Hengqi Liua,
Rui Xiongb,
Shengyu Maa,
Ran Wangc,
Zhiguo Liua,
Tai Yaoc and
Bo Song
*acdefgh
aSchool of Physics, Harbin Institute of Technology, Harbin 150001, China. E-mail: songbo@hit.edu.cn
bSchool of Future Technology, Harbin Institute of Technology, Harbin 150001, China
cNational Key Laboratory of Science and Technology on Advanced Composites in Special Environments, Harbin Institute of Technology, Harbin 150001, China
dNational Key Laboratory of Laser Spatial Information, Harbin Institute of Technology, Harbin 150001, China
eNational Key Laboratory of Space Environment and Matter Behaviors, Harbin Institute of Technology, Harbin 150001, China
fLaboratory for Space Environment and Physical Sciences, Harbin Institute of Technology, Harbin 150001, China
gZhengzou Research Institute, Harbin Institute of Technology, Zhengzhou 450046, China
hFrontiers Science Center for Matter Behave in Space Environment, Harbin Institute of Technology, Harbin 150001, China
First published on 18th November 2024
The oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) are crucial reactions in energy storage. However, the sluggish rate of these oxidation electrode reactions and the strong dependence of these technologies on precious metal-based electrocatalysts has greatly restricted further progress. In response to this challenge, researchers have widely investigated the preparation of high-performance ORR and OER electrocatalysts using non-precious metals, reporting substantial advancements in the last ten years. This article provides a concise overview of the latest advancements in oxygen electrocatalysts that are not based on precious metals. The review focuses on the benefits and drawbacks of carbon materials, transition metal compounds, and their composite structures. Moreover, the inherent sources of activity in these materials, techniques for enhancing the density and usage of active sites, and novel design approaches and regulation methods rooted in response mechanisms are examined. Then, a statistical examination of documented bifunctional electrocatalysts is carried out to reveal the correlation between composition, structure, and performance. This report provides a comprehensive analysis of catalyst preparation, element selection, and future directions, delivering significant insights to guide future research endeavors.
After decades of development, lithium-ion batteries have become widely utilized.9,10 However, these batteries increasingly struggle to meet the growing demands for higher energy density, longer lifespan, and lower costs, and they also face challenges such as limited resources, high production expenses, and safety concerns.11,12 In contrast, rechargeable zinc–air batteries (ZABs) are considered a promising alternative due to their high energy density (1086 W h kg−1), low cost, and intrinsic safety.13,14 Nevertheless, the kinetic limitations of the oxygen electrode during charge and discharge processes remain a significant barrier to the widespread adoption of ZABs.15 As a result, improving the stability and efficiency of oxygen electrode catalysts has become a central research focus. Given the complex mechanisms of ORR and OER, different catalysts are typically required for each process.16,17 Precious metal catalysts such as platinum–carbon (Pt/C), iridium oxide (IrO2), and ruthenium oxide (RuO2) exhibit exceptional half-reaction catalytic efficiencies but are insufficient for simultaneously catalyzing both ORR and OER in rechargeable ZABs.18,19 While it is possible to create bifunctional catalysts by combining ORR and OER active components, simply mixing these components together often does not result in achieving accurate catalytic selectivity for both reactions.20 Moreover, the practical implementation of ZABs is hindered by considerable hurdles such as high costs, limited catalytic activity, and poor durability.21 Recently, catalysts based on transition metals and carbon materials have attracted considerable attention as cost-effective alternatives to precious metals.22,23 Strategies such as surface functionalization,24 structural engineering,25 heteroatom doping,26 and defect engineering27 have been explored to enhance bifunctional activity and stability of these catalysts. However, non-precious metal bifunctional electrocatalysts still require substantial improvements to surpass the performance benchmarks set by precious metals. Addressing this critical challenge is urgent. It requires a comprehensive review of current advancements in ORR/OER electrocatalysts. This review should summarize existing progress and guide future research directions.
This paper provides a comprehensive review of the recent advancements in non-precious metal bifunctional oxygen electrocatalysts, focusing on the benefits and drawbacks of carbon materials, transition metal compounds, and their composite structures in catalytic processes. The review focuses on the inherent sources of activity in these materials, techniques to increase the density and usage of active sites, and design principles along with regulation approaches based on reaction processes. The regulation outcomes are displayed through linear sweep voltammetry (LSV) curves, which allow for a systematic evaluation of the ORR activity (E1/2, the potential at which the current density reaches half of its maximum value), OER activity (Ej10, the potential at 10 mA cm−2), and key bifunctional performance (ΔE, the difference between E1/2 and Ej10). Additionally, the paper examines catalyst preparation processes, element selection, the application scope of single-component versus composite structures, and future development prospects. The goal of this review is to provide readers with a comprehensive perspective on advancing bifunctional oxygen electrocatalysts.
Fig. 1 (a) The OER in metal–air batteries. Reprinted with permission.31 Copyright 2016, Wiley-VCH. (b) Theoretical specific energies, volumetric energy densities, and nominal cell voltages for various metal anodes in aqueous and non-aqueous metal–air batteries. Reprinted with permission.32 Copyright 2022, Royal Society of Chemistry. (c) Schematic illustration of the first alkaline zinc–air battery. In 1932, George W. Heise and Erwin A. Schumacher developed the first commercial alkaline zinc–air battery product. The battery employed a novel configuration, and utilized caustic soda as the electrolyte to prevent performance degradation. Reprinted with permission.33 Copyright 2022, The Author(s). (d) Schematic of an aqueous rechargeable zinc–air battery at charging status. Reprinted with permission.31 Copyright 2016, Wiley-VCH. (e) Schematic polarization curves of zinc–air cell. The equilibrium potential of the zinc–air cell (black line) is 1.65 V, but the practical voltage (red line) in discharge is lower than 1.65 V due to the sluggish ORR. A large potential is needed to charge zinc–air battery, higher than the equilibrium potential (blue line). Reprinted with permission.13 Copyright 2021, Wiley-VCH. (f) Schematic illustration of the proposed AEM pathway of OER in alkaline media on an active metal site. Reprinted with permission. Reprinted with permission.34 Copyright 2022, Royal Society of Chemistry. (g) Relationship between oxygen binding energy and catalytic activity of metals. (h) Relationship between catalytic activity and ΔGO/ΔGOH of metal oxides. Reprinted with permission.35 Copyright 2021, Elsevier Ltd. |
In contrast to many recent academic discoveries, the development of ZABs has taken nearly 140 years.32 This voyage began with the discovery of electrochemical principles in the early nineteenth century. In 1878, French engineer L. Maiche created the first zinc–air battery. In 1932, George W. Heise and Erwin A. Schumacher developed an improved alkaline zinc–air battery with enhanced performance, laying the groundwork for future technological developments in this field (Fig. 1(b)).36 From 1960 to 1990, ZABs became widely utilized in the U.S. space program.37 Advances in manufacturing processes and materials science further accelerated the development of these batteries.38 Nevertheless, between 1990 and 2010, further advancements were paused due to technological obstacles such as cathode catalyst instability, anode corrosion, and the rise of lithium-ion batteries. Consequently, research efforts shifted towards exploring alternative battery technologies.39,40 Since 2010, a resurgence of interest in ZABs has emerged due to breakthroughs in electrochemical theory and materials research.39 ZABs are extensively utilized in navigation equipment, signal lighting, maritime exploration, and railway signaling owing to their distinctive operating principles (Fig. 1(c)),33 high energy density, good safety, and environmental friendliness.
The primary components of ZABs include a zinc metal anode, an air electrode serving as the cathode (which consists of a gas diffusion layer, current collector, and catalyst layer), a separator, and an electrolyte (Fig. 1(d)).31 During the discharge process, the zinc anode suffers oxidation due to its interaction with hydroxide ions (OH−). This interaction to produces Zn(OH)42−, which then decomposes into ZnO. At the same time, oxygen from the air passes through the gas diffusion layer and attaches to the catalyst. Oxygen molecules undergo reduction at the three-phase interface by accepting electrons from the zinc anode, resulting in the formation of OH− ions. Subsequently, the hydroxide ions (OH−) are released from catalyst and move through the electrolyte to reach the zinc anode. As shown in Fig. 1(e), the polarization curves of ZABs demonstrate that the electrocatalysts responsible for ORR and OER in the air cathode play a crucial role in determining overall battery performance.13 However, most current catalysts do not meet the requirements for sustained operation, resulting in a substantial deficit in the actual energy density of zinc–air batteries compared to their theoretical value. Disappointingly, these batteries only achieve 40% to 50% of the anticipated energy density.41 Hence, to prepare ZABs with good performance and consistent stability, effective bifunctional electrocatalysts that can enhance the performance of the cathode must be developed.
The direct four-electron pathway by the following reaction:
O2 + 2H2O + 4e− → 4OH− | (2-1) |
In contrast, the stepwise two-electron process proceeds in two stages:
O2 + H2O + 2e− → HO2− + OH− | (2-2) |
HO2− + H2O + 2e− → 3OH− | (2-3) |
Under real-world conditions, the oxygen reduction process is more intricate than the simplified reactions shown in eqn (2-1)–(2-3) because both pathways can occur simultaneously. Preferably, Optimal ORR processes will preferably proceed via the more efficient four-electron pathway, which exhibits lower kinetic barriers. In contrast, the stepwise two-electron approach inhibits the total activity. The ORR process is initiated by the adsorption of oxygen molecules onto the electrocatalyst. This is followed by a sequence of reactions occurring on the surface or at the interface. These reactions involve the transfer of electrons, rearrangement of molecules, breaking of chemical bonds, and production of products.43 Ultimately, the reaction products are released from the surface of the catalyst. These complex reaction steps mean that the reaction process requires additional energy.
Additionally, because OER and ORR require different operating potentials, the charge states of the same catalyst or intermediate vary, leading to differences in electron transfer numbers for the same reversible steps. To design efficient ORR electrocatalysts, the binding energy of intermediates must be understood and controlled. Volcano plots, which are based on the Sabatier principle, are commonly used to illustrate the relationship between ORR activity and material properties despite their limitations in specific reactions.44 The oxygen binding energy (ΔEo) is a key descriptor for ORR processes, and an ideal catalyst will be located at the peak of the volcano plot.34 According to ORR volcano plot shown in Fig. 1(g), platinum (Pt) shows the best ORR performance. Although the equilibrium potentials for ORR and OER are both 1.23 V, different materials are typically selected as electrocatalysts for these reactions.45 While platinum shows exceptional ORR activity, studies have indicated that platinum has poor OER catalytic efficiency, suggesting different reaction mechanisms for these processes on platinum surfaces. To enhance OER activity, the free energy difference between O and OH (ΔGO − ΔGOH) must be optimized toward the peak position on the volcano plot shown in Fig. 1(h).46 Metal oxides are commonly used in OER electrocatalysts, and optimal activity is observed when ΔGO − ΔGOH is around 1.6 eV. Excessively high ΔGO − ΔGOH values lead to significant overpotential during O desorption, while low chemisorption energy results in a high overpotential during OH desorption.47 Thus, achieving both efficient bifunctional activity in a single catalyst remains a challenge.
During the charging process in ZABs, OER occurs at the air electrode.35 ZABs utilize alkaline solutions as electrolytes, and the OER in these batteries is a multi-step process that involves several electron transfer steps and the generation of intermediate. Each individual step has a significant influence on the total reaction efficiency. First, hydroxide ions (OH−) attach to the electrode catalyst, creating hydroxy intermediates (OH*). Subsequently, these intermediate undergo a reaction with additional hydroxide ions (OH−) to generate hydroperoxide intermediates (OOH*). Next, O–O bonds are generated. This step has significant influence on the overall reaction efficiency and determines the reaction rate. The hydroperoxide intermediates subsequently undergo decomposition, resulting in the release of oxygen molecules (O2). Consequently, active sites become available for further involvement in the OER cycle. The regenerated OH* intermediates initiate the cycle again, ensuring that the OER process proceeds uninterrupted. Thus, effective oxygen generation is promoted during battery recharging (Fig. 1(f)).48
In 2009, the ability of vertically aligned nitrogen-doped carbon nanotube arrays (VA-NCNTs) to perform the ORR under alkaline conditions was reported. Since then, nitrogen-doped carbon has garnered significant attention as a single-component bifunctional electrocatalyst. Adding nitrogen atoms changes how oxygen (O2) adsorbs onto carbon nanotubes. In pure carbon nanotubes (CCNTs), oxygen adsorbs end-on. In nitrogen-doped carbon nanotubes (NCNTs), it adsorbs side-on.54 The parallel diatomic adsorption mode efficiently weakens the oxygen–oxygen (O–O) bond, which promotes the ORR process on the surface of NCNTs. Subsequently, Wang et al. developed nitrogen-doped carbon spheres (N-CS) with a microporous structure via polymerization and pyrolysis methods. Using ZnCl2 and a foaming agent greatly enhanced the specific surface area of N-CS. Compared to Pt/C, N-CS-1 exhibited exceptional performance in ZABs.55 Meanwhile, Sun and colleagues doped pyridine nitrogen at graphene edge sites with precision. They developed a new nitrogen-doped graphene material. This material exhibited significantly improved bifunctional electrocatalytic performance.56
The chemical configuration and concentration of nitrogen atoms, including pyridine nitrogen, pyrrolic nitrogen, graphitic nitrogen, and oxidized nitrogen, have significant effects on the activity of nitrogen-doped carbon-based bifunctional electrocatalysts. Scientists have thoroughly investigated the catalytic activity associated with these various forms of doped nitrogen. Guo et al. used model catalyst construction and density functional theory (DFT) calculations to prove that carbon atoms adjacent to pyridine nitrogen serve as active sites for ORR. This finding provides theoretical support for the development of nitrogen-doped carbon catalysts with high activities.57 In addition to nitrogen-doped carbon materials, various heteroatoms can be integrated into carbon matrices to induce charge redistribution. Carbon-based substances doped with boron, sulfur, and phosphorus have been successfully produced and utilized as non-metallic bifunctional electrocatalysts. For instance, Huang et al. designed a porous electrocatalyst (mf-pClNC) without any metal components. Both chlorine and nitrogen were introduced into this electrocatalyst. As shown in Fig. 2(a), this modification greatly improved the catalytic activity for ORR. The addition of chlorine enhanced the proportion of sp3-hybridized carbon atoms and improved the adsorption and desorption properties of the involved substrates and intermediates.58 Subsequently, a range of non-metallic bifunctional carbon-based electrocatalysts with dual or multiple dopants has been synthesized, including binary (e.g., N–S,59,60 N–P,61 N–B62) and multiple (e.g., N–S–P,63 N–S–B64) dopant carbon materials. These studies demonstrate that secondary doping is an effective strategy for enhancing oxygen electrode catalytic activity.
Fig. 2 (a) The fabrication process of the metal-free porous electrocatalyst doped with Cl (mf-pClNC) electrocatalysts. Reprinted with permission.58 Copyright 2023, Wiley-VCH. (b) Schematic of the ORR mechanism for the model of the carbon vacancy coupling adjacent pentagons (VP) sites. Reprinted with permission.65 Copyright 2023, American Chemical Society. (c) Design and fabrication schematic diagram of high-defect density porous carbon (HDPC). Reprinted with permission.66 Copyright 2022, Elsevier. |
To enhance the catalytic activity of carbon-based materials, researchers have extensively explored various methods for introducing defects. Xia et al. fabricated two-dimensional carbon nanosheet (VP/CN) electrocatalysts with a high concentration of defects utilizing a template-assisted method.65 NaCl and Zn(NO3)2·6H2O served as templates, and after heat treatment and acid washing, a material with adjacent pentagon structures and carbon vacancies was obtained. The combined effect of these structures facilitated a two-site pathway during the ORR process, as depicted in Fig. 2(b). Compared to the single-site pathway, this dual-site pathway avoided the formation of the OOH intermediate, which lowered the reaction energy barrier. Additionally, the synergistic interaction between carbon vacancies and pentagon sites further facilitated O–O bond cleavage. As a result, this catalyst achieved a E1/2 of 0.86 V for ORR, nearly matching the 0.87 V of Pt/C. To the template method, substrate corrosion is another effective approach for introducing defects. Wu et al. investigated the diffusion and transformation behavior of CO2 in carbon cavities using finite element methods (FEM) and in situ thermogravimetric-mass spectrometry (TG-MS). High-density porous carbon (HDPC) was successfully synthesized by controlling defect density through carrier gas flow adjustment to provide the carbon cavities.66 This interfacial self-corrosion strategy significantly increased ORR active sites, offering an solution for enhancing the catalytic activity of carbon-based nanomaterials. The synthesized HDPC exhibited a E1/2 of 0.90 V. This research provides a promising and versatile strategy for developing a new generation of electrocatalysts.
The combination of defects with doping can significantly amplify the activity of metal-free carbon-based materials. Jiang et al. employed a spontaneous gas foaming method to synthesize ultrathin nitrogen-doped carbon nanosheets (NCNs).69 By optimizing the carbonization temperature and precursor mass ratio, they precisely controlled the morphology, resulting in a rich porous structure that enabled simultaneous ORR, OER, and hydrogen evolution reaction (HER) catalysis. The active sites were primarily located at the armchair edges near carbon atoms doped with graphitic nitrogen. The NCN-1000-5 sample exhibited trifunctional activity. As an air electrode for ZABs, NCN-1000-5 demonstrated a higher performance.
In summary, the introduction of defects in carbon-based materials increases the number of active sites. This provides bifunctional catalytic activity, even in carbon materials not modified with any additional elements. The discovery of these catalysts has enhanced our comprehension of edge chemistry and defect chemistry. However, accurately synthesizing and characterizing carbon defect structures at the atomic level continues to be a difficult challenge. Furthermore, because defects are frequently generated during heteroatom doping, it is crucial to differentiate between the impacts of defects and doping in order to comprehend fundamental chemical principles. This is particularly important when determining active structures such as dopant type, location, and distribution as well as the specific catalytic functions of these active structures. Additionally, the ORR activity of these catalysts still falls short of that seen in transition metal-containing catalysts, especially in acidic electrolytes, which continues to limit their practical application in energy conversion devices.
Fe–N–C electrocatalysts display superior ORR performance but show limited OER activity. In comparison to other M–N–C catalysts, Co–N–C exhibits improved OER performance and maintains robust ORR activity, as demonstrated by the theoretical calculations reported by Xu et al.74 Experimentally, Co–N–C has been extensively explored as a bifunctional catalyst. Sun et al. synthesized a Co–N–C with porous structure (Co–N–C/AC) by simultaneously heating a metal–organic framework (MOF) precursor and ammonium chloride (AC).75 The ammonium chloride enhanced the active surface area and enabled the precise control of active sites by adjusting the precursor ratio and temperature. Thus, pyridinic nitrogen sites were optimized and new cobalt species were generated. The resulting Co–N–C/AC catalyst exhibited excellent bifunctional performance, with a low potential difference of 0.72 V. Zhang et al. successfully synthesized oxygen-rich cobalt–nitrogen–carbon (O–Co–N/C) porous nanosheets using template-guided CoO@ZIF-8 nanosheet growth and high-temperature thermal treatment (Fig. 3(a)).76 The O–Co–N/C electrocatalyst, which contained Co–N–C and Co–O–C active centers and an ultra-thin porous carbon structure, exhibited an E1/2 of 0.85 V (ORR) and an OER overpotential of 0.29 V. Liang et al. first reported an onion-like carbon-coated cobalt and nitrogen-doped carbon material (OLC/Co–N–C) prepared by including surfactant micelles in the precursor, which resulted in a multilayer high-curvature nanocatalyst with a mesoporous structure (Fig. 3(b) and (c)).77 This OLC/Co–N–C catalyst demonstrated effective bifunctional ORR/OER activity due to its combined electrical and structural effects. Furthermore, the presence of the curved graphitic carbon significantly increased the reactivity of metal–carbon atoms in the vicinity of the graphitic nitrogen as well as that of ortho/meta carbon atoms near pyridinic nitrogen (Fig. 3(d)).
Fig. 3 (a) Illustration of the preparation steps of O–Co–N/C nanosheets. Reprinted with permission.76 Copyright 2022, Wiley-VCH. (b) SEM image and (c) HRTEM image of OLC/Co–N–C. (d) Structural reconstruction of NiFe-CNG during OER from separate Ni and Fe centers to Ni–O–Fe/Ni moieties. Reprinted with permission.77 Copyright 2021, Wiley-VCH. (e) Contour plot of OER overpotential as a function of Gibbs adsorption energies ΔG*OOH − ΔG*O along the x-axis and ΔG*OOH along the y-axis. The inset is the scaling relation ΔG*OOH = 0.84 (ΔG*OOH − ΔG*O) + 1.75. Reprinted with permission.78 Copyright 2023, The Author(s). (f) Structural reconstruction of NiFe-CNG during OER from separate Ni and Fe centers to Ni–O–Fe/Ni moieties. Reprinted with permission.79 Copyright 2021, The Author(s). Projected density of states of Co 3d and Cu 3d orbitals in (g) CoN4 and CuN4/CoN4, (h) CuN4 and CuN4/CoN4, and the corresponding d-band centers. (i) Charge density difference and Bader charge analysis of O2 adsorbed on CuN4/CoN4 (Co), CuN4/CoN4 (Cu), CuN4, and CoN4. Reprinted with permission.80 Copyright 2022, Wiley-VCH. |
Aside from Co–N–C, other M–N–C systems have shown considerable promise as bifunctional oxygen electrode catalysts. Liang et al. presented design guidelines for SACs based on graphene, using theoretical models and tests to confirm the influence of the local structure and chemical environment of Ni–N–C centers on catalytic activity.81 Jin and his coworkers created a catalyst made from a biomass aerogel precursor that is atomically distributed and doped with Mn and N in carbon aerogel.82 This catalyst demonstrated exceptional efficacy, indicating its potential for application in ZABs. Mei and colleagues developed a core–shell nanostructured catalyst consisting of MnO@Cu–N–C.83 Due to the combined influence of MnO and Cu–N–C, this catalyst demonstrated effective electron transport properties and a smaller overpotential. As a result, greatly improved ZAB performance was achieved. However, despite the significant advancements reported in the development of bifunctional oxygen electrode catalysts based on M–N–C materials, simultaneously enhancing ORR and OER activity within the same reaction mechanism is still a notable challenge. Rossmeisl et al. used theoretical calculations to predict the efficiency of M–N4 active sites.84 Their volcano plots indicated that selecting an active site requires an optimal balance between the adsorption energies of different reaction intermediates. However, because the optimal Gibbs free energy values differ for ORR and OER, optimizing one reaction often compromises the activity of the other reaction. Thus, simultaneously achieving high efficiency for both reactions is difficult.
One of the most effective solutions for reducing the energy requirements of the OER pathways is adjusting the coordination environment of metal atoms. Key measures include modifying the coordination number of metal atoms in the first coordination shell and introducing dopants or structural defects in the second coordination shell. For instance, Li et al. precisely tuned the coordination environment of iron SACs by employing a polymer coating, wet chemical adsorption, ammonia treatment, and calcination strategy.7 This approach resulted in the preparation of an atomically dispersed catalyst with an Fe1N4O1 structure. By adjusting the coordination environment of the iron atoms, the electronic metal-support interaction was optimized, enhancing the adsorption and desorption capabilities of this catalyst for oxygen intermediates. Fe1N4O1 demonstrated excellent ORR activity and a low OER overpotential across the entire pH range, outperforming the Pt/C + RuO2 benchmark in terms of ZAB performance and stability. In addition, the sluggish oxygen–oxygen coupling process in Fe–N–C materials leads to low OER activity. Chen et al. synthesized a phosphorus-containing Fe–N–C catalyst (P/Fe–N–C) by carbonizing iron-doped zeolitic imidazolate frameworks.78 Spectroscopic characterization revealed that an FeN4 structure was embedded in graphite carbon. The P/Fe–N–C catalyst exhibited enhanced OER activity, and excellent ORR performance was achieved. OER activity across can be described by Gibbs adsorption energy functions (ΔGOOH − ΔGO and ΔGOOH), where ΔGOOH − ΔGO represents the O–O coupling process (O → OOH), and ΔGOOH corresponds to the O2 release process (OOH → O2). The original Fe–N–C material exhibited an OER as high overpotential, indicating poor OER activity. In contrast, the P/Fe–N–C model displayed improved OER performance due to an increase in ΔGOOH and a decrease in ΔGOOH − ΔGO. Therefore, introducing phosphorus into the second coordination shell optimized the adsorption properties of *OOH and *O, achieving improved OER performance. To further regulate catalytic activity by generating defects in the second coordination shell, Li's group developed a multifunctional molten salt-assisted pyrolysis strategy for the construction of ultrathin porous carbon nanosheets loaded with Co SACs.85 The molten salt induced the formation of Co single atoms and porous graphene-like carbon, which ensured the full exposure of active centers and endowed the Co SACs with abundant defective Co–N4 structures. The prepared Co SACs exhibited excellent bifunctional activity and stability. DFT calculations showed that defects in the second coordination shell of the Co SACs facilitated the desorption of OH intermediates in ORR and the deprotonation of OH in OER. Therefore, these defects served as effective active centers for bifunctional oxygen catalysis.
Constructing dual-atom catalysts is an effective strategy for addressing these challenges. During OER, the synergistic effects of two metal elements can enhance the adsorption and reconstruction of reaction intermediates. Wan et al. proposed a molecular design strategy that facilitated the rational design of both single-center and dual-center catalysts.79 With this strategy, a catalyst featuring single Fe sites demonstrated excellent ORR activity, achieving a E1/2 of 0.89 V. During OER process, the coordination environment around the Ni atoms underwent reconstruction, forming Ni–O–Ni/Fe bonds (Fig. 3(f)). These newly formed Ni–O–Fe bonds provided a spin channel for electron transfer, which led to improved OER activity. A dual-site water oxidation pathway was enabled at Ni–O–Fe bridge sites, where *OH deprotonation at the Ni and Fe sites led to the generation of O2. Compared to single-atom catalysts, dual-atom catalysts retain the benefits of SACs while achieving multifunctional synergistic catalysis through geometric and electronic ligand effects. Catalysts with uniformly structured and flexible active sites show better performance in bifunctional catalysis. Tang et al. developed a Janus-type dual-metal active site catalyst (FeCo–N3O3@C). In this catalyst, the Fe and Co atoms were respectively coordinated with N and O atoms, and the atoms were connected through bridging N and O atoms.86 This unique coordination environment (Fe–N3 and Co–O3) allowed for the precise regulation of bifunctional active sites, optimizing the adsorption and desorption of oxidation intermediates and enhancing reaction kinetics.
Recently, Li's team proposed a method for constructing Cu–Co dual-atom site electrocatalysts on a highly porous nitrogen-doped carbon matrix (Cu–Co/NC).80 The CuN4/CoN4, CuN4/CoN4 system exhibited improved electronic conductivity, with a negative shift in the d-band center of Co leading to the increased occupancy of anti-bonding orbitals and promotion of molecular oxygen dissociation (Fig. 3(g)). Simultaneously, the larger negative shift in the d-band center of Cu may have weakened its binding affinity with molecular oxygen, affecting O2 activation (Fig. 3(h)). The O2 species adsorbed at the Co sites in the CuN4/CoN4 system acquired more charge (0.57 e−) compared to other models, demonstrating higher charge transfer efficiency (Fig. 3(i)). Thus, the synergistic effect of the Cu–Co bimetallic sites resulted in asymmetric charge distribution of oxygen intermediates, leading to moderate adsorption and desorption behavior. By optimizing the coordination environment of the metal active sites, the adsorption and desorption efficiency of oxidation intermediates was significantly improved, enhancing the catalytic performance and reaction kinetics. This catalyst shows excellent potential for application in bifunctional oxygen electrodes.
Overall, SACs can be used to effectively incorporate highly active atomic-level sites within conductive carbon frameworks, showing great promise as candidates for bifunctional oxygen electrocatalysis. The intrinsic activity of M–N–C sites surpasses that of most carbon-based active sites, such as those with non-metal dopants or defects. From the perspective of atomic utilization efficiency, atomically dispersed active sites represent the forefront of technology, approaching theoretical limits. Looking ahead, the well-defined chemical structures of these catalysts provide valuable opportunities for molecular design and activity optimization, which will further highlight their distinct advantages in bifunctional oxygen electrocatalysis. However, the stability of SACs under operational conditions remains a challenge. Therefore, efforts must focus on simplifying the synthesis process, further enhancing activity (particularly OER activity), improving stability, and increasing the content and utilization of active sites.
Fig. 4 (a) Schematic depiction of the hydrophobization engineering of Co3O4 NSs to enable the formation of more three-phase reaction interfaces and promoted oxygen diffusion on the obtained hydroophobic-Co3O4 NSs/CC cathode impregnated with electrolyte, in comparison with untreated-Co3O4 NSs/CC cathode. Reprinted with permission.90 Copyright 2022, Wiley-VCH. (b) and (c) Bader charges of Co (in blue), Mn (in magenta) and O (in red) of Co3O4 (001) and MCO (001) surfaces; (d) and (e) major bond lengths (in black) and formation energies of oxygen vacancy (in green); (f)–(i) top view of the pristine MCO (001) surface as well as the adsorption states of O*, OH*, OOH*, and H2O2* over Co1 site, respectively. Note: the blue, magenta and red balls represent Co, Mn and O elements, and the green dashed circles are the potential position for oxygen vacancies. Reprinted with permission.92 Copyright 2022, Elsevier. (j) Schematic illustration of the liquid-state Zn–air battery device. (k) Polarization curves and power density profiles, and (l) voltage vs. specific capacity curves at a constant current density of 5 mA cm−2 of the liquid-state rechargeable Zn–air batteries. (m) Hierarchical structure of the cable-type Zn–air battery. (n) Polarization curves and power density profiles, and (o) voltage vs. specific capacity curves at a constant current density of 7 mA cm−3 of the cable-type all-solid-state Zn–air batteries. Reprinted with permission.93 Copyright 2021, American Chemical Society. (p) Structures of manganese oxide reported here in this study: α-MnO2 (2 × 2 tunnel), β-MnO2 (1 × 1 tunnel, pyrolusite), δ-MnO2 (layered, birnessite), and amorphous manganese oxides (AMO). Reprinted with permission.94 Copyright 2014, American Chemical Society. (q) Oxygen electrode activities of the nanostructured Mn oxide thin film, nanoparticles of Pt, Ir, and Ru, and the GC substrate. The Mn oxide thin film shows excellent activity for both the ORR and the OER. Reprinted with permission.95 Copyright 2010, American Chemical Society. |
Single metal oxides typically exhibit high OER catalytic activity. However, ORR performance is impeded by the considerably weaker bond between oxygen atoms and the surfaces of TMOs compared to transition metals. This challenge is further exacerbated by the powerful electric field effect on the TMO, which inhibits the breaking of O–O bonds. Nørskov and colleagues screened 7798 TMO compositions for ORR catalysts, reporting that manganese-based oxides show great potential for achieving high activity.96 The catalytic properties of manganese oxides vary based on the structural connectivity of [MnO6] units, which can be linked by shared corners or edges. One common manganese oxide structure is layered δ-MnO2, where MnO6 units share edges and the interlayer spaces are occupied. One-dimensional MnO2 materials, form various configurations based on tunnel size. Manganese oxides with tunnel structures include β-MnO2, γ-MnO2, α-MnO2, OMS-1 (3 × 3), and OMS-5 (2 × 4). Among these, α-MnO2 has been extensively studied due to its superior adsorption, catalytic, and oxidation properties.
Meng's research team facilely synthesized MnO2 nanomaterials with different crystal structures (Fig. 4(p)) and systematically analyzed the structure–performance relationship of these materials as OER and ORR catalysts in alkaline media.94 The exceptional performance of α-MnO2 was attributed to its abundant bridging oxygen bonds, mixed valence states (average oxidation state = 3.7), and low charge transfer resistance (91.8 Ω). Compared to cobalt oxides, manganese oxides exhibit a broader range of valence states, with the presence and ratio of Mn4+ and Mn3+ playing a critical role in determining catalytic activity. Inspired by the oxygen-evolving center (OEC) of Photosystem II, Jaramillo and colleagues mimicked the structure of the active site CaMn4Ox to develop a manganese oxide film with coexisting Mn4+ and Mn3+, as shown in Fig. 4(q).95 This film demonstrated high ORR and OER activity, surpassing Ru and Ir nanoparticles in terms of ORR activity. In terms of OER activity, the manganese oxide film significantly outperformed Pt and approached the activity of Ir and Ru. Beyond crystal structure and valence state, manganese-based oxides can be prepared to selectively expose different crystal facets, which also exhibit varying catalytic activity. He and colleagues found that the MnO (100) facet showed significantly enhanced electrocatalytic performance due to its higher oxygen species adsorption energy.97 They synthesized MnO nanoflowers and multi-legged structures with specific orientations using a thermal decomposition method with oleic acid as a free ligand. The unique MnO exhibited a high mass activity.
Perovskite oxides are materials that can be described by the generic formula ABO3. The A-site ions of perovskites are usually larger alkaline earth or rare earth metals such as calcium, strontium, barium, or lanthanum. These ions are arranged in a twelve-fold polyhedral environment. The B-site ions are often smaller transition metals such as titanium, manganese, iron, or cobalt that are situated in an octahedral coordination environment, where they create a three-dimensional network of BO6 octahedra. Perovskite oxides offer a diverse platform for creating efficient electrocatalysts due to the variety of A-site and B-site ions, which enables the modification and control of oxygen vacancies and modify crystal structure.
Peng's group used oxygen vacancy-rich porous CaMnO3 nanofibers (V-CMO/rGO) as air electrode catalysts for low-temperature wearable ZABs (Fig. 5(a)).98 The synergistic effect of the metal atoms and oxygen vacancies combined with the improved kinetics, conductivity, and mass transfer inside the 3D rGO-coated nanofibers resulted in dramatically improved catalytic activity. The regulated electrical filling encouraged the desorption of ORR intermediates. Thus, V-CMO/rGO enabled the operation of a flexible ZABs at −40 °C, yielding a power density of 56 mW cm−2 and a cycling life of over 80 h. This novel and dynamically active catalyst shows good promise for ZAB applications in hostile settings (Fig. 5(b)).
Fig. 5 (a) Schematic diagram of the synthesis process of the vacancy-rich, reduced graphene oxidecoated porous CaMnO3 (V-CMO/rGO). (b) Photograph of a plasma ball powered by three flexible sandwich-type ZABs under −40 °C. Reprinted with permission.98 Copyright 2023, Wiley-VCH. (c) Chemical and electrochemical characterization of the solid-hydroxy silicon/perovskite oxide. A structural formula of the hydroxy silicon and a schematic diagram showing the deprotonation process on the PBCC surface. (d) DFT calculations and the electrocatalysis mechanism. Proposed OER mechanisms, including a AEM and b LOM. M1 (Co) and M2 (Si) are the adsorption centers of the AEM and LOM in the OER pathway. Other intermediates are also labeled. The empty square represents an oxygen vacancy. Reprinted with permission.99 Copyright 2023, The Author(s). (e) Schematic illustration of the formation process of ACo2O4/NCNTs (A = Mn, Co, Ni, Cu, Zn). (f) Schematic diagram of tetrahedral and octahedral interaction. Reprinted with permission.100 Copyright 2022, Wiley-VCH. |
The OER activity of perovskite oxides primarily depends on the charge transfer energy during the proton-coupled electron transfer (PCET) process. However, the limited energy of lone-pair electrons surrounding the oxygen atoms constrains the pace at which deprotonation occurs. Moreover, simultaneously enhancing OER and ORR is a significant challenge if both reactions follow the adsorbed-electron mechanism (AEM). To address this, Wang et al. introduced hydroxyl-containing solid base BaCaSiO4 (BCS) nanoparticles onto the surface of perovskite oxide nanofibers using a one-step calcination and desolation strategy.99 They investigated the role of Si doping in improving the physicochemical properties and OER activity of this material. During the formation of the BCS proton acceptors, the protons acted as Lewis acids, binding with electron donors. For instance, Ca2+/Ba2+ and SiO2 formed multi-anionic [SiO44−] silicates, where alkaline O2− ions were transferred from weakly acidic Ca2+/Ba2+ to strongly acidic Si4+. This increased the oxygen atom electron density, enhanced hydrogen bond formation, and stabilized the coordinated HO(Si), causing the Si–O–Si bond angle to deviate from 180°. The resulting hydrogen bonds stabilized structures such as OH–Si–O–Si–OH and OH–Si–O–Ca2+/Ba2+–Si–OH, which more readily accepted protons from OH* intermediates. Consequently, a tenfold increase in OER performance was achieved (Fig. 5(c)). Additionally, DFT was used to study the reaction mechanism and proton transfer pathways in the BCS-PBCC system during OER (Fig. 5(d)). DFT analysis revealed that the surface-functionalized Lewis base acted as an electron pair donor, enabling effective coordination with protons and a reduction in the energy barrier of OOH* deprotonation. Thus, kinetic limitations in proton transfer were overcome and the lattice oxygen mechanism (LOM) was enabled, significantly enhancing OER performance. The lattice oxygen mechanism refers to the process in which oxygen atoms in the catalyst lattice participate in oxidation-reduction reactions by accepting or releasing electrons to facilitate chemical reactions.
As bifunctional catalysts, the active sites of perovskite oxides mainly consist of surface oxygen vacancies, cationic defects, and transition metal centers. Crucially active locations such as surface oxygen vacancies can be used to catalyze both ORR and OER. Cationic flaws change the electronic structure of perovskite oxides, which improves activity. Moreover, important catalytic sites in perovskites that accelerate electrocatalytic processes via interactions with reactants include transition metal centers (e.g., Fe, Co, Mn). Thus, research on perovskite catalysts has concentrated on techniques for optimizing their crystal structure, cation control, anion management, nanostructure design, surface engineering, and electrocatalytic performance.
Transition metal oxides with spinel structures (i.e., those based on nickel, cobalt, and manganese) have attracted considerable interest as electrocatalysts, which is mainly due to their unique d-electron configuration and crystal structure. These materials usually adhere to the standard formula AB2O4, with A and B denoting distinct metal ions. The A-site metal ions typically occupy tetrahedral sites, while the B-site metal ions are located in octahedral sites, creating a three-dimensional network. The origins of the active sites and regulation techniques in perovskites are comparable to those in spinel-structured transition metal oxides. The metal ions, oxygen vacancies, and surface electronic structures of spinel oxides make them promising candidates as bifunctional oxygen catalysts.
For example, the spinel-structured catalyst CoFe2O4 exhibit bifunctional catalytic activity, which is often achieved through electronic structure regulation. Li's research group introduced La to adjust the electronic structure within a spinel, increasing the electron density at Fe sites. La acted as an “electron pump,” transferring electrons from Co sites to the catalytically active Fe sites. This optimized the adsorption process and enhanced bifunctional oxygen catalytic performance.101 Consequently, La0.2CoFe1.8O4/3D-G demonstrated outstanding ORR catalytic performance, with an E1/2 = 0.86 V. Moreover, this catalyst also showed a low potential and overpotential in OER (η = 320 mV), reflecting its excellent bifunctional performance (ΔE = 0.69 V). Tang et al. synthesized hollow spinel-type NiCo2O4 nanomaterials using a CTAB-assisted solvothermal method with ZIF-67 as a precursor.102 Highly active three-dimensional porous NiOOH–NiCo2O4 structures were formed on the surface of these catalysts via electrochemical reconstruction, which enhanced ORR and zinc–air battery performance. In situ Raman analysis revealed that during the OER process, this electrochemical reconstruction significantly increased pseudocapacitance compared to ORR. Theoretically, substituting atoms at the tetrahedral sites of AB2O4 spinel oxides optimizes charge distribution at the octahedral sites through ATd–O–BOh interactions. Zhao et al. proposed straightforward solvothermal process to control the structure of spinel oxides and used this method to prepare composite materials (ACo2O4/NCNTs, where A = Mn, Co, Ni, Cu, Zn) for use as oxygen electrocatalysis (Fig. 5(e)).100 The optimized MnCo2O4/NCNTs exhibited high ORR/OER activity and excellent stability. Fig. 5(f) illustrates the ATd–O–Co3+Oh and CoTd–O–Co3+Oh configurations in the spinel structure, where tetrahedral A ions were bonded to octahedral Co ions via a shared oxygen atom. The spin states of ATd and Co3+Oh are also depicted. Metal substitution adjusted the Co3+/Co2+ ratio, modulating the electronic structure. Compared to the other prepared catalysts, Co3+Oh in MnCo2O4 demonstrated more favorable binding with oxygen-containing species, which led to significantly enhanced oxygen electrocatalytic performance. Moreover, combining spinel oxides with other materials (e.g., carbon nanotubes or alloys) can markedly improve conductivity and catalytic activity. Alternatively, a one-step pyrolysis reduction strategy can be employed to prepare spinel oxide-modified alloy composites with superior catalytic performance.
Fig. 6 (a) Synthetic procedure selection for (CoxFe1−x)3N@NPC catalysts. Reprinted with permission.106 Copyright 2023, Wiley-VCH. (b) TEM images of the Ni–Co–Mn–P (NCMP). (c) Comparison of the overpotentials to drive a current density of 10 mA cm−2 for the NCMP catalyst with some other transition metal phosphide-based electrocatalysts namely, (A) Co2P/Co-, N-, and P-doped carbons; (B) CoFeO@black phosphorus; (C) CoP@P, N co-doped carbon; (D) CoP N-doped carbon@CNT; (E) CoP NFs; (F) CoP/T3C2; (G) Mo–Ni3S2/NixPy/NF; (H) (Fe0.1Ni0.9)2P(O)/NF; and (I) NF@Fe2–Ni2P/C. (d) ZABs performance comparison. Reprinted with permission.107 Copyright 2024, Royal Society of Chemistry. (e) and (f) SEM images, (g) HR-TEM images, and (h) HAADF-STEM image of Co3S4/FeS@CoFe/NC. (i) The real-space distributions of spin density (the first row, the blue and orange colors of the electron cloud indicate two different spin orientations, respectively) and charge density difference (the second row, the charge accumulation and consumption regions are represented by yellow and blue, respectively) by DFT calculations for O adsorbed on Co3S4/FeS@CoFe/NC, Co3S4@Co/NC, FeS@Fe/NC, Co3S4/FeS. Dark blue, orange, yellow, blue gray, brown, red and white spheres represent cobalt, iron, sulfur, nitrogen, carbon, oxygen and hydrogen atoms, respectively. Reprinted with permission.108 Copyright 2024, Science Press and Dalian Institute of Chemical Physics. |
In conclusion, transition metal compounds, including oxides, sulfides, nitrides, and phosphides, offer significant advantages as bifunctional oxygen electrode catalysts. The use of these compounds shows substantial potential in electrochemical storage systems. To traditional noble metal, transition metal compounds are abundant, cost-effective, and exhibit excellent chemical and structural tunability. The unfilled d orbitals of transition metals serve as primary active sites, allowing for kinetic optimization between the OER and ORR processes through the modulation of the adsorbed intermediate bonding energies. The coordination environment and chemical bonding characteristics, such as M–O or M–S bonds, strongly influence the electronic structure and catalytic performance of these materials.
Enhancing active site density, improving electrical conductivity, and boosting catalyst stability and resistance to poisoning can be achieved through strategies such as heteroatom doping, the introduction of lattice defects, the regulation of exposed crystal facets, nanostructure optimization, and compositing with conductive carbon materials. Elemental doping optimizes reaction barriers by adjusting the electron density around metal centers, while the introduction of lattice defects and nanostructuring strategies increase the density and exposure of active sites. However, significant challenges remain, including maintaining long-term efficiency and stability under practical conditions, further reducing catalyst costs, and improving scalability. Therefore, in addition to focusing on the discovery of novel transition metal compounds, future research should also seek to gain a deeper understanding of active site mechanisms, reaction pathways, and key rate-limiting steps. This will require a combined approach involving both theoretical calculations and experimental characterization to develop more precise structure–performance tuning strategies. Additionally, validating the performance and durability of these catalysts in full-cell devices is a crucial next step for enabling large-scale commercial applications.
Research on transition metal compound composites frequently focuses on combining various transition metal oxides and hydroxides. Yu et al. developed a CuCo2O4/NiFe LDH catalyst by adjusting the metal ion ratios and valence states within Ni–Fe–Co oxides/hydroxides.117 This approach facilitated the reconstruction of the catalyst surface via charge transfer between Ni, Fe, and Co, which optimized the catalytic performance. Experimental results indicated that ZABs prepared using CuCo2O4/NiFe LDH electrodes demonstrated higher efficiency compared to those prepared using a single CuCo2O4 electrode. Specifically, the charge/discharge voltage gap decreased from 1.02 V to 0.78 V, while the charge/discharge efficiency increased from 54% to 60%. Amorphous materials, which are characterized by a higher number of non-coordinated atoms and defect structures, provide more active sites for electrochemical processes compared to crystalline materials. Moreover, combining amorphous and crystalline phases in heterostructures can leverage the advantages of both states, leading to superior electrocatalytic performance. Zhang's research group drew inspiration from the atomic diffusion and reordering processes that occur when Co(OH)2 is heat-treated to form CoO.118 As shown in Fig. 7(a), a volcano plot was employed to illustrate the thermodynamic limitation potentials (UL) related to ΔGOOH* (for ORR) and ΔGOH* (for OER). At these UL potentials, the free energy diagrams for all fundamental reactions indicated that they were thermodynamically favorable. Ideally, ORR and OER should occur close to these thermodynamic potentials. The observed catalytic activity trend for ORR and OER was as follows: O@sur-1Ov > vac@sur-1Ov > CoO (111) > sub-1Ov. This trend demonstrates that the existence of surface vacancies significantly enhanced the intrinsic ORR/OER activity of CoO. Internal oxygen vacancies improved the performance by changing the electronic structure to facilitate electron transfer. Consequently, the precise control of oxygen defect concentrations and the formation of amorphous/crystalline heterostructures through vacuum calcination (Fig. 7(b) and (c)) effectively enhanced the electrocatalytic performance of CoO. The OER overpotential and ORR half-wave potential difference (ΔE) for this crystalline/amorphous heterostructure were 0.745 V. Additionally, during the OER process, the catalyst surface often reconstructed to form a hydroxide layer that acted as an active site. The synergistic combination of this restructured surface with the original structure provided enhanced catalytic performance.
Fig. 7 (a) Calculated UL ORR (vs. ΔGOOH*) and UL OER (vs. ΔGOH*) of different catalyst models. (b) Schematic illustration for the preparation of ODAC-CoO nanosheets. (c) Enlarged images of the selected regions in (d) and the corresponding FFT patterns in oxygen-defective amorphous-crystalline (ODAC-CoO-30). Reprinted with permission.118 Copyright 2021, Wiley-VCH. (e) and (f) Bragg-filtered image of the (111) planes of Ni and (113) planes of MnFe2O4 at the Ni35%/MnFe2O4 interface. The yellow color arising at the interface and the black curved line in the Ni particle in panel e is an artifact generated by the delocalized signal of the Bragg-filtered (111) planes. Spin polarization difference diagram of NiOOH/MnFeOOH (g) and NiOOH (h). The green contour around the atoms represents spin-up electrons. (i) OER process with and without the spin-aligned process. Reprinted with permission.119 Copyright 2024, Wiley-VCH. |
Yang's team constructed a Ni/MnFe2O4 heterojunction for use in rechargeable ZABs. HAADF-STEM images and EELS elemental distribution maps (Fig. 7(e)) revealed a uniform distribution of both phases, which resulted in the presence of numerous Ni/MnFe2O4 heterojunctions.119 The Bragg-filtered image, shown in the magnified red box region in Fig. 7(f), displays the relationship between the Ni (red) (111) plane and the MnFe2O4 (green) (113) plane at an angle of 66.18°. It was observed that these two crystal domains frequently grew in this relative orientation, reducing the geometric strain to 3%. During OER, the surface of this electrocatalyst reconstructed to NiOOH/MnFeOOH, and the high-spin Ni effectively tuned the spin of oxygen intermediates (Fig. 7(g) and (h)). The NiOOH/MnFeOOH electrocatalyst exhibited an overpotential of 261 mV and showed good stability for over 1000 cycles. Additionally, rechargeable ZABs prepared using this catalyst demonstrated an exceptional stability. The optimal spin state of Ni active sites with oxygen intermediates facilitated spin-selective charge transfer (Fig. 7(i)), optimizing reaction kinetics and reducing the energy barrier for oxygen evolution.
In addition to transition metal oxide and hydroxide composites, the combination of transition metal oxides with other transition metal compounds can also integrate different types of ORR and OER active sites. Xi's team reported NiFe2O4/FeNi2S4 heterostructure nanosheets (HNSs), which were used as efficient catalysts to significantly enhance the recharging performance of neutral aqueous zinc–air batteries.120 The NiFe2O4/FeNi2S4 HNSs featured a surface rich in oxide/sulfide interfaces, which effectively modulated the oxygen binding ability and activity of catalyst. In a 0.2 M PBS, NiFe2O4/FeNi2S4 HNSs exhibited excellent activity and stability. However, most transition metal sulfides, phosphides, and nitrides are typically used in combination with carbon-based materials, which will be discussed in Section 5.4.
In summary, transition metal compound composites offer notable advantages due to their tunable structures. These composites have the potential to achieve synergistic effects, show exceptional stability. However, compared to state-of-the-art bifunctional electrocatalysts, the performance remains limited. This limitation is primarily due to factors such as inadequate conductivity, and excessively high intermediate adsorption free energies during ORR. Consequently, the development of transition metal compound composites is still in its early stages, and further research and optimization are required.
Optimizing the morphology can enhance the accessibility of active sites, while doping heteroatoms into carbon substrates can further optimize conductivity and electronic structure. Ren et al. reported a pyrolysis method utilizing dual MOFs. Specifically, FeNiM@ZnCoZ was utilized as a precursor to synthesize FeNiCo@NC-P anchored onto carbon nanotubes (CNTs).121 Compared to carbon materials derived from single MOFs, this synthesized catalyst incorporated N dopants as well as three metals (Fe, Co, and Ni). Additionally, the two MOF precursors successfully yielded carbon materials with microporous/mesoporous structures while preserving a one-dimensional carbon nanorod morphology during high-temperature pyrolysis. These features provided FeNiCo@NC-P with abundant active sites. Consequently, exceptional bifunctional electrocatalytic activity and outstanding ZAB performance were achieved. The Fe–N sites on the carbon framework exhibited excellent ORR activity, while cobalt nanoparticles significantly enhanced the OER process. The synergistic effect of these features further improved the ORR and OER performance of this catalyst. Xie et al. employed MOFs to synthesize a series of carbon-based catalysts featuring cobalt nanoparticles and highly dispersed iron species with a controlled doping strategy.122 Using ZIF-8 as the precursor, iron and cobalt sources were introduced through in situ doping and a dual-solvent method, which led to the successful anchoring of highly dispersed iron species and the confined growth of cobalt nanoparticles. The optimized catalyst (CoNP@FeNC-0.05) demonstrated a high E1/2 and a low η10 in alkaline electrolytes.
In conventional synthesis of M–N–C structures, the presence of metal particles is unavoidable. These metal particles have incomplete d-orbitals and can readily accept electrons from oxygen reaction intermediates. Thus, activation energy barriers are lowered. However, a significant issue with Fe–N–C materials is their highly symmetrical electronic structure, resulting in sluggish reaction kinetics. Therefore, electronic modulation and orbital engineering are essential for the precise design of advanced Fe–N–C. Zhang et al. prepared a composite (FeSA/AC@HNC) by utilizing a dual-ligand zinc-based zeolite framework to anchor monodisperse iron atoms and adjacent iron clusters within nitrogen-doped layered porous hollow carbon.123 The introduction of iron clusters lowered the electron density of Fe–N4 and increased the effective magnetic moment of single-atom active sites. The FeSA/AC@HNC composite demonstrated excellent ORR activity. The interaction between iron clusters and single atoms also enhanced the stability and performance of this composite in rechargeable zinc–air batteries. A persistent challenge in metal–nitrogen–carbon systems is increasing the metal loading of SACs (>5 wt%). Xu's team employed a dual-melamine pyrolysis method and a layered g-C3N4 sacrificial template to co-anchor Co single atoms and Co clusters on nitrogen-rich porous graphene, achieving a Co single-atom loading of up to 14.0 wt%.124 Due to the modulated electronic structures and optimized chemical bonding of this composite material, catalytic activity was significantly improved. Peng et al. first synthesized iron-based MOFs with a nanotube structure via co-precipitation.125 On this basis, Fe-ACSA@NC was obtained by polymerization and pyrolysis (Fig. 8(a)). The optimized electron redistribution of these materials facilitated the easier desorption of intermediates, resulting in superior ORR activity compared to nanoparticle-modified materials. A metal–air battery was assembled with Fe-ACSA@NC, and this battery demonstrated exceptional power density and capacity.
Fig. 8 (a) Synthesis procedure and morphology characterizations of Fe-ACSA@NC. Reprinted with permission.125 Copyright 2021, Wiley-VCH. (b) Charge density difference between the metal cluster and metal-NC of Fe4@Fe–N4, N4@Ni–N4 and Fe4@Ni/Fe–N4 (Fe4 on the Ni–N4 site). X-ray absorption to determine the chemical state and atomic local structure. DSAs: dual single atoms; DSAs/NCs: dual single atoms and nanoclusters. Reprinted with permission.126 Copyright 2024, Royal Society of Chemistry. (c) Experimental and fitted FT-EXAFS curves at the Ni Kedge. (d) Experimental and fitted FT-EXAFS curves at the Fe K-edge. Insets are the corresponding atom configurations. Reprinted with permission. (e) Cycling performance of the ZABs made from three types of composite catalysts (5 mA cm−2; 4 h duration per cycle). Reprinted with permission.127 Copyright 2023, American Chemical Society. |
Dual-atom catalysts (DACs), which contain multiple atomic combinations arranged at local sites, address some of the limitations of SACs. As an extension of SACs, DACs retain the advantages of SACs while also exhibiting the geometric ensemble effects and electronic ligand effects typically observed in nanoparticle catalysts. DACs offer diverse and unique local electronic structures at surface sites. This can enhance the adsorption and activation of various intermediates, enabling multifunctional and synergistic catalysis while maximizing atomic utilization. Consequently, DACs represent a more promising strategy for the preparation of bifunctional oxygen electrode catalysts.
Xu's group employed a ZIF phase strategy transformation combined with thermal fixation to prepare nitrogen-doped carbon nanorods containing embedded bimetallic CoFe single atoms and clusters (CoFe–N–C).128 The unique one-dimensional structure of these nanorods and the synergistic of the FeCo system led to enhanced performance. Meng et al. developed a bimetallic NiFe–N–C containing Fe nanoclusters via the simple pyrolysis of metal phthalocyanine and a nitrogen-doped carbon precursor.126 As shown in Fig. 8(b), the charge transfer between the Fe nanoclusters in Fe4@Ni/Fe–N4 and the active sites was significantly stronger than that in Fe4@Fe–N4 and Ni4@Ni–N4. The synergistic between Fe and Ni–N4 and Fe–N4 sites optimized the electronic structure, leading to excellent catalytic activity. Wang et al. proposed a one-step, solvent-free annealing method to construct active centers consisting of dual atoms and clusters in scalable activated carbon by leveraging the strong anchoring effect of phenanthroline on metal ions.129
Li et al. developed an efficient isolation-confinement strategy to synthesize porous carbon nanofibers (β-FeCo-PCNF) loaded with Fe–Co atomic pairs and FeCo alloy clusters.130 Fan's team also achieved the synthesis of bimetallic single atoms and atomic clusters.127 They prepared Ni, Fe-DSAs/NCs by coordinating Ni atoms are coordinated with four N atoms to form a Ni–N4 structure while neighboring Ni atoms were coordinated with three additional Ni atoms to form Ni4 clusters (Fig. 8(c)). A similar configuration was observed for Fe (Fig. 8(d)), which formed both Fe–N4 and Fe4 clusters. In contrast, other samples exhibited only single Ni–N4 or Fe–N4 structures. EXAFS and FT-EXAFS analyses ruled out the possibility of NiFe alloy clusters, indicating that the Fe and Ni in Ni, Fe-DSAs/NCs were able to coexist as stable M–N4 and M4 nanoclusters. In long-term operation tests, a composite catalyst was used to fabricate ZABs. The performance of these ZABs remained stable, with minimal degradation observed after 500 h of cycling (Fig. 8(e)).
Heteroatom-doped and defect-rich carbon-based catalysts exhibit high ORR activity, particularly single- or dual-atom catalysts. However, these materials generally show lower OER activity. Consequently, combining these carbon materials with transition metal oxides can effectively achieve bifunctional catalysis. Zhao et al. reported a nitrogen-doped bimetallic oxide-coated carbon nanotube material (FeNiO@NCNT) derived from MOFs using a straightforward pyrolysis strategy.131 FeNiO@NCNT demonstrated excellent activity. Notably, the lattice oxygen and OH− in FeNiO provide additional OER active sites beyond the metal centers, resulting in outstanding OER performance (Ej10 = 205 mV). The ultimate goal is to be able to combine transition metal oxides with carbon materials by selecting appropriate elements and employing simple preparation methods. Wang et al. developed a nickel–iron oxide material supported on foam carbon with bamboo-like carbon nanotubes.132 During high-temperature synthesis, the edges of the carbon sheets in the foam carbon structure folded inward, forming short carbon nanotubes. Raman spectroscopy revealed that the addition of bimetallic elements enhanced the graphitization (Fig. 9(a)), providing a good conductivity. The supported NiFe2O4 phase improved the ORR activity of this electrocatalyst by anchoring additional N atoms, while the high-valent nickel oxide facilitated O–O bond formation, enhancing OER performance. Shao's team employed an in situ growth strategy to prepare a nitrogen-doped graphene (NG)-supported Fe–Co bimetallic oxide (FeCoOx) with a 3D-on-2D stacked structure (Fig. 9(b) and (c)) and investigated the effect of catalytic performance.133 The optimal FeCoOx@NG composite prepared with a 20% FeCoOx loading demonstrated exceptional ORR and OER electrocatalytic performance. Yoo's group successfully synthesized a NiCoMoO4 (NCMO) catalyst with a hierarchical nanosheet structure on a rGO substrate.134 The selective substitution of Mo at specific sites was used to modulate the electrochemical environment of the active centers (Fig. 9(d)). The introduction of Mo significantly enhanced the redox activity of NiCoMoO4 (Fig. 9(e)). Further analysis revealed that the hierarchical nanostructured NiCoMoO4 supported on reduced graphene oxide demonstrated excellent bifunctional performance in alkaline solutions, with a low overpotential gap of only 0.75 V for ORR and OER. A ZABs fabricated with this catalyst exhibited good cycling stability and a higher power density of up to 125.1 mW cm−2. However, despite the inherent advantages of oxide/hydroxide and carbon-based materials, their bifunctional performance and active site utilization are limited. Rational structural design to effectively integrate oxides/hydroxides with carbon materials is crucial for maximizing their complementary properties and achieving synergistic effects.
Fig. 9 (a) Raman spectroscopy of Fe/NCF, Ni/NCF and FeNi/TNCF-2 samples. Reprinted with permission.132 Copyright 2022, Wiley-VCH. (b) SEM images of FeCoOx@NG. (c) Colorized TEM image of FeCoOx@NG. Reprinted with permission.133 Copyright 2022, The Author(s). (d) Typical experimental setup for an in situ XAS analysis (NiCoMoO4). Reprinted with permission. (e) The setup for in situ XRD analysis and the occurrence of the X-ray diffraction peak took place in the Zn–air battery. Reprinted with permission.134 Copyright 2023, Wiley-VCH. Reprinted with permission. Data-driven analysis of bifunctional oxygen electrocatalysts (f) performance distribution of the reported bifunctional electrocatalysts within the category of M–N–C electrocatalyst (red label), transition metal hydroxides (blue label), and other types of electrocatalysts (gray label). (g) Performance distribution of the specific active sites in “M–N–C” and “Hydroxides” subcategory. The dash lines refer to the requested overpotential to achieve DE < 0.63 V for OER and ORR, respectively. (h) The diagram for bifunctional electrocatalytic activity comparison among FeNC@LDH, Pt/C + Ir/C, and other reported bifunctional electrocatalysts tested in 0.10 M KOH (marked as the green points). Reprinted with permission.135 Copyright 2024, Elsevier. |
Under hydrothermal conditions, a porous, ultrathin nitrogen-doped carbon material (NCM) with enlarged interlayer spacing was prepared via self-assembly. Although this material exhibited excellent activity for ORR, its OER catalytic performance remained limited. To design a bifunctional catalyst, Wang et al. proposed embedding nitrogen-doped nanocarbon materials into Ni–Co layered double hydroxide (NiCo-LDH) nanosheets.136 The integration of these two materials provided NiCo-LDH with enhanced electrical conductivity a larger electroactive surface area, and significantly improved oxygen vacancy formation, which greatly boosted its bifunctional catalytic activity. At 50 mA cm−2, OER overpotential of NiCo-LDH was as low as 0.352 V, and ORR performance was also improved. Therefore, the combination of layered double hydroxides (LDH) with heteroatom-doped carbon materials offers unique advantages for improving both OER and ORR performance.
Achieving record-breaking bifunctional oxygen electrocatalytic activity is a key goal in the development of next-generation ZABs. Historically, the ΔE metric for bifunctional electrocatalytic activity has stagnated at ΔE > 0.60 V. Zhang's team designed and synthesized a composite catalyst with record-breaking bifunctional activity (ΔE = 0.57 V).135 Based on data-driven analysis (Fig. 9(f) and (g)), Zhang's team identified a rational pathway for integrating ORR and OER active sites. They selected Fe–N–C and NiFeCe-LDH as the ORR and OER active centers, respectively, and a composite catalyst (Fe–N–C@LDH) containing both of these active sites was prepared. This catalyst exhibited a groundbreaking ΔE of 0.57 V (Fig. 9(h)). A peak power density of 176 mW cm−2 and a discharge rate of 50 mA cm−2 were achieved in rechargeable ZABs, confirming the exceptional electrocatalytic performance of Fe–N–C@LDH. This catalyst also showed stable operation under harsher conditions. Furthermore, a practical ampere-hour-level ZAB with a unit capacity of 6.4 A h was constructed, demonstrating 20 stable cycles at 1.0 A and 1.0 A h. This work set a new benchmark for electrocatalytic activity and advanced the development of ampere-hour-level ZAB devices for practical applications. Additionally, the study introduced a new material design methodology through data-driven analysis, offering a broadly applicable research paradigm.
In summary, composite bifunctional electrocatalysts combining oxides/hydroxides with carbon materials are extensively studied catalyst materials for OER and ORR. These composites typically demonstrate excellent catalytic performance. Oxides/hydroxides and carbon materials possess high intrinsic activity, providing a natural advantage for bifunctional catalysis. Moreover, the distinct properties of these materials offer diverse opportunities for structural design, enabling the optimization of their interactions and desirable synergistic effects. As a result, composite materials based on oxides/hydroxides and carbon are widely considered as promising alternatives.
Fig. 10 (a) TEM images of the CoSnOH/S NPs. Reprinted with permission.137 Copyright 2024, Wiley-VCH. (b) Bifunctional oxygen electrocatalytic performance comparison of some reported cobalt sulfide-based electrocatalysts. Reprinted with permission.139 Copyright 2021, Royal Society of Chemistry. (c) The process structures of ORR on C–N, Co2N, and Co2N/C–N. Reprinted with permission. Reprinted with permission.140 Copyright 2021, Zhengzhou University. (d) XRD patterns of CoN/UNG, CoN/NG-PEI and CoN/NG-o-phen and UNG. Reprinted with permission. Reprinted with permission.141 Copyright 2022, Elsevier. (e) HRTEM and (f) STEM images of the selected area for EDS elemental mapping images of C, P, and Co in the CoP@C/CNSs. Reprinted with permission.142 Copyright 2023, American Chemical Society. (g) The crystal structures of FeP, Fe2P, Cu3P in Fe–P/Cu3P-NPC. Reprinted with permission.143 Copyright 2023, Wiley-VCH. |
Transition metal nitrides and phosphides often exhibit excellent OER activity and possess metallic properties that provide superior electronic conductivity. By leveraging these characteristics, high-performance bifunctional electrocatalysts can be developed by combining OER-active transition metal nitrides and phosphides with carbon materials that demonstrate ORR activity. Research on nitride–carbon composite catalysts highlights the significant advantages of cobalt-based nitrides. Liu et al. proposed a strategy that utilized the electronic interactions between transition metal nitrides and biomass-derived carbon to design high-performance catalysts.140 They prepared composites consisting of Co2N nanoparticles (NPs) anchored to the ends of nitrogen-doped carbon chains. The ORR reaction pathways for C–N, Co2N, and CoxNBC were structurally similar, with variations in CoxNBC activity attributed to differences in the size and dispersion of Co2N. The electronic states of Co2N were notably enhanced in the nitrogen-doped carbon layers near the Fermi level. The synergistic effect between the Co2N NPs and nitrogen atoms increased the charge density of intermediates. This improved adsorption and charge transfer, leading to significantly enhanced ORR activity. Combined with the controllable preparation method used to prepare these Co2N composites, this innovative catalytic performance offers a promising approach for developing metal-containing catalytic carbon from biomass resources.
Wang's team reported a novel dual construction method that utilized cobalt chloride as a template to prepare a two-dimensional structure containing cobalt single-atom active sites.144 This method enabled the scalable production of CoN3 single atoms dispersed on nitrogen-doped graphene nanosheets, resulting in a bifunctional catalyst (CoSA/NCs). During synthesis, cobalt chloride served as the cobalt source, template, porogen, and graphite catalyst, and this cobalt chloride was able to be recycled. The presence of cobalt chloride endowed the CoSA/NCs with multiple active sites (CoN3 and abundant nitrogen-doped species), large graphite interlayer spacing, a high specific surface area, nanoporosity, and high graphite crystallinity. Controlling the size and the thickness during catalyst preparation is often challenging. Xu et al. synthesized CoN anchored on ultrathin nitrogen-doped graphene using a strategy that combined 1,10-phenanthroline coordination with polyethyleneimine intercalation effects.141 Polyethyleneimine was intercalated between graphene layers, which prevented graphene stacking and resulted in an ultrathin nanosheet structure. The nitrogen lone pairs in 1,10-phenanthroline coordinated with Co2+ to form Co–N bonds, which inhibited Co aggregation. XRD and SEM analyses confirmed the successful anchoring of smaller CoN nanoparticles on the ultrathin nitrogen-doped graphene. CoN/UNG exhibited excellent ORR electrochemical performance, with a E1/2 of 0.87 V and high stability. A zinc–air battery prepared using CoN/UNG demonstrated a power density of 149.3 mW cm−2.
In addition to transition metal nitrides, transition metal phosphides are commonly employed in bifunctional electrocatalysts due to their excellent electronic conductivity and inherent OER activity. Sun et al. developed a carbon dot (CD)-assisted method to fabricate CoP/C nanocomposites with a refined microstructure (Fig. 10(e) and (f)).142 In this process, the CDs acted as a stabilizing platform, adsorbing and immobilizing Co precursors while preventing the agglomeration of CoP nanoparticles. This resulted in a well-structured distribution of carbon-coated CoP particles on carbon nanosheets (CoP@C/CNSs). ZABs prepared using CoP@C/CNSs as the electrocatalyst exhibited excellent performance. Yang et al. utilized ZIF-8 as a sacrificial template and employed a nontoxic phytic acid “in situ etching-adsorption-phosphidation” process to construct a series of carbon-supported metal phosphides with hollow structures and “wall-sharing” features (e.g., Mn2P-NPC, FeP/Fe2P-NPC, CoP-NPC, Ni2P-NPC, Cu3P-NPC, Fe–P/Cu3P-NPC).143 Among these metal phosphides, Fe–P/Cu3P-NPC exhibited the best bifunctional catalytic activity. The electronic interaction between Fe and Cu enhanced the intrinsic catalytic activity of Fe–P/Cu3P-NPC, while the heterogeneous interfaces of FeP, Fe2P, and Cu3P (Fig. 10(g)) increased the number of active sites. Moreover, the ultrathin NPC framework facilitated mass transport. Liquid ZABs prepared using Fe–P/Cu3P-NPC achieved a peak power density of 158.5 mW cm−2. Additionally, excellent cycling stability was achieved in flexible ZABs, which showed stable operation for 81 h at 2 mA cm−2 without bending and for 26 h at various bending angles.
In summary, carbon–sulfide composite bifunctional electrocatalysts can be prepared using a wide variety of material combinations, and these materials demonstrate exceptional electrocatalytic performance. Enhancing the overall bifunctional activity of these electrocatalysts can be achieved by optimizing their intrinsic properties or structure. However, sulfides generally exhibit poor stability under operational conditions, leading to their conversion into oxides, hydroxides, or hydrosulfides. Currently, composite materials that combine transition metal nitrides or phosphides with carbon are receiving increasing attention. Innovations in modulating the activity of these components are emerging as key advancements. Research is evolving from a sole focus on nitrides and phosphides to exploring multi-cationic and multi-anionic compounds, as well as heterogeneous interfaces. Similar to sulfides, nitrides and phosphides also exhibit poor stability in oxidative environments. Understanding the chemical structure evolution of these materials will be crucial.
Catalysts | E1/2 (V vs. RHE) | η@10 mA cm−2 (V) | ΔE (V) | Power density (mW cm−2) | Cycling stability (h) | Ref. |
---|---|---|---|---|---|---|
N-HC@G-900 | 0.85 | 1.58 | 0.73 | — | — | 56 |
mf-pClNC | 0.91 | — | — | 277 | 200 | 58 |
VP/CNs | 0.86 | — | — | 155 | 250 | 65 |
NCN-1000-5 | 0.82 | 1.55 | 0.73 | 207 | 330 | 69 |
Co–N–C/AC | 0.87 | 1.59 | 0.72 | 203 | 72 | 75 |
OLC/Co–N–C | 0.85 | 1.58 | 0.73 | 238 | 100 | 77 |
Ni7Fe3-CNG | 0.89 | 1.50 | 0.61 | — | — | 79 |
FeCo–N3O3@C | 0.93 | 1.52 | 0.59 | 143 | 200 | 86 |
Co3O4-CP-3 | 0.69 | 1.57 | 0.88 | — | 600 | 91 |
V-CMO/rGO | 0.86 | — | — | 145 | 40 | 98 |
Porous NiCo2O4 | 0.79+ | 1.61 | 0.82 | 124 | 160 | 102 |
Co3Mo3N | 0.75 | 1.52 | 0.77 | — | 240 | 104 |
(Co0.17Fe0.83)3N@NPC | 0.88 | 1.56 | 0.68 | 126 | 160 | 106 |
NiCoMnP | 0.72 | 1.53 | 0.81 | 148 | 1000 | 107 |
FeCoNiSx | — | 1.43 | — | 257 | 200 | 112 |
FeNiCo@NC-P | 0.84 | 1.54 | 0.70 | 112 | 120 | 121 |
Fe-ACSA@NC | 0.90 | — | — | 140 | 35 | 125 |
Fe4@Fe–N4 | 0.89 | 1.68 | 0.78 | 217 | 60 | 127 |
FeNC@LDH | 0.88 | 1.45 | 0.57 | 176 | — | 135 |
Fig. 11 (a) Diffusivities simulated by molecular dynamics (MD) at high temperatures and by learning-on-the-fly (LOTF-MD) at intermediate temperatures and corresponding LOTF-MD snapshots. Reprinted with permission.145 Copyright 2020, American Chemical Society. (b) Schematic diagram of geometric health indicators (solid lines denote the selected data; dash lines denote the fitted circle). Reprinted with permission.146 Copyright 2022, Elsevier. (c) Schematic diagram of phase transformation process during pyrolysis. Reprinted with permission.147 Copyright 2024, The Author(s). (d) Compared to furnace annealing, our method shows a time reduction by at least ∼5 orders of magnitude, a size reduction from 27 to 1.9nm, and an increase in electrochemical surface area (ECSA) of 3.8 times. Reprinted with permission. (e) The general synthesis of ultrasmall nanoparticles by coordinated CTS beyond delicate metal–organic framework (MOF) chemistry. Reprinted with permission.148 Copyright 2023, The Author(s). (f) HAADFSTEM images of the Pt = N2 = Fe ABA catalyst. The structure evolution identified by in situ XAFS. (g) The Pt L3-edge and (h) XANES spectra are the corresponding local magnifications of the white-line peaks. (i) 800–1300 cm−1 range of in situ SR-FTIR characterizations for Pt = N2 = Fe ABA. Reprinted with permission.148 Copyright 2022, The Author(s). |
Additionally, in situ nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR) techniques are valuable for detecting molecular-level changes in electronic structures and coordination environments, enabling the further elucidation of complex reaction mechanisms. Integrating these techniques with multi-scale theoretical simulations and data analysis shows promise for providing a more comprehensive understanding of catalytic processes and better guidance for more precise catalyst design. Future advancements should focus on enhancing the spatial, temporal, and energy resolution of in situ techniques to enable the detailed analysis of catalysts under conditions closer to those used during practical operation. Moreover, developing comprehensive analytical platforms that integrate multiple characterization methods will significantly enhance the understanding of bifunctional oxygen electrocatalysts, reduce research biases, and provide critical insights for designing efficient and stable catalysts.
In summary, this review offers a comprehensive analysis of recent advancements in non-precious metal bifunctional oxygen electrode catalysts, with a particular emphasis on the advantages and limitations of carbon materials, transition metal compounds, and their composite structures in catalytic processes. It explores the sources of catalytic activity, including methods for increasing active site density, as well as design principles and regulation strategies based on reaction mechanisms. Through statistical analysis of reported bifunctional electrocatalysts, the review reveals correlations between composition, structure, and performance, offering new insights into how different materials influence electrocatalytic efficiency. Looking to the future, further advancements in this field are expected to pave new pathways for the development of next-generation metal–air batteries and other energy storage and conversion technologies, contributing to a greener and more sustainable future.
This journal is © The Royal Society of Chemistry 2024 |