhang, Q.; Titirici, M.-M.; Wei, F. Topological Defects in Metal-Free Nanocarbon for Oxygen Electrocatalysis. Adv. Mater. 2016, 28, 6845–6851. [Google Scholar] [CrossRef]
  15. Guan, B.Y.; Zhang, S.L.; Lou, X.W. Realization of Walnut-Shaped Particles with Macro-/Mesoporous Open Channels through Pore Architecture Manipulation and Their Use in Electrocatalytic Oxygen Reduction. Angew. Chem. Int. Ed. 2018, 57, 6176–6180. [Google Scholar] [CrossRef]
  16. Jasinski, R. A New Fuel Cell Cathode Catalyst. Nature 1964, 201, 1212–1213. [Google Scholar] [CrossRef]
  17. Gu, Y.; Xi, B.J.; Zhang, H.; Ma, Y.C.; Xiong, S.L. Activation of Main-Group Antimony Atomic Sites for Oxygen Reduction Catalysis. Angew. Chem. Int. Ed. 2022, 61, e202202200. [Google Scholar] [CrossRef]
  18. Hu, S.; Li, W.-X. Sabatier principle of metal-support interaction for design of ultrastable metal nanocatalysts. Science 2021, 374, 1360–1365. [Google Scholar] [CrossRef] [PubMed]
  19. Li, L.; Tang, X.; Huang, S.; Lu, C.; Lützenkirchen-Hecht, D.; Yuan, K.; Zhuang, X.; Chen, Y. Longitudinally Grafting of Graphene with Iron Phthalocyanine-based Porous Organic Polymer to Boost Oxygen Electroreduction. Angew. Chem. Int. Ed. 2023, 62, e202301642. [Google Scholar] [CrossRef]
  20. Liu, J.; Liu, Y.; Nan, B.; Wang, D.; Allen, C.; Gong, Z.; He, G.; Fu, K.; Ye, G.; Fei, H. A Two-in-One Strategy to Simultaneously Boost the Site Density and Turnover Frequency of Fe−N−C Oxygen Reduction Catalysts. Angew. Chem. Int. Ed. 2025, 64, e202425196. [Google Scholar] [CrossRef] [PubMed]
  21. Jiao, L.; Wan, G.; Zhang, R.; Zhou, H.; Yu, S.H.; Jiang, H.L. From metal-organic frameworks to single-atom Fe implanted N-doped porous carbons: Efficient oxygen reduction in both alkaline and acidic media. Angew. Chem. Int. Ed. 2018, 130, 8661–8665. [Google Scholar] [CrossRef]
  22. Wu, L.; Zhao, R.; Du, G.; Wang, H.; Hou, M.; Zhang, W.; Sun, P.; Chen, T. Hierarchically porous Fe/N/S/C nanospheres with high-content of Fe-Nx for enhanced ORR and Zn-air battery performance. Green Energy Environ. 2023, 8, 1693–1702. [Google Scholar] [CrossRef]
  23. Peng, L.; Yang, J.; Yang, Y.; Qian, F.; Wang, Q.; Sun-Waterhouse, D.; Shang, L.; Zhang, T.; Waterhouse, G.I.N. Mesopore-Rich Fe–N–C Catalyst with FeN4-O-NCSingle-Atom Sites Delivers Remarkable Oxygen Reduction Reaction Performance in Alkaline Media. Adv. Mater. 2022, 34, 2202544. [Google Scholar] [CrossRef]
  24. Shao, C.; Wu, L.; Wang, Y.; Qu, K.; Chu, H.; Sun, L.; Ye, J.; Li, B.; Wang, X. Engineering asymmetric Fe coordination centers with hydroxyl adsorption for efficient and durable oxygen reduction catalysis. Appl. Catal. B Environ. 2022, 316, 121607. [Google Scholar] [CrossRef]
  25. Ji, B.; Gou, J.; Zheng, Y.; Zhou, X.; Kidkhunthod, P.; Wang, Y.; Tang, Q.; Tang, Y. Metalloid-Cluster Ligands Enabling Stable and Active FeN4-Ten Motifs for the Oxygen Reduction Reaction. Adv. Mater. 2022, 34, 2202714. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, Y.; Ji, S.; Wang, Y.; Dong, J.; Chen, W.; Li, Z.; Shen, R.; Zheng, L.; Zhuang, Z.; Wang, D.; et al. Isolated Single Iron Atoms Anchored on N-Doped Porous Carbon as an Efficient Electrocatalyst for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2017, 56, 6937–6941. [Google Scholar] [CrossRef]
  27. Chen, G.; Liu, P.; Liao, Z.; Sun, F.; He, Y.; Zhong, H.; Zhang, T.; Zschech, E.; Chen, M.; Wu, G.; et al. Zinc-Mediated Template Synthesis of Fe-N-C Electrocatalysts with Densely Accessible Fe-Nx Active Sites for Efficient Oxygen Reduction. Adv. Mater. 2020, 32, 1907399. [Google Scholar] [CrossRef]
  28. Xin, C.; Shang, W.; Hu, J.; Zhu, C.; Guo, J.; Zhang, J.; Dong, H.; Liu, W.; Shi, Y. Integration of Morphology and Electronic Structure Modulation on Atomic Iron-Nitrogen-Carbon Catalysts for Highly Efficient Oxygen Reduction. Adv. Funct. Mater. 2022, 32, 2108345. [Google Scholar] [CrossRef]
  29. Sun, H.; Wang, M.; Du, X.; Jiao, Y.; Liu, S.; Qian, T.; Yan, Y.; Liu, C.; Liao, M.; Zhang, Q.; et al. Modulating the d-band center of boron doped single-atom sites to boost the oxygen reduction reaction. J. Mater. Chem. A 2019, 7, 20952–20957. [Google Scholar] [CrossRef]
  30. Zhou, W.; Su, H.; Li, Y.; Liu, M.; Zhang, H.; Zhang, X.; Sun, X.; Xu, Y.; Liu, Q.; Wei, S. Identification of the Evolving Dynamics of Coordination-Unsaturated Iron Atomic Active Sites under Reaction Conditions. ACS Energy Lett. 2021, 6, 3359–3366. [Google Scholar] [CrossRef]
  31. Hu, L.; Dai, C.; Chen, L.; Zhu, Y.; Hao, Y.; Zhang, Q.; Gu, L.; Feng, X.; Yuan, S.; Wang, L.; et al. Metal-Triazolate-Framework-Derived FeN4Cl11 Single-Atom Catalysts with Hierarchical Porosity for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2021, 60, 27324–27329. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, K.; Lu, Z.; Lei, J.; Liu, Z.; Li, Y.; Cao, Y. Modulation of Ligand Fields in a Single-Atom Site by the Molten Salt Strategy for Enhanced Oxygen Bifunctional Activity for Zinc-Air Batteries. ACS Nano 2022, 16, 11944–11956. [Google Scholar] [CrossRef] [PubMed]
  33. Zhao, X.; Wang, J.; Wang, J.; Yang, M.; Yan, C.; Zou, G.; Tse, J.S.; Fernandez, C.; Peng, Q. Atomically dispersed quintuple nitrogen and oxygen co-coordinated zirconium on graphene-type substrate for highly efficient oxygen reduction reaction. Cell Rep. Phys. Sci. 2022, 3, 100773. [Google Scholar] [CrossRef]
  34. 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]
  35. Yang, L.; Shi, L.; Wang, D.; Lv, Y.; Cao, D. Single-atom cobalt electrocatalysts for foldable solid-state Zn-air battery. Nano Energy 2018, 50, 691–698. [Google Scholar] [CrossRef]
  36. Kong, Z.; Liu, T.; Hou, K.; Guan, L. Atomically dispersed Mn–N4 electrocatalyst with high oxygen reduction reaction catalytic activity from metal–organic framework ZIF-8 by minimal-water-assisted mechanochemical synthesis. J. Mater. Chem. A 2022, 10, 2826–2834. [Google Scholar] [CrossRef]
  37. Wang, X.; An, Y.; Liu, L.; Fang, L.; Liu, Y.; Zhang, J.; Qi, H.; Heine, T.; Li, T.; Kuc, A. Atomically dispersed pentacoordinated-zirconium catalyst with axial oxygen ligand for oxygen reduction reaction. Angew. Chem. Int. Ed. 2022, 134, e202209746. [Google Scholar] [CrossRef]
  38. Chen, Y.; Gao, R.; Ji, S.; Li, H.; Tang, K.; Jiang, P.; Hu, H.; Zhang, Z.; Hao, H.; Qu, Q.; et al. Atomic-Level Modulation of Electronic Density at Cobalt Single-Atom Sites Derived from Metal–Organic Frameworks: Enhanced Oxygen Reduction Performance. Angew. Chem. Int. Ed. 2021, 60, 3212–3221. [Google Scholar] [CrossRef]
  39. Shang, H.; Jiang, Z.; Zhou, D.; Pei, J.; Wang, Y.; Dong, J.; Zheng, X.; Zhang, J.; Chen, W. Engineering a metal–organic framework derived Mn–N4–CxSy atomic interface for highly efficient oxygen reduction reaction. Chem. Sci. 2020, 11, 5994–5999. [Google Scholar] [CrossRef]
  40. Jiang, Z.; Sun, W.; Shang, H.; Chen, W.; Sun, T.; Li, H.; Dong, J.; Zhou, J.; Li, Z.; Wang, Y.; et al. Atomic interface effect of a single atom copper catalyst for enhanced oxygen reduction reactions. Energy Environ. Sci. 2019, 12, 3508–3514. [Google Scholar] [CrossRef]
  41. Yang, Z.; Wang, X.; Zhu, M.; Leng, X.; Chen, W.; Wang, W.; Xu, Q.; Yang, L.-M.; Wu, Y. Structural revolution of atomically dispersed Mn sites dictates oxygen reduction performance. Nano Res. 2021, 14, 4512–4519. [Google Scholar] [CrossRef]
  42. Han, X.; Zhang, T.; Chen, W.; Dong, B.; Meng, G.; Zheng, L.; Yang, C.; Sun, X.; Zhuang, Z.; Wang, D.; et al. Mn@N4 Oxygen Reduction Electrocatalyst: Operando Investigation of Active Sites and High Performance in Zinc–Air Battery. Adv. Energy Mater. 2021, 11, 2002753. [Google Scholar] [CrossRef]
  43. Zhang, T.; Wang, F.; Yang, C.; Han, X.; Liang, C.; Zhang, Z.; Li, Y.; Han, A.; Liu, J.; Liu, B. Boosting ORR performance by single atomic divacancy Zn-N3C-C8 sites on ultrathin N-doped carbon nanosheets. Chem Catal. 2022, 2, 836–852. [Google Scholar] [CrossRef]
  44. Yin, L.; Zhang, S.; Sun, M.; Wang, S.; Huang, B.; Du, Y. Heteroatom-Driven Coordination Fields Altering Single Cerium Atom Sites for Efficient Oxygen Reduction Reaction. Adv. Mater. 2023, 35, 2302485. [Google Scholar] [CrossRef] [PubMed]
  45. Sun, Z.; Zhang, H.; Cao, L.; Liu, X.; Wu, D.; Shen, X.; Zhang, X.; Chen, Z.; Ru, S.; Zhu, X.; et al. Understanding Synergistic Catalysis on Cu-Se Dual Atom Sites via Operando X-ray Absorption Spectroscopy in Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2023, 62, e202217719. [Google Scholar] [CrossRef] [PubMed]
  46. Li, Z.; Tian, Z.; Cheng, H.; Wang, T.; Zhang, W.; Lu, Y.; Lai, Y.; He, G. Engineering d-band center of FeN4 moieties for efficient oxygen reduction reaction electrocatalysts. Energy Storage Mater. 2023, 59, 102764. [Google Scholar] [CrossRef]
  47. Li, Z.; Wang, Q.; Cheng, H.; Wang, M.; Hu, S.; He, G.; Tian, Z. Rectification effect: A universal strategy for single-atom electrocatalysts to enhance oxygen reduction reaction. Energy Storage Mater. 2025, 76, 104121. [Google Scholar] [CrossRef]
  48. Li, Z.; Cheng, H.; Lu, Y.; Wang, T.; Li, Y.; Zhang, W.; He, G.; Tian, Z. Potent Charge-Trapping for Boosted Electrocatalytic Oxygen Reduction. Adv. Energy Mater. 2023, 13, 2203963. [Google Scholar] [CrossRef]
  49. Asset, T.; Atanassov, P. Iron-Nitrogen-Carbon Catalysts for Proton Exchange Membrane Fuel Cells. Joule 2020, 4, 33–44. [Google Scholar] [CrossRef]
  50. Ma, Q.; Jin, H.; Zhu, J.; Li, Z.; Xu, H.; Liu, B.; Zhang, Z.; Ma, J.; Mu, S. Stabilizing Fe–N–C Catalysts as Model for Oxygen Reduction Reaction. Adv. Sci. 2021, 8, 2102209. [Google Scholar] [CrossRef]
  51. Jiao, L.; Li, J.; Richard, L.L.; Sun, Q.; Stracensky, T.; Liu, E.; Sougrati, M.T.; Zhao, Z.; Yang, F.; Zhong, S.; et al. Chemical vapour deposition of Fe–N–C oxygen reduction catalysts with full utilization of dense Fe–N4 sites. Nat. Mater. 2021, 20, 1385–1391. [Google Scholar] [CrossRef]
  52. Qiao, Y.; Yuan, P.; Hu, Y.; Zhang, J.; Mu, S.; Zhou, J.; Li, H.; Xia, H.; He, J.; Xu, Q. Sulfuration of an Fe–N–C Catalyst Containing FexC/Fe Species to Enhance the Catalysis of Oxygen Reduction in Acidic Media and for Use in Flexible Zn–Air Batteries. Adv. Mater. 2018, 30, 1804504. [Google Scholar] [CrossRef]
  53. Tian, H.; Song, A.; Zhang, P.; Sun, K.; Wang, J.; Sun, B.; Fan, Q.; Shao, G.; Chen, C.; Liu, H.; et al. High Durability of Fe–N–C Single-Atom Catalysts with Carbon Vacancies toward the Oxygen Reduction Reaction in Alkaline Media. Adv. Mater. 2023, 35, 2210714. [Google Scholar] [CrossRef]
  54. Chen, Z.; Niu, H.; Ding, J.; Liu, H.; Chen, P.H.; Lu, Y.H.; Lu, Y.R.; Zuo, W.; Han, L.; Guo, Y. Unraveling the origin of sulfur-doped Fe-N-C single-atom catalyst for enhanced oxygen reduction activity: Effect of iron spin-state tuning. Angew. Chem. Int. Ed. 2021, 133, 25608–25614. [Google Scholar] [CrossRef]
  55. Dai, Y.; Liu, B.; Zhang, Z.; Guo, P.; Liu, C.; Zhang, Y.; Zhao, L.; Wang, Z. Tailoring the d-Orbital Splitting Manner of Single Atomic Sites for Enhanced Oxygen Reduction. Adv. Mater. 2023, 35, 2210757. [Google Scholar] [CrossRef]
  56. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  57. He, Y.; Liu, S.; Priest, C.; Shi, Q.; Wu, G. Atomically dispersed metal–nitrogen–carbon catalysts for fuel cells: Advances in catalyst design, electrode performance, and durability improvement. Chem. Soc. Rev. 2020, 49, 3484–3524. [Google Scholar] [CrossRef] [PubMed]
  58. Zhong, L.; Li, S. Unconventional Oxygen Reduction Reaction Mechanism and Scaling Relation on Single-Atom Catalysts. ACS Catal. 2020, 10, 4313–4318. [Google Scholar] [CrossRef]
  59. Lu, X.F.; Xia, B.Y.; Zang, S.Q.; Lou, X.W. Metal–organic frameworks based electrocatalysts for the oxygen reduction reaction. Angew. Chem. Int. Ed. 2020, 132, 4662–4678. [Google Scholar] [CrossRef]
  60. Kulkarni, A.; Siahrostami, S.; Patel, A.; Nørskov, J.K. Understanding Catalytic Activity Trends in the Oxygen Reduction Reaction. Chem. Rev. 2018, 118, 2302–2312. [Google Scholar] [CrossRef] [PubMed]
  61. Xiong, Y.; Li, H.; Liu, C.; Zheng, L.; Liu, C.; Wang, J.-O.; Liu, S.; Han, Y.; Gu, L.; Qian, J.; et al. Single-Atom Fe Catalysts for Fenton-Like Reactions: Roles of Different N Species. Adv. Mater. 2022, 34, 2110653. [Google Scholar] [CrossRef]
  62. Li, P.; Jiao, Y.; Ruan, Y.; Fei, H.; Men, Y.; Guo, C.; Wu, Y.; Chen, S. Revealing the role of double-layer microenvironments in pH-dependent oxygen reduction activity over metal-nitrogen-carbon catalysts. Nat. Commun. 2023, 14, 6936. [Google Scholar] [CrossRef]
  63. Yang, H.; Chen, Z.; Kou, S.; Lu, G.; Chen, D.; Liu, Z. Carbon-supported catalysts with atomically dispersed metal sites for oxygen electroreduction: Present and future perspectives. J. Mater. Chem. A 2021, 9, 15919–15936. [Google Scholar] [CrossRef]
  64. Liu, L.; Corma, A. Confining isolated atoms and clusters in crystalline porous materials for catalysis. Nat. Rev. Mater. 2021, 6, 244–263. [Google Scholar] [CrossRef]
  65. Sun, J.-F.; Xu, Q.-Q.; Qi, J.-L.; Zhou, D.; Zhu, H.-Y.; Yin, J.-Z. Isolated Single Atoms Anchored on N-Doped Carbon Materials as a Highly Efficient Catalyst for Electrochemical and Organic Reactions. ACS Sustain. Chem. Eng. 2020, 8, 14630–14656. [Google Scholar] [CrossRef]
  66. Zhang, H.; Hwang, S.; Wang, M.; Feng, Z.; Karakalos, S.; Luo, L.; Qiao, Z.; Xie, X.; Wang, C.; Su, D.; et al. Single Atomic Iron Catalysts for Oxygen Reduction in Acidic Media: Particle Size Control and Thermal Activation. J. Am. Chem. Soc. 2017, 139, 14143–14149. [Google Scholar] [CrossRef]
  67. Kou, Z.; Zang, W.; Ma, Y.; Pan, Z.; Mu, S.; Gao, X.; Tang, B.; Xiong, M.; Zhao, X.; Cheetham, A.K.; et al. Cage-confinement pyrolysis route to size-controlled molybdenum-based oxygen electrode catalysts: From isolated atoms to clusters and nanoparticles. Nano Energy 2020, 67, 104288. [Google Scholar] [CrossRef]
  68. Jeong, H.; Shin, S.; Lee, H. Heterogeneous Atomic Catalysts Overcoming the Limitations of Single-Atom Catalysts. ACS Nano 2020, 14, 14355–14374. [Google Scholar] [CrossRef]
  69. Yi, J.-D.; Xu, R.; Chai, G.-L.; Zhang, T.; Zang, K.; Nan, B.; Lin, H.; Liang, Y.-L.; Lv, J.; Luo, J.; et al. Cobalt single-atoms anchored on porphyrinic triazine-based frameworks as bifunctional electrocatalysts for oxygen reduction and hydrogen evolution reactions. J. Mater. Chem. A 2019, 7, 1252–1259. [Google Scholar] [CrossRef]
  70. Wan, W.; Triana, C.A.; Lan, J.; Li, J.; Allen, C.S.; Zhao, Y.; Iannuzzi, M.; Patzke, G.R. Bifunctional Single Atom Electrocatalysts: Coordination–Performance Correlations and Reaction Pathways. ACS Nano 2020, 14, 13279–13293. [Google Scholar] [CrossRef] [PubMed]
  71. Zhuo, H.-Y.; Zhang, X.; Liang, J.-X.; Yu, Q.; Xiao, H.; Li, J. Theoretical Understandings of Graphene-based Metal Single-Atom Catalysts: Stability and Catalytic Performance. Chem. Rev. 2020, 120, 12315–12341. [Google Scholar] [CrossRef] [PubMed]
  72. Shaari, N.; Kamarudin, S.K. Graphene in electrocatalyst and proton conductiong membrane in fuel cell applications: An overview. Renew. Sustain. Energy Rev. 2017, 69, 862–870. [Google Scholar] [CrossRef]
  73. Shu, X.; Chen, S.; Chen, S.; Pan, W.; Zhang, J. Cobalt nitride embedded holey N-doped graphene as advanced bifunctional electrocatalysts for Zn-Air batteries and overall water splitting. Carbon 2020, 157, 234–243. [Google Scholar] [CrossRef]
  74. Chen, S.; Zhang, N.; Narváez Villarrubia, C.W.; Huang, X.; Xie, L.; Wang, X.; Kong, X.; Xu, H.; Wu, G.; Zeng, J.; et al. Single Fe atoms anchored by short-range ordered nanographene boost oxygen reduction reaction in acidic media. Nano Energy 2019, 66, 104164. [Google Scholar] [CrossRef]
  75. Chen, G.; Qiu, X.; Liu, S.; Cui, Y.; Sun, Y.; Zhang, Y.; Liu, Y.; Liu, G.; Kim, Y.; Xing, W.; et al. Mn–N–C with High-Density Atomically Dispersed Mn Active Sites for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2025, 64, e202503934. [Google Scholar] [CrossRef]
  76. Zhang, H.; Chung, H.T.; Cullen, D.A.; Wagner, S.; Kramm, U.I.; More, K.L.; Zelenay, P.; Wu, G. High-performance fuel cell cathodes exclusively containing atomically dispersed iron active sites. Energy Environ. Sci. 2019, 12, 2548–2558. [Google Scholar] [CrossRef]
  77. Lai, Q.; Zheng, L.; Liang, Y.; He, J.; Zhao, J.; Chen, J. Metal–Organic-Framework-Derived Fe-N/C Electrocatalyst with Five-Coordinated Fe-Nx Sites for Advanced Oxygen Reduction in Acid Media. ACS Catal. 2017, 7, 1655–1663. [Google Scholar] [CrossRef]
  78. Cui, X.; Li, W.; Ryabchuk, P.; Junge, K.; Beller, M. Bridging homogeneous and heterogeneous catalysis by heterogeneous single-metal-site catalysts. Nat. Catal. 2018, 1, 385–397. [Google Scholar] [CrossRef]
  79. Qu, Y.; Li, Z.; Chen, W.; Lin, Y.; Yuan, T.; Yang, Z.; Zhao, C.; Wang, J.; Zhao, C.; Wang, X.; et al. Direct transformation of bulk copper into copper single sites via emitting and trapping of atoms. Nat. Catal. 2018, 1, 781–786. [Google Scholar] [CrossRef]
  80. Hu, Y.; Li, B.; Yu, C.; Fang, H.; Li, Z. Mechanochemical preparation of single atom catalysts for versatile catalytic applications: A perspective review. Mater. Today 2023, 63, 288–312. [Google Scholar] [CrossRef]
  81. Zheng, Y.; Yang, D.-S.; Kweun, J.M.; Li, C.; Tan, K.; Kong, F.; Liang, C.; Chabal, Y.J.; Kim, Y.Y.; Cho, M.; et al. Rational design of common transition metal-nitrogen-carbon catalysts for oxygen reduction reaction in fuel cells. Nano Energy 2016, 30, 443–449. [Google Scholar] [CrossRef]
  82. Peng, H.; Liu, F.; Liu, X.; Liao, S.; You, C.; Tian, X.; Nan, H.; Luo, F.; Song, H.; Fu, Z.; et al. Effect of Transition Metals on the Structure and Performance of the Doped Carbon Catalysts Derived From Polyaniline and Melamine for ORR Application. ACS Catal. 2014, 4, 3797–3805. [Google Scholar] [CrossRef]
  83. Gao, L.; Xiao, M.; Jin, Z.; Liu, C.; Zhu, J.; Ge, J.; Xing, W. Correlating Fe source with Fe-N-C active site construction: Guidance for rational design of high-performance ORR catalyst. J. Energy Chem. 2018, 27, 1668–1673. [Google Scholar] [CrossRef]
  84. Zhang, D.; She, F.; Chen, J.; Wei, L.; Li, H. Why Do Weak-Binding M–N–C Single-Atom Catalysts Possess Anomalously High Oxygen Reduction Activity? J. Am. Chem. Soc. 2025, 147, 6076–6086. [Google Scholar] [CrossRef]
  85. Ruan, Q.-D.; Feng, R.; Liu, L.; Zhang, L.; Feng, J.-J.; Shi, Y.; Gao, Y.-J.; Wang, A.-J. Atomically dispersed dual-active-site catalyst for enhanced oxygen reduction reaction: Synergistic effect of Fe and V in M−N−C structures. Chem. Eng. J. 2025, 509, 161441. [Google Scholar] [CrossRef]
  86. Zhang, N.; Zhou, T.; Chen, M.; Feng, H.; Yuan, R.; Zhong, C.a.; Yan, W.; Tian, Y.; Wu, X.; Chu, W.; et al. High-purity pyrrole-type FeN4 sites as a superior oxygen reduction electrocatalyst. Energy Environ. Sci. 2020, 13, 111–118. [Google Scholar] [CrossRef]
  87. Ha, Y.; Fei, B.; Yan, X.; Xu, H.; Chen, Z.; Shi, L.; Fu, M.; Xu, W.; Wu, R. Atomically Dispersed Co-Pyridinic N-C for Superior Oxygen Reduction Reaction. Adv. Energy Mater. 2020, 10, 2002592. [Google Scholar] [CrossRef]
  88. He, Y.; Guo, H.; Hwang, S.; Yang, X.; He, Z.; Braaten, J.; Karakalos, S.; Shan, W.; Wang, M.; Zhou, H.; et al. Single Cobalt Sites Dispersed in Hierarchically Porous Nanofiber Networks for Durable and High-Power PGM-Free Cathodes in Fuel Cells. Adv. Mater. 2020, 32, 2003577. [Google Scholar] [CrossRef] [PubMed]
  89. Mun, Y.; Lee, S.; Kim, K.; Kim, S.; Lee, S.; Han, J.W.; Lee, J. Versatile Strategy for Tuning ORR Activity of a Single Fe-N4 Site by Controlling Electron-Withdrawing/Donating Properties of a Carbon Plane. J. Am. Chem. Soc. 2019, 141, 6254–6262. [Google Scholar] [CrossRef] [PubMed]
  90. Chao, G.; Zhang, L.; Wang, D.; Chen, S.; Guo, H.; Xu, K.; Fan, W.; Liu, T. Activation of graphitic nitrogen sites for boosting oxygen reduction. Carbon 2020, 159, 611–616. [Google Scholar] [CrossRef]
  91. 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]
  92. Berwanger, J.; Polesya, S.; Mankovsky, S.; Ebert, H.; Giessibl, F.J. Atomically Resolved Chemical Reactivity of Small Fe Clusters. Phys. Rev. Lett. 2020, 124, 096001. [Google Scholar] [CrossRef]
  93. Li, Y.; Liu, X.; Zheng, L.; Shang, J.; Wan, X.; Hu, R.; Guo, X.; Hong, S.; Shui, J. Preparation of Fe–N–C catalysts with FeNx (x = 1, 3, 4) active sites and comparison of their activities for the oxygen reduction reaction and performances in proton exchange membrane fuel cells. J. Mater. Chem. A 2019, 7, 26147–26153. [Google Scholar] [CrossRef]
  94. Zitolo, A.; Goellner, V.; Armel, V.; Sougrati, M.-T.; Mineva, T.; Stievano, L.; Fonda, E.; Jaouen, F. Identification of catalytic sites for oxygen reduction in iron- and nitrogen-doped graphene materials. Nat. Mater. 2015, 14, 937–942. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, K.; Wu, G.; Wang, G. Role of Local Carbon Structure Surrounding FeN4 Sites in Boosting the Catalytic Activity for Oxygen Reduction. J. Phys. Chem. C 2017, 121, 11319–11324. [Google Scholar] [CrossRef]
  96. Jia, Q.; Ramaswamy, N.; Hafiz, H.; Tylus, U.; Strickland, K.; Wu, G.; Barbiellini, B.; Bansil, A.; Holby, E.F.; Zelenay, P.; et al. Experimental Observation of Redox-Induced Fe–N Switching Behavior as a Determinant Role for Oxygen Reduction Activity. ACS Nano 2015, 9, 12496–12505. [Google Scholar] [CrossRef]
  97. Jiao, L.; Zhang, R.; Wan, G.; Yang, W.; Wan, X.; Zhou, H.; Shui, J.; Yu, S.-H.; Jiang, H.-L. Nanocasting SiO2 into metal-organic frameworks imparts dual protection to high-loading Fe single-atom electrocatalysts. Nat. Commun. 2020, 11, 2831. [Google Scholar] [CrossRef] [PubMed]
  98. Jin, H.; Kou, Z.; Cai, W.; Zhou, H.; Ji, P.; Liu, B.; Radwan, A.; He, D.; Mu, S. P–Fe bond oxygen reduction catalysts toward high-efficiency metal–air batteries and fuel cells. J. Mater. Chem. A 2020, 8, 9121–9127. [Google Scholar] [CrossRef]
  99. Chen, Y.; Ji, S.; Zhao, S.; Chen, W.; Dong, J.; Cheong, W.-C.; Shen, R.; Wen, X.; Zheng, L.; Rykov, A.I.; et al. Enhanced oxygen reduction with single-atomic-site iron catalysts for a zinc-air battery and hydrogen-air fuel cell. Nat. Commun. 2018, 9, 5422. [Google Scholar] [CrossRef]
  100. Banham, D.; Ye, S.; Pei, K.; Ozaki, J.-i.; Kishimoto, T.; Imashiro, Y. A review of the stability and durability of non-precious metal catalysts for the oxygen reduction reaction in proton exchange membrane fuel cells. J. Power Sources 2015, 285, 334–348. [Google Scholar] [CrossRef]
  101. Choi, C.H.; Baldizzone, C.; Grote, J.-P.; Schuppert, A.K.; Jaouen, F.; Mayrhofer, K.J.J. Stability of Fe-N-C Catalysts in Acidic Medium Studied by Operando Spectroscopy. Angew. Chem. Int. Ed. 2015, 54, 12753–12757. [Google Scholar] [CrossRef]
  102. Herranz, J.; Jaouen, F.; Lefèvre, M.; Kramm, U.I.; Proietti, E.; Dodelet, J.-P.; Bogdanoff, P.; Fiechter, S.; Abs-Wurmbach, I.; Bertrand, P.; et al. Unveiling N-Protonation and Anion-Binding Effects on Fe/N/C Catalysts for O2 Reduction in Proton-Exchange-Membrane Fuel Cells. J. Phys. Chem. C 2011, 115, 16087–16097. [Google Scholar] [CrossRef]
  103. Choi, J.-Y.; Yang, L.; Kishimoto, T.; Fu, X.; Ye, S.; Chen, Z.; Banham, D. Is the rapid initial performance loss of Fe/N/C non precious metal catalysts due to micropore flooding? Energy Environ. Sci. 2017, 10, 296–305. [Google Scholar] [CrossRef]
  104. Wang, Y.-C.; Lai, Y.-J.; Wan, L.-Y.; Yang, H.; Dong, J.; Huang, L.; Chen, C.; Rauf, M.; Zhou, Z.-Y.; Sun, S.-G. Suppression Effect of Small Organic Molecules on Oxygen Reduction Activity of Fe/N/C Catalysts. ACS Energy Lett. 2018, 3, 1396–1401. [Google Scholar] [CrossRef]
  105. Banham, D.; Kishimoto, T.; Sato, T.; Kobayashi, Y.; Narizuka, K.; Ozaki, J.-i.; Zhou, Y.; Marquez, E.; Bai, K.; Ye, S. New insights into non-precious metal catalyst layer designs for proton exchange membrane fuel cells: Improving performance and stability. J. Power Sources 2017, 344, 39–45. [Google Scholar] [CrossRef]
  106. Lu, Y.; Zhu, Y.; Chen, Z.; Chen, C.; Li, X.; Yu, H.; Peng, K.; Tian, Z. Boosting the Performance of Aluminum–Air Batteries by Interface Modification. ACS Appl. Mater. Interfaces 2024, 16, 37818–37828. [Google Scholar] [CrossRef]
  107. Jiang, W.-J.; Gu, L.; Li, L.; Zhang, Y.; Zhang, X.; Zhang, L.-J.; Wang, J.-Q.; Hu, J.-S.; Wei, Z.; Wan, L.-J. Understanding the High Activity of Fe–N–C Electrocatalysts in Oxygen Reduction: Fe/Fe3C Nanoparticles Boost the Activity of Fe–Nx. J. Am. Chem. Soc. 2016, 138, 3570–3578. [Google Scholar] [CrossRef] [PubMed]
  108. Li, Z.; Wei, L.; Jiang, W.-J.; Hu, Z.; Luo, H.; Zhao, W.; Xu, T.; Wu, W.; Wu, M.; Hu, J.-S. Chemical state of surrounding iron species affects the activity of Fe-Nx for electrocatalytic oxygen reduction. Appl. Catal. B Environ. 2019, 251, 240–246. [Google Scholar] [CrossRef]
  109. Zhang, Y.-X.; Zhang, S.; Huang, H.; Liu, X.; Li, B.; Lee, Y.; Wang, X.; Bai, Y.; Sun, M.; Wu, Y.; et al. General Synthesis of a Diatomic Catalyst Library via a Macrocyclic Precursor-Mediated Approach. J. Am. Chem. Soc. 2023, 145, 4819–4827. [Google Scholar] [CrossRef]
  110. Zhou, X.; Gao, J.; Hu, Y.; Jin, Z.; Hu, K.; Reddy, K.M.; Yuan, Q.; Lin, X.; Qiu, H.-J. Theoretically Revealed and Experimentally Demonstrated Synergistic Electronic Interaction of CoFe Dual-Metal Sites on N-doped Carbon for Boosting Both Oxygen Reduction and Evolution Reactions. Nano Lett. 2022, 22, 3392–3399. [Google Scholar] [CrossRef]
  111. Li, H.; Di, S.; Niu, P.; Wang, S.; Wang, J.; Li, L. A durable half-metallic diatomic catalyst for efficient oxygen reduction. Energy Environ. Sci. 2022, 15, 1601–1610. [Google Scholar] [CrossRef]
  112. Li, L.; Huang, S.; Cao, R.; Yuan, K.; Lu, C.; Huang, B.; Tang, X.; Hu, T.; Zhuang, X.; Chen, Y. Optimizing Microenvironment of Asymmetric N,S-Coordinated Single-Atom Fe via Axial Fifth Coordination toward Efficient Oxygen Electroreduction. Small 2022, 18, 2105387. [Google Scholar] [CrossRef]
  113. Chen, K.; Liu, K.; An, P.; Li, H.; Lin, Y.; Hu, J.; Jia, C.; Fu, J.; Li, H.; Liu, H.; et al. Iron phthalocyanine with coordination induced electronic localization to boost oxygen reduction reaction. Nat. Commun. 2020, 11, 4173. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, Y.; Tang, Y.-J.; Zhou, K. Self-Adjusting Activity Induced by Intrinsic Reaction Intermediate in Fe–N–C Single-Atom Catalysts. J. Am. Chem. Soc. 2019, 141, 14115–14119. [Google Scholar] [CrossRef] [PubMed]
  115. Han, Y.; Wang, Y.; Xu, R.; Chen, W.; Zheng, L.; Han, A.; Zhu, Y.; Zhang, J.; Zhang, H.; Luo, J.; et al. Electronic structure engineering to boost oxygen reduction activity by controlling the coordination of the central metal. Energy Environ. Sci. 2018, 11, 2348–2352. [Google Scholar] [CrossRef]
  116. Choi, C.H.; Lim, H.-K.; Chung, M.W.; Chon, G.; Ranjbar Sahraie, N.; Altin, A.; Sougrati, M.-T.; Stievano, L.; Oh, H.S.; Park, E.S.; et al. The Achilles’ heel of iron-based catalysts during oxygen reduction in an acidic medium. Energy Environ. Sci. 2018, 11, 3176–3182. [Google Scholar] [CrossRef]
  117. Charreteur, F.; Jaouen, F.; Dodelet, J.-P. Iron porphyrin-based cathode catalysts for PEM fuel cells: Influence of pyrolysis gas on activity and stability. Electrochim. Acta 2009, 54, 6622–6630. [Google Scholar] [CrossRef]
  118. Choi, C.H.; Baldizzone, C.; Polymeros, G.; Pizzutilo, E.; Kasian, O.; Schuppert, A.K.; Ranjbar Sahraie, N.; Sougrati, M.-T.; Mayrhofer, K.J.J.; Jaouen, F. Minimizing Operando Demetallation of Fe-N-C Electrocatalysts in Acidic Medium. ACS Catal. 2016, 6, 3136–3146. [Google Scholar] [CrossRef]
  119. Liu, H.; Zeng, W.; Yang, Y.; Chen, J.; Zhao, Y.; Mu, S. Synchronously improved graphitization and surface area in a 3D porous carbon network as a high capacity anode material for lithium/sodium-ion batteries. J. Mater. Chem. A 2021, 9, 1260–1268. [Google Scholar] [CrossRef]
  120. Gao, W.; Havas, D.; Gupta, S.; Pan, Q.; He, N.; Zhang, H.; Wang, H.-L.; Wu, G. Is reduced graphene oxide favorable for nonprecious metal oxygen-reduction catalysts? Carbon 2016, 102, 346–356. [Google Scholar] [CrossRef]
  121. Lee, J.M.; Han, H.; Jin, S.; Choi, S.M.; Kim, H.J.; Seo, M.H.; Kim, W.B. A Review on Recent Progress in the Aspect of Stability of Oxygen Reduction Electrocatalysts for Proton-Exchange Membrane Fuel Cell: Quantum Mechanics and Experimental Approaches. Energy Technol. 2019, 7, 1900312. [Google Scholar] [CrossRef]
  122. Fu, X.; Hassan, F.M.; Zamani, P.; Jiang, G.; Higgins, D.C.; Choi, J.-Y.; Wang, X.; Xu, P.; Liu, Y.; Chen, Z. Engineered architecture of nitrogenous graphene encapsulating porous carbon with nano-channel reactors enhancing the PEM fuel cell performance. Nano Energy 2017, 42, 249–256. [Google Scholar] [CrossRef]
  123. Xia, D.; Yang, X.; Xie, L.; Wei, Y.; Jiang, W.; Dou, M.; Li, X.; Li, J.; Gan, L.; Kang, F. Direct Growth of Carbon Nanotubes Doped with Single Atomic Fe–N4 Active Sites and Neighboring Graphitic Nitrogen for Efficient and Stable Oxygen Reduction Electrocatalysis. Adv. Funct. Mater. 2019, 29, 1906174. [Google Scholar] [CrossRef]
  124. Han, G.-F.; Xiao, B.-B.; Kim, S.-J.; Li, F.; Ahmad, I.; Jeon, I.-Y.; Baek, J.-B. Tuning edge-oxygenated groups on graphitic carbon materials against corrosion. Nano Energy 2019, 66, 104112. [Google Scholar] [CrossRef]
  125. Zhu, J.; Mu, S. Defect Engineering in Carbon-Based Electrocatalysts: Insight into Intrinsic Carbon Defects. Adv. Funct. Mater. 2020, 30, 2001097. [Google Scholar] [CrossRef]
  126. Jiang, Y.; Yang, L.; Sun, T.; Zhao, J.; Lyu, Z.; Zhuo, O.; Wang, X.; Wu, Q.; Ma, J.; Hu, Z. Significant Contribution of Intrinsic Carbon Defects to Oxygen Reduction Activity. ACS Catal. 2015, 5, 6707–6712. [Google Scholar] [CrossRef]
  127. Proietti, E.; Jaouen, F.; Lefèvre, M.; Larouche, N.; Tian, J.; Herranz, J.; Dodelet, J.-P. Iron-based cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells. Nat. Commun. 2011, 2, 416. [Google Scholar] [CrossRef]
Figure 1. Typical strategies for improving the ORR performance of M-N-C catalysts. Images for “Carbon-rich vacancies increase the density of active sites”: Reproduced with permission [46]. Copyright 2023, Elsevier. Images for “Space-charged localization enhances electronic effect”: Reproduced with permission [47]. Copyright 2025, Elsevier. Images for “Potent-trapping strategy enhances intrinsic activity”: Reproduced with permission [48]. Copyright 2023, Wiley-VCH. Images for “Defect regulation strategies enhance antioxidant properties”: Reproduced with permission [49]. Copyright 2020, Elsevier.
Figure 1. Typical strategies for improving the ORR performance of M-N-C catalysts. Images for “Carbon-rich vacancies increase the density of active sites”: Reproduced with permission [46]. Copyright 2023, Elsevier. Images for “Space-charged localization enhances electronic effect”: Reproduced with permission [47]. Copyright 2025, Elsevier. Images for “Potent-trapping strategy enhances intrinsic activity”: Reproduced with permission [48]. Copyright 2023, Wiley-VCH. Images for “Defect regulation strategies enhance antioxidant properties”: Reproduced with permission [49]. Copyright 2020, Elsevier.
Nanomaterials 15 01257 g001
Figure 2. (a) Structure image of M-N-C catalysts. Reproduced with permission [50]. Copyright 2021, Wiley-VCH. (b) ORR hybridization process image. (c) Calculated charge density differences for M-N-C catalysts. (b,c) Reproduced with permission [55]. Copyright 2023, Wiley-VCH.
Figure 2. (a) Structure image of M-N-C catalysts. Reproduced with permission [50]. Copyright 2021, Wiley-VCH. (b) ORR hybridization process image. (c) Calculated charge density differences for M-N-C catalysts. (b,c) Reproduced with permission [55]. Copyright 2023, Wiley-VCH.
Nanomaterials 15 01257 g002
Figure 3. (a) Schematic depiction of three distinct oxygen molecule adsorption models. Reproduced with permission [57]. Copyright 2020, Royal Society of Chemistry. (b) Diagram depicting the *O pathway and the 2*OH reaction mechanism. Reproduced with permission [58]. Copyright 2020, American Chemical Society.
Figure 3. (a) Schematic depiction of three distinct oxygen molecule adsorption models. Reproduced with permission [57]. Copyright 2020, Royal Society of Chemistry. (b) Diagram depicting the *O pathway and the 2*OH reaction mechanism. Reproduced with permission [58]. Copyright 2020, American Chemical Society.
Nanomaterials 15 01257 g003
Figure 4. Volcano image of M-N-C catalysts. Reproduced with permission [63]. Copyright 2013, Royal Society of Chemistry.
Figure 4. Volcano image of M-N-C catalysts. Reproduced with permission [63]. Copyright 2013, Royal Society of Chemistry.
Nanomaterials 15 01257 g004
Figure 5. (a) The catalyst models and charge densities for FePc/G, FePc/NG, and FePc/G585. Reproduced with permission [74]. Copyright 2019, Elsevier. (b) Preparation illustrations of Fe-SAs@CTCs. (c) Typical structural configurations observed at the interface during the ORR process in alkaline environments. (b,c) Reproduced with permission [62]. Copyright 2023, Springer Nature.
Figure 5. (a) The catalyst models and charge densities for FePc/G, FePc/NG, and FePc/G585. Reproduced with permission [74]. Copyright 2019, Elsevier. (b) Preparation illustrations of Fe-SAs@CTCs. (c) Typical structural configurations observed at the interface during the ORR process in alkaline environments. (b,c) Reproduced with permission [62]. Copyright 2023, Springer Nature.
Nanomaterials 15 01257 g005
Figure 7. (a) Atomic structures and O2 adsorption at the ORR active sites of Fe-N4-C10 (D1), Fe-N4-C12 (D2), and Fe-N4-C8 (D3); (b) atomic structures of the *OOH decomposition reaction’s initial, transition, and final states on D1, D2, and D3; (c) optimized free energy profiles for ORR; (d) PDOS for different catalysts. (ad) Reproduced with permission [95]. Copyright 2017, American Chemical Society.
Figure 7. (a) Atomic structures and O2 adsorption at the ORR active sites of Fe-N4-C10 (D1), Fe-N4-C12 (D2), and Fe-N4-C8 (D3); (b) atomic structures of the *OOH decomposition reaction’s initial, transition, and final states on D1, D2, and D3; (c) optimized free energy profiles for ORR; (d) PDOS for different catalysts. (ad) Reproduced with permission [95]. Copyright 2017, American Chemical Society.
Nanomaterials 15 01257 g007
Figure 8. Factors contributing to the degradation of Fe-N-C catalyst performance. Images for “Fe atoms leaching” and “Carbon corrosion”: Reproduced with permission [101]. Copyright 2015, Wiley-VCH. Images for “N-group protonation”: Reproduced with permission [102]. Copyright 2011, American Chemical Society. Images for “Micropore flooding”: Reproduced with permission [103]. Copyright 2017, Royal Society of Chemistry. Images for “Micropore blockage”: Reproduced with permission [104]. Copyright 2018, American Chemical Society. Images for “Resistance loss”: Reproduced with permission [105]. Copyright 2017, Elsevier. Images for “Al–air battery”: Reproduced with permission [106]. Copyright 2024, American Chemical Society.
Figure 8. Factors contributing to the degradation of Fe-N-C catalyst performance. Images for “Fe atoms leaching” and “Carbon corrosion”: Reproduced with permission [101]. Copyright 2015, Wiley-VCH. Images for “N-group protonation”: Reproduced with permission [102]. Copyright 2011, American Chemical Society. Images for “Micropore flooding”: Reproduced with permission [103]. Copyright 2017, Royal Society of Chemistry. Images for “Micropore blockage”: Reproduced with permission [104]. Copyright 2018, American Chemical Society. Images for “Resistance loss”: Reproduced with permission [105]. Copyright 2017, Elsevier. Images for “Al–air battery”: Reproduced with permission [106]. Copyright 2024, American Chemical Society.
Nanomaterials 15 01257 g008
Figure 10. (a) Schematic illustrations and stability of PANI-Fe-MCS. Reproduced with permission [122]. Copyright 2017, Elsevier. (b) Schematic illustrations and stability. Reproduced with permission [123]. Copyright 2019, Wiley-VCH. (c) Schematic illustrations and stability of EOG-O2, EOG-H2O, EOG-CO2, EOG-H2. Reproduced with permission [124]. Copyright 2019, Elsevier.
Figure 10. (a) Schematic illustrations and stability of PANI-Fe-MCS. Reproduced with permission [122]. Copyright 2017, Elsevier. (b) Schematic illustrations and stability. Reproduced with permission [123]. Copyright 2019, Wiley-VCH. (c) Schematic illustrations and stability of EOG-O2, EOG-H2O, EOG-CO2, EOG-H2. Reproduced with permission [124]. Copyright 2019, Elsevier.
Nanomaterials 15 01257 g010
Table 1. A systematic comparative analysis of the performance of ORR catalysts documented in the recent literature in 0.1 M KOH.
Table 1. A systematic comparative analysis of the performance of ORR catalysts documented in the recent literature in 0.1 M KOH.
CatalystE1/2 (V vs. RHE)Tafel Slope (mV dec−1)Stability
(ΔE1/2 After CV Cycling)
Reference
Fe-N-CNH4I0.925212 mV@20K cycles[20]
FeSA-N-C0.89--6 mV@5K cycles[21]
FeNx/NC-S0.9261--[22]
FeN4-O-NCR0.94545 mV@5K cycles[23]
FeNSC-2Fe0.91597 mV@8K cycles[24]
FeN4-Ten0.8786--[25]
Fe-ISAs/CN0.90522 mV@5K cycles[26]
SA-Fe-NHPC0.90721 mV@10K cycles[27]
Fe-N/C-SAC0.91527 mV@5K cycles[28]
FeSA/B, N-CNT0.9362--[29]
Fe-SAs/N-C0.90562 mV@10K cycles[30]
FeN4Cl1/NC0.91362 mV@5K cycles[31]
MS-CoSA-NC0.868914 mV@10K cycles[32]
Zr-N/O-C0.91715 mV@10K cycles[33]
S-Cu-ISA/SNC0.92502 mV@5K cycles[34]
CoN4/NG0.8770--[35]
Mn-N-C-OAc0.946411 mV@5K cycles[36]
O-Zr-N-C0.91662 mV@40K cycles[37]
Co1-N3PS/HC0.92311 mV@10K cycles[38]
MnSAs/S-NC0.92621 mV@5K cycles[39]
Cu-SA/SNC0.89--1 mV@10K cycles[40]
Mn SAs-N40.90690 mV@8K cycles[41]
Mn-SAS/CN0.91690 mV@5K cycles[42]
Zn-SAs/UNCNs0.91386 mV@5K cycles[43]
Ce SAs/PSNC0.90474 mV@5K cycles[44]
Cu-Se DAs0.913116 mV@10K cycles[45]
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

Chen, J.; Li, Z.; Du, X.; Wang, M.; Li, S.; Wang, Q.; Zhou, Y.; Lai, Y. Recent Advances in Improving the Alkaline Oxygen Reduction Performance of Atomically Dispersed Metal–Nitrogen–Carbon Catalysts. Nanomaterials 2025, 15, 1257. https://doi.org/10.3390/nano15161257

AMA Style

Chen J, Li Z, Du X, Wang M, Li S, Wang Q, Zhou Y, Lai Y. Recent Advances in Improving the Alkaline Oxygen Reduction Performance of Atomically Dispersed Metal–Nitrogen–Carbon Catalysts. Nanomaterials. 2025; 15(16):1257. https://doi.org/10.3390/nano15161257

Chicago/Turabian Style

Chen, Jian, Zheng Li, Xiong Du, Mengran Wang, Simin Li, Qiyu Wang, Yangen Zhou, and Yanqing Lai. 2025. "Recent Advances in Improving the Alkaline Oxygen Reduction Performance of Atomically Dispersed Metal–Nitrogen–Carbon Catalysts" Nanomaterials 15, no. 16: 1257. https://doi.org/10.3390/nano15161257

APA Style

Chen, J., Li, Z., Du, X., Wang, M., Li, S., Wang, Q., Zhou, Y., & Lai, Y. (2025). Recent Advances in Improving the Alkaline Oxygen Reduction Performance of Atomically Dispersed Metal–Nitrogen–Carbon Catalysts. Nanomaterials, 15(16), 1257. https://doi.org/10.3390/nano15161257

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