DOI:
10.1039/D4QM00169A
(Review Article)
Mater. Chem. Front., 2024,
8, 2394-2419
Cobalt-containing ZIF-derived catalysts for Zn–air batteries
Received
4th March 2024
, Accepted 17th April 2024
First published on 18th April 2024
Abstract
Zinc–air batteries (ZABs) are safe, environmentally friendly and stable energy storage devices. However, the two important electrochemical reactions (ORR and OER) in ZABs have high reaction barriers, and thus require the help of a catalyst. At present, the efficient commercial bifunctional catalyst (Pt + RuO2) has some drawbacks such as high price and poor stability. Thus, it is extremely important to find efficient and inexpensive bifunctional electrocatalysts for ZABs. In this case, zeolitic imidazolate frameworks (ZIFs) are emerging functional materials with highly ordered pore structures and extremely strong catalytic activity, which have attracted much attention due to their high specific surface area, tunable pore size, and structural diversity. Especially, cobalt-containing catalysts derived from ZIFs containing cobalt have excellent bifunctional electrocatalytic properties. Cobalt-containing ZIFs with different amounts of metal possess different synergistic effects and active sites. This review summarizes the preparation and modification methods and active sources of ZIF-derived catalysts containing different amounts of cobalt. Furthermore, we summarized the application of ZIF-derived catalysts containing different amounts of cobalt in ORR and OER. Finally, we briefly discussed the development prospects of ZIF-derived catalysts containing different amounts of cobalt in the field of oxygen electrocatalysis and ZABs.

Yansheng Fan
| Yansheng Fan received his Undergraduate Degree from the Applied Technology College of Soochow University in 2021. Currently, he is a Master's student at Hubei University of Automotive Technology. His current research direction is zinc–air battery electrocatalysts. |

Wenhui Wang
| Wenhui Wang is currently an undergraduate student at Hubei University of Automotive Technology. Her current undergraduate major is New Energy Materials and Devices, and her research interest focuses on zinc–air battery materials and their properties. |

Yixin Chen
| Yixin Chen received her Undergraduate degree from Henan Institute of Science and Technology in 2021. She is currently a Master's student at Hubei University of Automotive Technology. Her current research direction is electrocatalysis. |

Zhenyi Xu
| Xu Zhenyi received his Undergraduate degree from Hubei University of Automotive Technology in 2023. He is currently a Master's student at Hubei University of Automotive Technology. His current research direction is optoelectronic energy materials and devices. |

Miao Xu
| Miao Xu received her Master's Degree from Hubei University in 2016, and now works as a Lecturer at the College of Mathematics, Physics and Optic-electronic Engineering, Hubei University of Automotive Technology. Her research interests are focused on micro- and nano-scale energy materials. |

Rui Tong
| Rui Tong obtained his doctorate from the Institute of Applied Physics and Materials Engineering (IAPME) at University of Macau in 2020. He is currently working as an Associate Professor at the College of Mathematics, Physics and Optic-electronic Engineering, Hubei University of Automotive Technology. His research interests are focused on the synthesis of nanomaterials as well as their applications in catalysis, energy storage and conversion. |
1. Introduction
In the 21st century of global modernization, fossil fuels (including non-renewable fuel resources such as coal, oil, and natural gas) are still widely used by countries worldwide.1 However, the continuous use of fossil fuels has resulted in a series of problems such as energy shortage, global warming, environmental pollution, and resource scarcity.2 Thus, presently, to reduce the consumption of non-renewable fossil fuels, the effective utilization of renewable energy sources, especially solar, wind, and hydro energy, is a top priority for society.3,4
In current energy storage technology, various types of batteries have become an important medium for energy acquisition and utilization.5 However, although Li-ion batteries, which are well known to the general public, are widely used in electric vehicles, aerospace, and other applications, alternatives have emerged due to their shortcomings such as high cost, insufficient energy density, poor safety and low recycling rate.6,7 Particularly, ZABs have emerged as vital energy storage candidates because they have many advantages in practical applications. Firstly, ZABs have low-cost electrode materials because their zinc negative and oxygen positive electrodes have richer reserves worldwide.8 Secondly, ZABs have a theoretical energy density of 1086 W h kg−1, which is significantly higher than that of conventional lithium-ion batteries.9 Thirdly, the electrolyte used in ZABs is a safe and environmentally friendly aqueous electrolyte.10 However, ZABs have some drawbacks (such as growth of zinc dendrites and low power) in practical applications.11 The key material affecting the performances of ZABs is the cathode oxygen catalyst.12 Noble metal-based materials (such as Pt/C,13 RuO2,14 and IrO215) are considered superior oxygen electrocatalysts; however, their scarcity and high cost limit their practical large-scale application. Thus, recently, researchers have devoted their efforts to studying non-precious metal electrocatalysts (transition metal oxides,16–18 sulfides,19–21 phosphides,22–24 nitrides,25–27 carbides,28–30 alloys,31–33 and carbon materials34–36).
Metal–organic frameworks (MOFs), a new class of porous materials that emerged at the end of the 20th century, have37 become one of the most promising candidates as functional and novel carbon materials due to their highly controllable structural and surface properties.38 MOFs are organic–inorganic hybrid materials with intramolecular pores formed by the self-assembly of organic ligands and metal ions or clusters through coordination bonds.39 By changing their different coordination modes and adjusting the size of their organic ligands, materials with different functions can be obtained, which have been widely used in many fields such as gas storage, gas separation, catalysis, medicine, sensors, and energy storage devices.40 MOFs have advantages in electrocatalysis, such as high specific surface area, porous structure, higher proportion of metal sites, catalytically active atoms, and easy structural characterization.41 Thus, based on these advantages, MOFs are compatible with electrochemical applications, especially batteries.42
One of the prominent types of MOFs is zeolitic imidazolate frameworks (ZIFs), which are preferentially used in electrochemical applications.43 ZIFs are emerging functional materials with highly ordered pore structures and extremely strong catalytic activity.44 ZIFs are organic–inorganic hybrid systems of zeolites and MOFs, which consist of an inorganic porous skeletal system tightly coupled with imidazolium-type organic linkers, resulting in a highly ordered and spatially regular pore structure.45 ZIF derivatives are often used as ideal catalysts for battery electrodes because they not only retain the original advantages of ZIFs, such as high specific surface area, adjustable pore size, and structural diversity, but also the unique advantages of zeolites, such as adsorption, catalytic, and ion exchange properties and thermal stability.46 In recent years, simple transition metal cobalt materials have shown high ORR and OER activity due to their excellent stability as non-precious metals with heat resistance, abrasion resistance, corrosion resistance, ferromagnetism, flexibility, and good electrochemical properties. Also, cobalt can be more readily bonded to form ZIFs than other metals.47
Due to their excellent bifunctional electrocatalytic performance in OER and ORR, there are many reports on the application of cobalt-ZIFs and their derivatives in the field of electrocatalysis.48 This review summarizes the cobalt-containing ZIF-derived catalysts applied in Zn–air batteries.49 Currently, most reviews only focused on ZIF derivatives with a fixed quantity of metal.50 Therefore, in this review, we systematically summarize the methods for the preparation and modification of cobalt-based ZIF derivatives, cobalt–M-based ZIF derivatives, and cobalt-based trimetallic ZIF derivatives as ORR electrocatalysts.51 Regarding the strategies for the preparation and modification of cobalt-based monometal derivatives, we mainly summarize them based on five aspects including pyrolysis, oxidation, sulfuration, doping and hybridization.52 Among them, hybridization is an important strategy to improve the properties of cobalt-based ZIF derivatives, including compounding with conductive carbon (carbon nanotubes, carbon fibers, etc.) to improve their conductivity and binding with highly active ORR electrocatalysts (Co3O4, NiCo-layered double hydroxides (LDH), etc.) to increase their number of active sites, and hybridizing with organics (supramolecular gels (SMG), benzimidazole (BIm), etc.) to enhance their structural stability.53 Similarly, in this review, we summarize methods for the preparation and modification of cobalt-based trimetallic ZIF derivatives (mainly ZIFs formed by cobalt metal and transition metals such as Fe, Ni, and Zn).54 Finally, the challenges and prospects of developing cobalt-containing ZIF-derived electrocatalysts with bifunctional ORR and OER activity are outlined to provide a comprehensive understanding.55
2. Fundamentals of Zn–air batteries
ZABs mainly include two types, i.e., primary ZABs and rechargeable/secondary ZABs, and mainly comprised of three parts including a zinc plate or zinc powder electrode (anode), air electrode (cathode), and alkaline electrolyte. Primary zinc–air batteries convert chemical energy into electrical energy for external circuits and are powered only by internal chemical reactions.56 Alternatively, secondary ZABs are discharged and charged by atmospheric oxygen through ORR and OER at the air electrodes, respectively.
2.1 Zn electrodes
Zinc is abundant in the Earth's crust, and simultaneously zinc is fully recoverable. Furthermore, it is cheaper and more environmentally friendly than other metal materials, making it an excellent anode material for ZABs. However, zinc reacts violently in acidic electrolyte, leading to corrosion, and thus alkaline electrolyte is generally employed in ZABs. During the discharge process, Zn loses electrons and is converted to Zn2+. The O2 at the cathode receives electrons from the anode to undergo ORR to obtain OH−, subsequently forming Zn(OH)42− with the Zn2+ produced at the anode, and when the Zn(OH)42− deposition exceeds the saturation limit, it is further decomposed into insoluble ZnO and deposited on the surface of the anode.57 The continuous accumulation of ZnO in the insoluble insulating layer on the zinc electrode produces a passivation layer, which blocks the pores and active sites, leading to an increase in internal resistance and a gradual decrease in conductivity, discharge capacity, and power capacity. Ultimately, this blocks the entry of ions, further affecting the utilization and reversibility of zinc and adversely affecting the overall reaction of the cell.58 Thus, to address the problem of passivation, many reports employed different approaches, such as increasing the concentration of zinc oxide, causing it to be greater than a critical value, accelerating the dissolution of the metal to avoid passivation;59 and adding surfactants to the electrolyte to reduce the surface tension and the surface free energy, eliminating the interface between the two phases and loosening the passivation layer.60 In addition, during the charging process, ZnO gains electrons and undergoes a reduction reaction to be converted into Zn, but the deposition process of Zn2+ into Zn is not uniform due to the uneven flow of current in the cell during this process, with further problems such as the growth of dendrites (dendrites: the part of the needle protrusion that the deposits gather together and the deposits gather together to form a small, hard tree structure). Clade crystals and dendrites can penetrate the diaphragm inside the battery and cause a short circuit. There are many reports on alleviating dendrites, such as by optimizing the electric field distribution on the electrode surface, resulting in the uniform deposition of zinc.59 Also, the addition of additives to the electrolyte reduces the rate of dendrite formation.60 In addition, in the electrolyte on the anode surface, the OH− generated in the electrolyte through ionization of water will compete with ZnO for electrons, and then H2 is generated through a reduction reaction, and this process is known as the hydrogen evolution reaction (HER), i.e., Zn + H2O → ZnO + H2.61 Given that it competes with the anode for electrons, it can impact the charging performance of the battery, greatly reducing the charging efficiency. At the same time, it can also cause corrosion of the zinc anode, and hydrogen generation increases the risk of the expansion or even explosion in ZABs. In this case, the concentration of hydrogen ions is adjusted, namely changing the electrode potential of hydrogen ions;59 reducing their ability to grab electrons, namely appropriately increasing the pH;60 and adding corrosion inhibitors and other methods to inhibit or slow down the occurrence of HER.62 The above-mentioned four issues (passivation of zinc during discharge, formation of dendrites during charging, and hydrogen evolution during charging) limit the development of ZABs.
2.2 Air electrodes
The air electrode, as the cathode of the secondary ZABs, is accompanied by OER and ORR during the charging and discharging operation of secondary ZABs, respectively, where their reaction rate directly determines the performance of the battery.
The ORR is divided into two-electron and four-electron paths. The two-electron path is represented by the following equation:
| O2 + H2O + 2e− → HO2− + OH− | (1) |
| HO2− + H2O + 2e− → 3OH− | (2) |
According to these equations, the two-electron path involves the conversion of O2 into intermediate HO2− firstly, which later produces OH−. Due to the corrosive nature of peroxides, they can harm the catalyst, leading to a decrease in the performance of the oxygen electrocatalyst and ZABs.63 Therefore, the four-electron path is the ideal ORR path. The equations for the four-electron path are as follows:
| * + O2 + H2O + e− → OOH* + OH− | (3) |
| O* + H2O + e− → OH* + OH− | (5) |
Overall reaction:
where * denotes the catalyst surface and O*, OH*, and OOH* denote the oxygen intermediates produced when oxygen binds to the active site. The adsorption and dissociation of these intermediates are effective for the ORR and OER performances. Given that the reduction potential of the four-electron path (1.23 V) is higher than that of the two-electron path (0.695 V), ORR is more prone to undergo the two-electron path when the catalytic efficiency of the cell is low. In the case of ZABs, the most desirable reaction is still the four-electron path, given that it avoids cell corrosion while improving the cell performance with efficient use of the ORR reaction.
During charging, OER is the opposite reaction ORR and is the key reaction for rechargeable ZABs. The reaction equation for OER is expressed as follows:
| OH* + OH− → O* + H2O + e− | (9) |
| OOH* + OH− → * + O2 + H2O + e− | (11) |
Overall reaction:
| 4OH− → O2 + 2H2O + 4e− | (12) |
According to these equations, OER is a reaction in which OH
− is converted to oxygen and water by transferring four electrons in sequence through four steps. In the standard case, the thermodynamic equilibrium potential of OER is 1.23 V. However, in practice, due to the hindrance of some kinetic processes, the working potential of the actual reaction often needs to be higher than the value of the equilibrium potential, and the voltage exceeding the theoretical value is called overpotential. Therefore, efficient redox electrocatalysts with optimal binding energy are needed to reduce the overpotential.
2.3 Overall reactions
ZABs are chemical power sources with oxygen in the air as the positive active substance and zinc metal as the negative active substance, with the following reaction formula when discharged at the cathode: | O2 + 4e− + 2H2O → 4OH− | (13) |
The reactions at the anode are as follows: | Zn + 4OH− → Zn(OH)42− + 2e− | (14) |
| Zn(OH)42− → ZnO + H2O + 2OH− | (15) |
The overall reaction is:ZABs work on the principle that Zn on the anode is oxidized by oxygen to ZnO and releases electrons, and the oxygen on the cathode gains the electrons and is reduced to OH−. This process produces a potential difference between the cathode and anode, generating a current to realize charging and discharging.
2.4 OER/ORR evaluation parameters
In OER and ORR tests, a three-electrode system is employed, which is comprised of a working electrode, reference electrode, and counter electrode.64 Firstly, we introduce the working electrode, which is usually a glass carbon electrode because of its good electrical conductivity, easy cleaning, strong durability, and good stability. The second is the reference electrode, which is usually Ag/AgCl, calomel, and mercury oxide electrodes under acidic, neutral, and alkaline conditions, respectively. Finally, as the counter electrode, a graphite rod is usually preferred, and Pt filament or Pt sheet can be selected. However, because Pt will dissolve in the electrolytic cell for a long time, the dissolved Pt ions will be deposited on the working electrode, which will have a certain impact on the test results. Therefore, graphite rods are usually used as the counter electrode.
To reasonably evaluate the activity of electrocatalysts and obtain information about the reaction mechanism, it is necessary to determine the evaluation indexes of the electrocatalytic kinetics. The main kinetic indicators currently in use include are presented below.
2.4.1 OER overpotential.
The overpotential can directly reflect the catalytic activity of OER catalysts, which is the difference between the actual potential and the theoretical potential when a certain current density is reached during the catalytic reaction.65 The overpotential can be calculated using the following equation: | Ej = Ei (actual potential) − Et (theoretical potential) | (17) |
According to this equation, it can be seen that the overpotential cannot be zero, but theoretically, the smaller the overpotential, the better the performance of the catalyst, the lower the actual voltage required to reach the related current density, the relatively smaller the energy consumption, and the higher the catalytic activity. However, it has to be compared at the same current density or the same potential, otherwise the comparison will be meaningless. The performance of catalysts is generally compared using a current density of 10 mA cm−2 as the benchmark, and the corresponding overpotential is noted as Ej=10.
2.4.2 Onset potential.
The onset potential is a key parameter for evaluating the catalytic activity of ORR. The starting potential is the equilibrium potential reached by an electrode in contact with a solution in the absence of any external interference. The potential at a current density of 0.1 mA cm−2 or the potential corresponding to 5% of the limiting current is usually determined as the starting potential.66 The onset potential is affected by various factors, such as pH, temperature, and ion concentration. Among them, one of the most important factors is pH. When the pH changes, the nature of the oxide or reductant film on the metal surface also changes, resulting in a shift in the onset potential.
2.4.3 Half-wave potential.
The half-wave potential is another key parameter reflecting the catalytic activity of ORR, which is expressed as E1/2.67E1/2 refers to the value of the potential corresponding to more than half of the limiting current in ORR. The factors that affect the half-wave potential are the type of electrolyte and its concentration, temperature, etc. When the electrolyte concentration and testing temperature remain constant, the value of the half-wave potential is also unchanged.
2.4.4 Limiting current density.
The limiting current density is the maximum current density of the electrode reaction,68 which is mainly affected by the pH value, gas concentration, and temperature, and all are proportional to it. Additionally, is an essential parameter to characterize the reaction speed of the electrode.
2.4.5 Potential gap.
The potential gap is a key parameter reflecting the catalytic activity of a catalyst and is expressed as ΔE.69 It is the potential difference between the OER potential at a current density of 10 mA cm−2 and the half-wave potential of ORR. The smaller the potential gap, the better the bifunctional catalytic performance.
2.4.6 Tafel equation and Tafel slope.
The Tafel equation is an essential equation in electrochemical kinetics that relates the electrochemical current density to the overpotential. Tafel analysis is a powerful tool for evaluating or discussing the steps involved in the determination of the electrocatalytic rate. Tafel analysis enables the analysis of the Tafel slope, from which the relationship between current density and overpotential can be known further to understand the rate and mechanism of the reaction. The relationship between the Tafel slope and overpotential can be expressed mathematically by the Tafel equation as follows: | η = a + b log j | (18) |
where η and j are absolute values, η is the overpotential, and j is the current density. Also, a represents the overpotential value at a unit value of current density (1A cm−2) and b is called the Tafel slope.70 Depending on the magnitude of the a-value, it is possible to compare the ease of performing the electron transfer step in different electrode systems. The other essential parameters that can be obtained based on the magnitude of b include the electron mobility number, charge mobility coefficient, and reaction rate. When the Tafel slope (b) is positive, it indicates that the electrochemical reaction controls the electrode reaction rate. When the Tafel slope (b) is negative, it indicates that the electrochemical reaction controls the diffusion rate of the electrolyte. The smaller the Tafel slope, the better the kinetic performance of the response, the smaller the overpotential in the catalytic process, and the more desirable the electrocatalytic activity.
2.4.7 Real surface area.
The real surface area is the surface area with catalytic sites. There are two methods to calculate the catalyst surface area, namely the BET test and calculation of the electrochemical active surface area (ECSA).71 The BET test is a gas adsorption method that uses the adsorption of gas on the material surface to evaluate its specific surface area and pore structure. However, not all the electrochemical active sites measured by this method are catalytically active. In practical applications, BET testing is usually used with other characterization methods to allow a more comprehensive evaluation of the material properties. In addition, the ECSA refers to the electrochemically active area, which is the effective activated surface area involved in the electrochemical reaction. The electrochemical active area of the catalyst is calculated based on the following equation:where Cdl is the bilayer capacitance measured electrochemically and Cs is the capacitance in the ideal plane of the catalyst. The measurement of the electrochemical surface area can give information and signals, namely, the changing trend of the catalyst activity, the reaction rate of the electrode surface, and the capacity of the electrochemical reaction, which are the main parameters affecting the electrochemical reaction.
2.4.8 Electron transfer number and HO2− percentage.
The electron transfer number of ORR can be expressed by the Koutecky–Levich equation (K–L equation for short) as follows: | 1/j = 1/jL + 1/jK = 1/Bω1/2 + 1/jK | (20) |
where j is the experimentally measured electrode current; jL and jK are the limiting and kinetic current densities, respectively; B is regarded as a constant and written as 0.620 nFC0D02/3ν−1/6; ω is the disk electrode rotation rate; n is the electron transfer number in ORR; F is the Faraday constant (96
485 C mol−1); C0 is the concentration of O2 in solution; D0 is the diffusivity of oxygen molecules; and ν is the kinetic viscosity of the solution.72 Therefore, based on the experimental data, at the same voltage, different rotational speeds and their corresponding current densities are plotted, yielding the scatterplot of 1/j and 1/ω1/2. These points are all on a straight line with a slope of 1/K, and the electron transfer number, n, can be determined. The ORR test set is generally RRED. The yield of HO2− can be determined using the following equation: | %(HO2−) = 200IR/N/(ID + IR/N) | (21) |
where ID is the disk current, IR is the ring current, and N is the collection efficiency of RRED.
2.4.9 Turnover frequency.
The turnover frequency (TOF) is used to characterize the intrinsic activity of the active sites of different electrocatalysts,73 which is the number of molecules of H2 or O2 generated per unit of time at one catalyst active site, with units of 1/h. In practice, it is impossible to calculate the number of active sites involved in the reaction because the active sites cannot come into contact with the reactants when the catalyst is covered. Thus, practical methods for accurately determining the total active sites on the surface of electrocatalysts still need to be discovered. Most research calculations assume that all the active sites can contact the reactants and the TOF calculated accordingly. The calculated values will be smaller than the actual values, but they can still be compared.
2.5 ZAB evaluation parameters
Many studies have applied catalysts to ZABs to test their performance in practical applications. Here, we present a few of the key performance parameters of ZABs.
2.5.1 Open circuit voltage.
The open circuit voltage (OCV) is a characteristic parameter of battery systems. The terminal voltage of the battery is in the open circuit state, namely the difference between its positive and negative electrode potentials when disconnected. The OCV is calculated using the following equation:where Ec and Ep are the positive and negative electrode potentials of the battery, respectively. The main factors affecting the OCV are the nature of the system constituting the poles of the battery, the material nature, the electrolyte composition and concentration, and the temperature.
2.5.2 Charge–discharge performance.
The charging and discharging performance of a battery is a capacity limit based on comprehensive physical and chemical considerations of safety performance. This depends on the type of material, composition ratio, structural design, the battery itself, etc. The charge–discharge curve represents the relationship between the charge/discharge current and voltage of the battery. Representing the battery material charging and discharging behavior, the analysis of the battery charging and discharging curves is of great significance in understanding the properties and electrochemical behavior of materials, especially for the analysis of half-cell charging and discharging curves, which can be targeted to analyze the behavior of the characteristics of a specific type of material.
2.5.3 Peak power density.
The power of the battery is obtained from the product of voltage and current. In the I–V curve of the battery, the voltage value at each point is multiplied by the corresponding current density value to obtain the power density curve. The maximum value in the peak power density curve is the fitted power density, where a higher peak power density indicates more efficient battery operation and less energy consumption.
2.5.4 Specific capacity.
The specific capacity is one of the leading performance indicators of a battery. There are two types of specific capacity, i.e., weight-specific capacity and volume-specific capacity. Zinc–air batteries are measured using weight-specific capacity, which is the amount of electricity that can be discharged per unit weight of battery or active substance. The formula for calculating the weight-specific capacity of a battery is as follows: | Specific capacity = current density × service hours/weight of consumed zinc | (23) |
Generally, the higher the specific capacity, and the more stable the voltage and the better the discharge performance of the battery. Conversely, the lower the particular capacity and the greater the voltage fluctuation, the worse the discharge performance of the battery. The specific capacity of a battery is usually affected by its structure, electrolyte, electrodes, temperature, and other factors. Thus, to improve the specific capacity of the battery, these factors need to be fully considered, and the battery structure, electrolyte, and electrode materials need to be rationally selected to achieve a higher specific capacity.
2.5.5 Charge–discharge cycling performance.
The charge–discharge cycling performance testing of batteries is an essential test method. Charge–discharge cycling simulates the actual use of the battery by repeatedly charging and discharging it to test its performance, life, and stability to optimize its performance.
The cycling performance of a battery is measured by three leading indicators, i.e., number of cycles, first discharge capacity, and retained capacity. The cycle number is the number of times the battery has cycled through multiple charge–discharge cycles; the discharge capacity is the discharge capacity of the battery during the first charge–discharge test; and the retention capacity is the discharge capacity maintained by the battery after it has completed a certain number of cycles of charging and discharging. With the same number of cycles, the greater the ratio of retained capacity to first discharge capacity, the better the cycling performance of the battery.
3. Recent research progress in cobalt-containing ZIF-derived catalysts
3.1 Co-Based ZIF derivatives
Cobalt-ZIFs can be directly used as ORR and OER electrocatalysts. However, their catalytic performance is relatively poor, which is possibly due to their organic linkers blocking their active site and the poor interactions between their metal center and the ligand. Thus, to improve the catalytic performance of cobalt-ZIFs as ORR and OER electrocatalysts, many reports have studied various strategies, including pyrolysis, oxidation, sulfuration, doping, and hybridization.
(1) Pyrolysis.
Cobalt-based ZIF derivatives usually require pyrolysis to fully expose their active sites. Liu et al. designed highly dispersed cobalt quantum dot-embedded (diameter of 4.57 nm) ultra-thin nitrogen-doped carbon layer (2 nm), denoted as Co@NCL.74 It showed excellent catalytic activity against ORR and OER due to the unique structural features of the obtained cobalt with NCL catalysts, including highly dispersed cobalt quantum dots, a carbon substrate with high nitrogen doping content, and a large specific surface area. Co@NCL had an ORR half-wave potential of 0.84 VRHE (V vs. RHE) and OER overpotential of 1.63 VRHE at 10 mA cm−2. Guan et al. reported the synthesis of unique cobalt/cobalt nitride nanoparticles (NPs) with a two-dimensional nanoarray-based catalyst (NC–Co/CoNx).75 The ORR half-wave potential and OER overpotential of NC–Co/CoNx were determined to be 0.87 VRHE and 289 mVRHE, respectively. Furthermore, the covered NC–Co/CoNx nanoarrays were used as the cathode, carbon fiber as the anode, and gel as the electrolyte to assemble coaxial fiber solid-state ZABs. These coaxial fiber solid-state ZABs exhibited a significantly enhanced volume power density and good flexibility, with broad application prospects in flexible energy storage devices. The solid-state ZABs with the NC–Co/CoNx cathode had a peak power density of 41.5 mW cm−3, and the cells containing the NC–Co/CoNx cathode operated longer than Pt/C in both flat and curved states, showing better structural deformation tolerance.
Many reports suggest that cobalt-ZIFs may produce carbon nanotubes (CNTs) during pyrolysis. Chen et al. investigated the effects of the structure of cobalt-ZIFs and the subsequent carbonization atmosphere on the production of CNTs and catalyst performance.76 The surface of the two-dimensional nanosheet ZIF-L was covered with large amounts of CNTs, and the tip of CNTs was coated with highly dispersed cobalt nanoparticles. The abundant Co–N–C active sites and large specific surface area endowed Co–NCS@CNTs with a half-wave potential of 0.86 VRHE. Furthermore, the ZABs assembled by the catalyst had a peak power density of 90 mW cm−2 and a specific capacity of 798 mA h g−1.
Traditional heat treatment methods are used to convert cobalt-ZIFs into carbon materials, which require high temperatures, long preparation time, and a unique gas environment. Zou et al. used a laser-induced carbonization strategy (LIC) to rapidly convert cobalt-ZIFs into a conductive carbon grid structure in milliseconds (Fig. 1a–c).77 More interestingly, LIC-ZIF-67-M10 exhibited a better bifunctional electrocatalytic performance than the carbon materials obtained by the conventional carbonization process. Specifically, LIC-ZIF-67-M10 had an ORR half-wave potential of 0.77 VRHE and OER overpotential of 390 mVRHE. The ZAB assembled by this catalyst exhibited a high specific capacity of 780 mA h g−1 and a long cycle stability of 220 h (Fig. 1d). These excellent properties most likely originated from their high specific surface area and unique honeycomb open channel structure, which facilitated electrolyte storage and mass transfer processes.
 |
| Fig. 1 (a) Illustration of the fabrication process of LIC-ZIF-67-M10. (b) and (c) SEM images of LIC-ZIF-67-M10. (d) Galvanostat charge–discharge of LIC-ZIF-67-M10 and Pt/C + RuO2. Copyright 2022, Elsevier. (e) Schematic demonstration of the process for the fabrication of NP-Co3O4/CC. (f) SEM images of NP-Co3O4/CC. (g) ORR and OER potential gaps of various samples. (h) Aqueous zinc–air battery voltage and power density with NP-Co3O4/CC and Pt/C + Ir/C as the air electrode. (i) Galvanostatic discharge–charge cycling curves at 5 mA cm−2 for aqueous rechargeable Zn–air batteries with NP-Co3O4/CC and Pt/C + Ir/C catalysts as the air electrode, respectively. Copyright 2020, Elsevier. (j) Schematic representation of the formation of Co/Co9S8-NCL. (k) SEM images of Co/Co9S8-NCL. (l) LSV curves of Co/Co9S8-NCL, Co–N–C, and Pt/C. Copyright 2022, Elsevier. (m) Synthesis route of Mn/Co–N–C-0.02–800. (n) Photograph of an LED lit by a series connection of three all-solid-state ZABs. (o) Galvanostatic discharge–charge cycling profiles of Mn/Co–N–C-0.02–800 all-solid-state ZAB at 2 mA cm−2 with a per cycle period of 10 min. Copyright 2019, the American Chemical Society. | |
The use of cobalt ZIFs as precursors for pyrolysis may produce single atoms of cobalt. Sun et al. reported the use of cobalt-ZIFs to form nitrogen-doped carbon nano/microtubes confined to metallic cobalt nanoparticles and single atoms (Co@NCNT/Co-SA@NCMT).78 The NCMT was only 1.6 μm, while ordinary carbon nanotubes were 200 nm. These microtubules facilitated the better diffusion of oxygen molecules to the active site, improved the local concentration, and promoted the catalytic performance at a high current density. In addition to microtubule production, cobalt single atoms were found in the final product of carbonization. Therefore, the final catalyst named Co@NCNT/Co-SA@NCMT exhibited an excellent ORR catalytic performance (E1/2 = 0.87 VRHE) and OER catalytic performance (Ej=10 = 1.6 VRHE). The ZABs assembled by this catalyst had a peak power density of 155 mW cm−2 and a specific capacity of 846 mA h g−1. Furthermore, they showed a stable voltage gap during charging and discharging for 300 h at a current density of 5 mA cm−2.
(2) Oxidation.
In recent years, an increasing number of reports has studied the role of cobalt oxides in ORR and OER. Zhang et al. designed a high-performance ORR/OER electrocatalyst with a pore structure combining Co–Nx and cobalt oxide using an in situ sodium borohydride reduction strategy (Co3O4–Co@NC-2).79 The in situ sodium borohydride reduction strategy employed in this study solved the problem of metal agglomeration in ZIF-67 during pyrolysis at high temperatures. In addition, the XPS and BET characterization results showed that Co3O4–Co@NC-2 had the highest proportion of pyridine-N and the largest specific surface area. This finding suggests that the in situ sodium borohydride reduction strategy can effectively increase the active site and specific surface area of ZIF-67-derived catalysts. Thus, Co3O4–Co@NC-2 had a half-wave potential of 0.86 VRHE, the potential difference between the OER and ORR was 0.72 VRHE, and the liquid ZABs based on Co3O4–Co@NC-2 had a peak power density of 158 mW cm−2, a high specific capacity of 758 mA h g−1. In addition, constant current charging and discharge were performed for 200 h at a current density of 10 mA cm−2. Wang et al. reported the synthesis of ultrafine nitrogen-doped cobalt oxide nanoparticles on carbon cloth (NP-Co3O4/CC) via a mild oxidation method (Fig. 1e and f).80 Owing to its abundant accessible surface active sites and pores as well as optimized oxygen adsorption, NP-Co3O4/CC delivered a particularly low potential gap of 0.66 VRHE (Fig. 1g). As a difunctional oxygen electrode for rechargeable liquid zinc–air batteries and all-solid-state flexible ZABs, the ultra-high power density of the NP-Co3O4/CC-based batteries reached up to 200 mW cm−2 and 99.8 mW cm−3, respectively (Fig. 1h). In addition, the liquid ZABs had a long cycle life of 400 h, showing excellent battery stability (Fig. 1i).
(3) Sulfuration.
In addition to using cobalt-ZIFs as a precursor, many reports convert cobalt-ZIFs into sulfide as efficient electrocatalysts. For instance, Huang et al. reported the simple sulfation of dithioxamide (DTO) with two-dimensional leaf-shaped ZIF-L, preparing Co/Co9S8 nanoparticle-decorated nitrogen and sulfur double-doped carbon materials (Co/Co9S8-NCL) with excellent ORR performance (Fig. 1j and k).81 The core–shell structure of this catalyst could shorten the three-phase interface reaction pathway, expose more active sites, and solve the problem of possible structure collapse of ZIF-L during carbonization, thus improving the catalytic performance of ORR. The Co/Co9S8-NCL had an ORR half-wave potential of 0.852 VRHE, and the ZAB assembled by this catalyst has a high power density of 112 mW cm−2 and high specific capacity of 799 mA h g−1 (Fig. 1l).
(4) Doping.
Pyrolysis of cobalt-ZIFs usually causes problems such as the uneven decomposition and severe accumulation of metals, resulting in insufficient electrocatalytic activity of the final compound. Many reports have addressed this issue by tuning the composition and microstructure of cobalt-ZIF precursors. For example, doping cobalt-ZIFs with other metal ions can provide a practical approach for synthesizing high-performance electrocatalysts. Pendashteh et al. reported the preparation of an efficient bifunctional ZIF-9_Fe3_Pyrol electrocatalyst by adjusting the iron doping and calcination temperature with iron-doped cobalt-ZIF as the precursor.82 In this catalyst, high-speed doping could be achieved without changing its crystal structure, and iron doping could significantly affect its electrocatalytic activity by increasing the number of transferred electrons. ZIF-9_Fe3_Pyrol exhibited an ORR start potential of 1.55V at 10 mA cm−2 and ORR half-wave potential of 0.81 VRHE. In addition, the ZAB assembled by the catalyst showed a specific capacity of 815 mA h g−1 and good stability. Similarly, Wei et al. reported a simple dipping solution method to prepare Mn/Co–NC-0.02–800 with a stable dodecahedral morphology using manganese-doped cobalt-ZIF as the pyrolysis precursor (Fig. 1m).83 Due to the enhanced conductivity and activity of the catalyst after Mn doping, it exhibited good electrocatalytic activity for ORR/OER, with an ORR half-wave potential of 0.80 VRHE and OER overpotential of 1.66 VRHE. The battery assembled with this catalyst was cycled for 250 h at the current density of 5 mA cm−2. The peak power density was 136 mW cm−2, which also could power a light-emitting diode (LED) viewing screen with three batteries in series (Fig. 1n). After charging and discharging for 60 cycles at the current density of 2 mA cm−2, the increase in the charge and discharge potential difference was negligible (Fig. 1o).
(5) Hybridization.
In addition to separate pyrolysis, oxidation, sulfation, and doping, many reports use hybridization to design catalysts. Cobalt-based ZIF catalysts with multiple material advantages can be obtained through the hybridization of different substances with cobalt ZIF derivatives.
Many reports have hybridized cobalt ZIFs with carbon materials and carbon sources (CNTs,84 supramolecular gels (SMG),85 benzimidazole (BIm),86 α-amylase,87 polyaniline (PANI),88 Pluronic F127,89 glucose,90etc.) to improve their conductivity, structural stability, and adsorption capacity for oxygen and water. The carbonization of cobalt-ZIF materials often results in poor electrical conductivity, while carbon nanotubes are carbon materials with good electrical conductivity. Thus, many reports have structured ZIF materials with carbon nanotubes to improve their electrical conductivity and electrocatalytic properties. The traditional method usually involves simply mixing cobalt-ZIFs with CNTs, resulting in their uneven and insufficient combination. Lv et al. used a simple solvent thermal method to synthesize cobalt-ZIFs and CNT suspensions into flexible Co–N–C/CNT films (Fig. 2a and b).84 This method of film preparation avoided the use of adhesives and prevented the stripping of the catalyst particles. The ORR half-wave potential of the catalyst was 0.87 VRHE (Fig. 2c). The flexible ZAB assembled with this flexible Co–N–C/CNT film showed a stable platform discharge at 1.15V and charge at 1.97 V under a current density of 1 mA cm−2, respectively.
 |
| Fig. 2 (a) Schematic showing the steps for the preparation of Co–N–C/CNT films. (b) Photograph of Co–N–C/CNT film. (c) ORR LSV curves of Co–N–C/CNT films. Copyright 2019, Elsevier. (d) Schematic illustration of the formation mechanism for the decussation-shaped D-Co@NC nanostructure. (e) SEM image of D-Co@NC. (f) ORR LSV curves of D-Co@NC. Copyright 2022, Elsevier. (g) Schematic diagram of Co–O–ZIF/PANI synthesis path. (h) SEM image of Co–O–ZIF/PANI. (i) Charging and discharging curves. (j) Photograph of a red LED powered by two liquid ZABs in series. Copyright 2022, Elsevier. (k) Scheme illustration for synthesizing CoS2@MoS2@NiS2 double-shelled polyhedron. TEM image (l) and HRTEM image (m) of CoS2@MoS2@NiS2. (n) OER LSV curves. Copyright 2022 Elsevier. | |
Zhang et al. designed a simple enzyme-assisted synthesis strategy to convert Co-ZIF precursors into structurally stable crossover shapes, obtaining Co nanoislands grafted on decussate N-doped carbon nanoleaves (D-Co@NC) (Fig. 2d and e).87 The structural mediator of α-amylase plays the key role in forming reciprocal crossover frames by coordinating competition with organic ligands and regulating the nucleation and growth process of the ZIF precursors. Due to its powerful interpenetration structure and unique composition, D-Co@NC showed stronger catalytic activity in ORR and OER compared to catalysts derived from ZIFs and conventional three-dimensional ZIFs, and its ORR half-wave potential was E1/2 = 0.852 VRHE (Fig. 2f). In the ZAB, the peak power density was 115.4 mW cm−2. The energy density was 879.6 W h kg−1 and the long cycle stability exceeded 200 h, which was better than that of the benchmark catalyst of Pt/C + RuO2. Lei et al. reported the combination of PANI with a cobalt-ZIF as a bifunctional electrocatalyst (Co–O–ZIF/PANI) (Fig. 2g and h).88 This Co–O–ZIF/PANI exhibited excellent surface adsorption capacity and a suitable Co3+/Co2+ ratio, which facilitated the progress of the electrocatalytic reaction. In situ Raman spectroscopy showed that the surface chemisorption properties of O2 and H2O brought about by the Co–O bond and PANI also enhanced the catalytic properties. The half-wave potential of Co–O–ZIF/PANI was 0.7 VRHE. The voltage gap of ZABs based on Co–O–ZIF/PANI had no significant change after 300 cycles (100 h) at a current density of 5 mA cm−2 (Fig. 2i). As a demonstration of practical applications, two ZABs connected in series could illuminate a red LED with a stable working status, confirming the favorable application prospects of Co–O–ZIF/PANI-based ZABs (Fig. 2j).
Many reports have also hybridized cobalt ZIFs with metal alloys (NiFe alloy,91 CoFe alloy,92,93 CuCo alloy,94etc.). Metal alloys are one of the recent research hotspots of catalytic materials. Compared with single metal–NC, metal alloys have better bifunctional electrochemical properties due to the synergistic interaction of different metal sites. Therefore, many reports used doped metal alloys to obtain catalysts with polymetallic–NC active sites. For instance, Zhu et al. reported the successful synthesis a bifunctional electrocatalyst (NiFe–Co@NC-450) composed of NiFe alloy by simple solution immersion and pyrolysis using ZIF-67 as the template.91 This study explored the activity of the catalyst by changing the metal composition and content, and concluded that the samples with polymetallic and higher metal content had better electrocatalytic properties. NiFe–Co@NC-450 exhibited an ORR half-wave potential of 0.833 VRHE, and the potential difference between OER and ORR was 0.857 VRHE. The liquid ZABs had a high specific capacity of 798 mA h g−1 and a high power density of 100 mW cm−2 over 200 h at a current density of 10 mA cm−2. Du et al. developed a nitrogen-doped carbon material (CoFe/NC) of composed CoFe alloy using a simple pyrolysis method.92 CoFe/NC had a high surface area, nitrogen content, and well-dispersed CoFe alloy nanoparticles, also exhibiting an ORR half-wave potential of 0.82 VRHE. The liquid ZAB had a peak power density of 173 mW cm−2 and a stable cycle of 230 h (about 690 cycles) at a current density of 10 mA cm−2.
In addition, there have been many reports on the hybridization of cobalt sulfide derived from cobalt ZIFs and other metal sulfides. Liu et al. reported the preparation of CoS2@MoS2@NiS2 nanopolyhedra with a double-shell structure using cobalt-ZIFs as the precursor (Fig. 2k–m).95 The catalytic performance of CoS2@MoS2@NiS2 mainly originated from its double shell hollow structure, which exposed more active sites and provided a smooth diffusion path for the rapid transport of reactants and products during the reaction. The double shell structure could also limit the electrolyte in the shell, thus providing a high driving force for the electrochemical reaction. The OER overpotential of CoS2@MoS2@NiS2 was determined to be 200 mVRHE (Fig. 2n).
In conclusion, this chapter addressed electrocatalysts derived from cobalt-based ZIFs. Among them, the NP-Co3O4/CC electrocatalyst exhibited the best difunctional oxygen electrocatalytic performance. NP-Co3O4/CC was composed of ultrafine nitrogen-doped cobalt oxide nanoparticles on carbon cloth. The relevant electrochemical properties of Co-based ZIF derivatives are shown in Table 1.
Table 1 A comparison of cobalt-based ZIF derivatives for oxygen electrocatalysis and ZABs
Electrocatalysts |
ORR |
OER |
ΔE (VRHE) |
ZABs |
Ref. |
E
1/2 (VRHE) |
E
j=10 (VRHE) |
OCV (V) |
Peak power density (mW cm−2) |
Specific capacity (mA h g−1) |
Cycling stability |
Electrolyte |
Co@NCL |
0.84 |
1.63 |
0.79 |
1.47 |
170 |
766.35 |
200 h (10 mA cm−2) |
6 M KOH |
74
|
−0.2 M Zn(Ac)2 |
NC–Co/CoNx |
0.87 |
1.52 |
0.65 |
1.40 |
41.5 |
— |
1500 min (10 mA cm−2) |
KOH–ZnO gel |
75
|
Co-NCS@CNT |
0.86 |
2.07 |
1.21 |
1.42 |
90.6 |
789 |
80 h (5 mA cm−2) |
6 M KOH |
76
|
−0.2 M Zn(Ac)2 |
LIC-ZIF-67-M10 |
0.77 |
1.49 |
0.72 |
1.36 |
80 |
780 |
220 h (10 mA cm−2) |
6 M KOH |
77
|
−0.2 M Zn(Ac)2 |
Co@NCNT/Co-SA@NCMT |
0.87 |
1.29 |
0.65 |
1.4 |
120 |
846 |
300 h (5 mA cm−2) |
6 M KOH |
78
|
−0.2 M Zn(Ac)2 |
Co3O4–Co@NC-2 |
0.86 |
1.58 |
0.72 |
1.46 |
158 |
758 |
200 cycles (10 mA cm−2) |
6 M KOH |
79
|
−0.2 M Zn(Ac)2 |
NP–Co3O4/CC |
0.9 |
1.56 |
0.66 |
1.57 |
200 |
— |
400 h (5 mA cm−2) |
6 M KOH |
80
|
−0.2 M Zn(Ac)2 |
Co/Co9S8-NCL |
0.852 |
— |
— |
1.48 |
112 |
799 |
— |
6 M KOH |
81
|
−0.2 M Zn(Ac)2 |
ZIF-9_Fe3_Pyrol |
0.89 |
1.7 |
0.81 |
1.25 |
32 |
815 |
15 h (2 mA cm−2) |
6 M KOH |
82
|
−0.02 M ZnSO4 |
Mn/Co–N–C-0.02-800 |
0.80 |
1.66 |
0.86 |
1.39 |
136 |
— |
250 h (5 mA cm−2) |
6 M KOH |
83
|
−0.2 M Zn(OAc)2 |
Co–N–C/CNT |
0.87 |
1.58 |
0.71 |
— |
— |
— |
4 h (1 mA cm−2) |
polyvinyl alcohol–KOH–Zn(OAc)2 gel |
84
|
Co@N–PCP/NB-CNF-2–800 |
0.85 |
2.22 |
1.37 |
1.47 |
143.8 |
700 |
110 h (10 mA cm−2) |
6 M KOH |
85
|
−0.2 M Zn(Ac)2 |
Co–N–PCD |
0.886 |
— |
— |
1.41 |
259.4 |
773.53 |
25 h (20 mA cm−2) |
6 M KOH |
86
|
−0.2 M Zn(Ac)2 |
Co–O–ZIF/PANI |
0.7 |
351 mV (50 mA cm−2) |
— |
1.40 |
123.1 |
748.2 |
100 h (5 mA cm−2) |
6 M KOH |
87
|
−0.2 M Zn(Ac)2 |
D-Co@NC |
0.852 |
1.718 |
0.866 |
1.41 |
115.4 |
745.5 |
200 h (10 mA cm−2) |
6 M KOH |
88
|
−0.2 M Zn(Ac)2 |
Co4N@NC-2 |
0.84 |
1.519 |
0.679 |
1.48 |
74.3 |
769.4 |
750 h (5 mA cm−2) |
6 M KOH |
89
|
−0.2 M Zn(Ac)2 |
Co/NGC-3 |
0.85 |
1.63 |
0.78 |
1.40 |
134.4 |
716 |
120 h (5 mA cm−2) |
6 M KOH |
90
|
|
−0.2 M Zn(Ac)2 |
NiFe–Co@NC-450 |
0.833 |
1.69 |
0.857 |
1.44 |
100 |
798 |
200 h (10 mA cm−2) |
6 M KOH |
91
|
−0.2 M Zn(Ac)2 |
CoFe/NC |
0.82 |
— |
— |
— |
173 |
— |
230 h (10 mA cm−2) |
6 M KOH |
92
|
−0.2 M Zn(Ac)2 |
CoFe–Co@PNC |
0.887 |
1.55 |
0.663 |
1.46 |
152.8 |
786 |
200 h (10 mA cm−2) |
6 M KOH |
93
|
−0.2 M Zn(Ac)2 |
CuCo@NC |
0.866 |
1.568 (50 mA cm−2) |
— |
1.23 |
303.7 |
751.4 |
100 h (2 mA cm−2) |
6 M KOH |
94
|
−0.2 M ZnCl2 |
CoS2@MoS2@NiS2 |
0.80 |
2.00 |
1.20 |
1.39 |
80.28 |
708 |
24 h (10 mA cm−2) |
6 M KOH |
95
|
−0.2 M Zn(Ac)2 |
3.2 Co–M (M = Zn, Ni, Fe, etc.)-based ZIF derivatives
Many studies investigated the addition of other metals for the preparation of cobalt-ZIFs to improve their electrocatalytic performance. In contrast to monometallic catalysts, bimetallic catalysts usually exhibit a synergistic effect, which improves the catalytic activity by adjusting the electronic structure to optimize the adsorption and desorption of the reactants. There are many similar reports, and next, we will summarize and analyze the recent cobalt–M (M = Zn, Ni, Fe, etc.) ZIF derivatives.
3.2.1 C–Zn-based ZIF derivatives.
N-Doped amorphous carbon frameworks derived from ZIF-8 have a large surface area and porosity but poor conductivity. The graphite-carbon frame encapsulated by cobalt nanoparticles (Co-NPs) of ZIF-67 has higher electrical conductivity but lower surface area. Therefore, ZIF-67 and ZIF-8 were combined as catalyst precursors. The CoZn-ZIF derivatives have abundant Co–Nx, high conductivity, and high surface area. These advantages can significantly improve the ORR and OER activity of the catalyst. However, although CoZn-ZIFs combine the advantages of both cobalt-ZIFs and zinc-ZIFs, some of their aspects can be improved. For example, their structure collapses after high temperature treatment and they possess a low-density of Co–NC active sites and poor electrocatalytic stability. Consequently, these problems severely limit the development of CoZn-ZIFs in the field of catalysis, and many reports propose novel approaches to address them.
At present, many metal sulfides derived from CoZn-ZIFs are used as oxygen electrocatalysts, but the oxygen electrocatalytic performance of zinc sulfide derived from CoZn-ZIFs still needs to be enhanced. Huang et al. reported the synthesis of cobalt nanoparticle and ZnS-decorated N,S co-doped carbon nanotubes (ZnS/Co-NSCNTs) (Fig. 3a and b).96 According to the XPS analysis, ZnS/Co-NSCNTs contained a C–S–C bond, indicating that the production of CNTs was associated with the C–S–C bond formed under the catalysis of cobalt. In this novel composite, the N-doped porous carbon substrate of ZIF-8 derivatives stimulated the diffusion dynamics of O2, which enabled the C–S–C bond and zinc sulfide to absorb O2. The samples were tested for their ORR performance, and it was found that ZnS/Co-NSCNTs had the highest half-wave potential of 0.871 VRHE, indicating its high ORR catalytic performance (Fig. 3c). The ZnS/Co-NSCNT-based ZABs had a peak power density of 182.6 mW cm−2 and specific capacity of 755.0 mA h g−1 (Fig. 3d).
 |
| Fig. 3 (a) Schematic illustration describing the synthesis of ZnS/Co-NSCNTs. (b) HRTEM image of ZnS/Co-NSCNTs. (c) ORR LSV curves of ZnS/Co-NSCNTs. (d) Typical galvanostatic discharge curves (5 mA cm−2) of ZnS/Co-NSCNTs. Copyright 2021, Elsevier. (e) Schematic illustration of the synthesis of Co–N–PHCNTs. (f) SEM image of Co–N–PHCNTs. (g) ORR LSV curves of Co–N–PHCNTs. Copyright 2019, Elsevier. (h) Schematic illustration of the preparation of NGPC@CoOx. (i) SEM image of NGPC@CoOx. (j) ORR LSV curves of NGPC@CoOx. Copyright 2022, Elsevier. (k) Schematic illustration of the synthesis of LDH@N–CoOx@C. ORR (l) and OER (m) LSV curves of LDH@N–CoOx@C. Copyright 2022, Elsevier. | |
CoZn-ZIFs have the problem of structural collapse after high temperatures. In this case, many reports use carbon materials (carbon fiber,97 bacterial cellulose (BC),98 and PVP99) to protect the structure of CoZn-ZIFs from destruction and improve their electrical conductivity and catalytic properties. Pan et al. reported the synthesis of a self-supported N-doped Co/Zn carbon nanofiber membrane (Co/Zn@NCF) with superior ORR activity and durability.97 The cycle life of the liquid ZABs based on the Co/Zn@NCF catalyst exceeded 666 h, with a charge–discharge voltage difference of 0.64 VRHE. The prepared flexible solid-state ZABs enabled it to charge and discharge stably at different bending angles. Due to the three-dimensional porous structure of BC and the adsorption of many hydroxyl groups on the fiber surface, it could be used as an ideal supporting material. Zhang et al. prepared a new and three-dimensional porous carbon nanofiber network rich in CoNxC active sites, which was denoted as CoNC@N-CNF.98 Furthermore, Guan et al. used PVP as a surfactant to synthesize cobalt and N codoped carbon nanotubes embedded in a nitrogen-doped hexagonal carbon nanosheet, denoted as Co–N–PHCNTs (Fig. 3e and f).99 PVP could inhibit the vertical growth of CoZn-ZIFs to form a favorable two-dimensional hexahedral structure, which remained unchanged during the subsequent high-temperature pyrolysis. Co–N–PHCNTs showed good ORR activity with a high starting potential (Eonset = 0.98 VRHE) and a better half-wave potential of Co–N–PHCNTs (E1/2 = 0.89 VRHE) than that of Pt/C (E1/2 = 0.87 VRHE) (Fig. 3g).
Many reports have hybridized CoZn-ZIFs with carbon materials (CNTs,100 GO,101etc.) For example, Wang et al. reported the synthesis of Co3O4 nanoparticle-embedded carbon nanotubes (CNT–Co3O4/NC) using CoZn-ZIF as the precursor.100 The cobalt and Co3O4 nanoparticles played an essential role in the ORR performance. In addition, the graphite carbon on the catalyst provided good conductivity and improved the internal charge conductivity, while the carbon nanotubes grafted on the carbon frame improved the external conductivity. Therefore, CNT–Co3O4/NC had an excellent ORR catalytic performance with an ORR half-wave potential of −0.124 V (vs. Ag/AgCl). Zhao et al. reported the preparation of CoO/Co3O4 nanoparticles by low-temperature oxidation evenly distributed in three-dimensional layered nitrogen-doped graphene mesh (NGPC@CoOx) (Fig. 3h and i).101 The results demonstrated that CoZn-ZIFs could serve as a protective layer to prevent the severe accumulation of reduced graphene oxide (rGO) and as a precursor to construct a functional nanoporous carbon layer modified with highly distributed nitrogen and CoOx dopants. The ORR half-wave potential and OER overpotential of NGPC@CoOx were 0.86 and 1.61 VRHE, respectively, and the ZABs based on NGPC@CoOx had a peak power density of 184.4 mW cm−2 and a high specific capacity of 805.3 mA h g−1 (Fig. 3j). In addition, during the charge–discharge performance test at 2 mA cm−2, the initial charge–discharge voltage gap of the ZABs based on NGPC@CoOx was 0.76 V, which was low after 256 charge–discharge cycles and kept at 1 V.
Some reports hybridized CoZn-ZIFs with carbon materials to prepare single-atom catalysts. Li et al. reported a simple spatial isolation strategy for fabricating atomic cobalt catalytic sites anchored on layered porous N-doped carbon (Co-SAs/N–C/rGO) as catalysts for ORR in a wide pH range.102 The resultant Co-SAs/N–C/rGO catalysts exhibited pH-universal ORR activity with half-wave potentials (E1/2) of 0.84, 0.77, and 0.65 VRHE in basic, acidic, and neutral media, while the half-wave potentials for commercial Pt/C were 0.85,0.79 and 0.65 VRHE, respectively. These superior ORR catalytic activities were attributed to the combined action of the atomic active sites, larger specific surface area, and hierarchical porous structures. The OCV of liquid ZABs with Co-SAs/N–C/rGO as the cathode catalyst was 1.52 V, and the discharge capacity was 671.94 mA h g−1, which were better than that of the commercial Pt/C + RuO2 (1.49 V and 657.32 mA h g−1, respectively).
As a substance with high-efficiency catalytic properties, NiFe-LDH was hybridized with CoZn-ZIF-derived oxide to prepare an efficient catalyst. For instance, Hao et al. used a simple chemical method to combine NiFe-LDH with N–CoOx@C derived from CoZn-ZIFs to synthesize a bifunctional catalyst with a core–shell structure (LDH@N–CoOx@C) (Fig. 3k).103 The core–shell structures of CoOx and NiFe-LDH increased the number of defects in the catalyst and exposed more active sites, as confirmed by the ID/IG value of 1.03 for LDH@N–CoOx@C. The OER and ORR characterization of LDH@N–CoOx@C showed that its OER starting potential and ORR half-wave potential were 273 mVRHE and 0.838 VRHE (Fig. 3l and m), respectively. The LDH@N–CoOx@C-based ZAB exhibited a peak power density of 155.5 mW cm−2 and specific capacity of 685.6 mA h g−1 at 5 mA cm−2.
3.2.1.1 ZIF-67@ZIF-8 (core–shell structure) derivatives.
ZIF-67@ZIF-8 functions as a CoZn-ZIF with a particular core–shell structure and has been attracting significant attention. Precisely, the preparation of ZIF-67@ZIF-8 mainly involves two steps, where the first step is to mix Zn2+ and 2-MeIm in water or methanol solution to form ZIF-8, and the second step involves wrapping ZIF-67 formed by Co2+ and 2-MeIm on the surface of ZIF-8. Compared to traditional CoZn-ZIF derivatives, ZIF-8@ZIF-67 derivatives have a particular core–shell structure. The core–shell structure has the following advantages: (1) the hollow structure of the ZIF-8 core can provide channels for the material transport required for the catalytic reaction, and its outer layer is ZIF-67-derived Co@NC, which can give a rich active site for the catalytic reaction. (2) Because Zn is inside the material, Zn evaporation from inside to outside in the pyrolysis process is more conducive to the formation of vacancies and defects, better exposing the active site of the material and improving its electrocatalytic performance. (3) The interacting forces between the core and shell components allow modifications to the chemical/electronic configurations of the catalytic sites, which are critical for optimizing the reaction intermediate bindings and improving the intrinsic activity. (4) Core–shell nanostructures with unique active sites are expected to synergistically address multiple fundamental steps in water oxidation reactions.
Currently, there are many reports on ZIF-67@ZIF-8-derived catalysts, but there needs to be more research on the impact of the incorporation of Co2+ on the overall electrochemical performance. Zhu et al. explored the effect of the Co2+ content on the overall catalytic performance by adjusting the amount of Co2+ incorporated in the catalyst.104 This study showed that the incorporation of an appropriate content of Co2+ allowed the catalyst to obtain more carbon defects and cobalt–Nx, which could achieve an enhanced ORR catalytic performance. Among the samples, Co–N/PCNs-2 had the best ORR activity, with the highest half-wave potential of 0.85 VRHE. The open-circuit voltage of the ZAB based on Co–N/PCNs-2 was 1.49 VRHE and its maximum power density was 135 mW cm−2, while there was almost no voltage loss during 7 h discharge at the current density of 20 mA cm−2.
Simple pyrolysis is usually required to expose the active sites in the ZIF-67@ZIF-8 material. For instance, Li et al. reported the preparation of N-CNT hollow polyhedra using simple epitaxial growth and pyrolysis methods.105 The activity of Co@N-CNT-HC mainly originated from the efficient integration of 0D cobalt nanoparticles, 2D carbon nanotubes, and 3D hollow nanocarbons, which synergistically enhanced the interfacial reaction dynamics of oxygen and accelerated the charge transfer. The ORR half-wave potential of Co@N-CNT-HC was 0.84 VRHE. Similarly, Zhang et al. reported the simple one-step pyrolysis of ZIF-8@ZIF-67 to prepare a porous carbon material codoped with cobalt and nitrogen (Co–N@CNT-C800) (Fig. 4a and b).106 It was comprised of abundant cobalt and N codoped carbon nanotubes, with a large surface area (428 m2 g−1). The unique three-dimensional shape of Co–N@CNT-C800 guaranteed a particular surface area and higher porosity, facilitating mass and electron transfer for ORR. Consequently, Co–N@CNT-C800 exhibited a high ORR half-wave potential (0.841 VRHE) (Fig. 4c).
 |
| Fig. 4 (a) Schematic illustration of the preparation of Co–N@CNT-C800. (b) SEM image of Co–N@CNT-C800. (c) ORR LSV curves. Copyright 2023, Elsevier. (d) Schematic representation of the formation of Co, N-PCL. (e) and (f) SEM images of Co, N-PCL. (g) ORR LSV curves. Copyright 2019, Elsevier. (h) Schematic illustration of the three-step synthesis of ZnCoFe–N–C. (i) XPS spectrum of Zn 2p for ZnCoFe–N–C. (j) ORR LSV curves. Copyright 2021, the American Chemical Society. (k) Schematic Illustration of the synthesis process for Co/NHCNT. (l) TEM image of Co/NHCNT-160. (m) Chronoamperometric measurement of Co/NHCNT-160 and Pt/C. Copyright 2023, Elsevier. | |
In addition to the traditional three-dimensional structure of ZIF-8@ZIF-67 as the precursor, there are many recent reports on the preparation of catalysts using two-dimensional ZIF-L@ZIF-67 as the precursor. For instance, Park et al. reported the preparation of a highly active ORR catalyst (Co, N-PCL) using ZIF-L@ZIF-67 as the precursor (Fig. 4d and e).107 Co, N-PCL possessed many features that favor ORR, such as many carbon nanotubes, large surface area, and large pore volume (Fig. 4f). The ORR half-wave potential of Co, N-PCL was 0.846 VRHE (Fig. 4g).
Furthermore, several studies reported the use doping to improve the catalytic performance of ZIF-8@ZIF-67 derivatives. For example, Li et al. used iron-doped CoZn-ZIFs as precursors, preparing Zn, Co, and Fe triple-doped nitrogen–carbon nanocages (ZnCoFe–N–C) by annealing at high temperature (Fig. 4h).108 The XPS spectrum of ZnCoFe–N–C showed the presence of Zn in the sample, possibly because some Zn elements had formed stable Zn–Nx species before gasification. Co–Nx and Fe–Nx also existed in the sample (Fig. 4i). Due to the synergistic action of the three metals, the catalyst showed enhanced ORR catalytic activity with a half-wave potential of 0.878 VRHE (Fig. 4j). Subsequently, ZnCoFe–N–C was assembled into liquid ZABs, which had a peak power density and specific capacity of 350.2 mW cm−2 and 794.7 mA h g−1, respectively. The ZnCoFe–N–C-based ZABs were stabilized at 1.36 V for 40 h at a current density of 2 mA cm−2, and these results are better than that of Pt/C-based ZABs.
Transition metal sulfides exhibit excellent oxygen electrocatalytic performances. Accordingly, many reports use ZIF-8@ZIF-67 precursors to prepare metal sulfides. Peng et al. reported Co9S8 nanoparticles embedded in nitrogen-sulfur co-doped hollow carbon nanosheets (Co9S8/NSC-3) as a high-efficiency oxygen electrocatalyst, where 3 represents the mass ratio of TAA to ZIF-8@ZIF-67 precursor.109 Four coexisting XPS nitrogen types were observed for Co9S8/NSC-3, namely pyridine N, pyrrole N, graphite N, and oxidized N, in which pyridine N commonly serves as the active catalytic site of ORR. Co9S8/NSC-3 had a half-wave potential of 0.82 VRHE. Furthermore, the ZAB based on Co9S8/NSC-3 had a peak power density of 85 mW cm−2 and a high specific capacity of 804 mA h g−1. The ZABs based on Co9S8/NSC-3 also had a stable charge–discharge performance, with the voltage gap of around 0.962 VRHE, which was more durable than the ZAB based on Pt/C + RuO2.
To improve the electrical conductivity and electrocatalytic properties of ZIF-67@ZIF-8 derivatives, many reports hybridized them with better conductive carbon materials and carbon sources (rGO,110 GO,111 PAN,112 PS/PVP,113etc.) For instance, Qi et al. reported the synthesis of nitrogen-doped carbon nanotubes containing Co/Co4N nanoparticles (Co/Co4N@N-CNTs/rGO).110 The TEM images of Co/Co4N@N-CNTs/rGO showed that its surface was very rough, and the cobalt nanoparticles were encapsulated on top of N-CNTs. N-CNTs could effectively inhibit the aggregation or shedding of cobalt nanoparticles during long-term cycling. In addition, the BET characterization indicated the presence of mesopores and micropores in the sample, which could fully expose the active sites. The ORR half-wave potential and OER overpotential of Co/Co4N@N-CNTs/rGO were 0.85 VRHE and 372.1 mVRHE, respectively. The liquid ZABs based on Co/Co4N@N-CNTs/rGO had a peak power density of 200 mW cm−2 and specific capacity of 783 mA h g−1. The ZABs based on Co/Co4N@N-CNTs/rGO completed 144 h of stability tests and showed good stability. He et al. prepared a catalyst (Co/NHCNT-160) by using polystyrene (PS) spheres as the sacrificial template and ZIF-8@ZIF-67 and PVP as cobalt–NC sources and adhesives, (Fig. 4k and l), respectively.113 Co/NHCNT-160 possessed a unique hollow core–shell structure and abundant carbon nanotubes with excellent stability, abundant electron conduction channels, high specific surface area and three-dimensional interconnection. Consequently, Co/NHCNT-160 had a high starting potential (0.955 VRHE), half-wave potential (0.88 VRHE), limited current density (6.3 mA cm−2), and current retention rate of 99.7% after 7 h (Fig. 4m).
3.2.2 Co–Ni-based ZIF derivatives.
As an efficient active site, Ni-NC has attracted significant attention. Moreover, the cooperation of cobalt and nickel nanoparticles enriches the active center and enhances the intrinsic catalytic activity, thus improving the catalytic performance. Therefore, increasing reports use CoNi-ZIFs as precursors to prepare efficient electrocatalysts.
CoNi-ZIFs usually require high-temperature pyrolysis to expose their active sites. For example, Ahmed et al. reported the preparation of an Ni-modified bimetallic cobalt nickel catalyst rich in carbon nanotubes (Ni/CNT/CoNi) via a simple pyrolysis process.114 The carbon nanotubes were observed in the SEM images of Ni/CNT/CoNi. In addition, the TEM characterization of the samples showed that the cobalt distribution in CNT/CoNi was uniform, and all Ni almost wholly overlapped with Co, indicating the presence of bimetallic CoNi. Ning et al. reported the preparation of porous N-doped carbon-encapsulated CoNi alloy nanoparticles (Co1Ni1@N–C) derived from CoNi-ZIF as efficient bifunctional oxygen electrocatalysts.115 This report explored the molar ratio of cobalt and nickel elements and the pyrolysis temperature. The results indicated that the optimal molar ratio of cobalt and nickel elements was 1
:
1, and the optimal pyrolysis temperature was 800 °C. The ORR half-wave potential of Co1Ni1@N–C was 0.82 VRHE.
The wrapping layer of SiO2 could effectively prevent the aggregation of metal nanoparticles and the collapse of the carbon matrix structure under high-temperature calcination conditions. Li et al. reported the use of an SiO2-wrapped CoNi-ZIF as the precursor to obtain CoNi@NPC with a core–shell structure (Fig. 5a).116 According to the TEM images, they observed that CoNi alloy nanoparticles were evenly distributed in the central region of NPC. NPC built a good conductive layered shell, which is conducive to accelerating electricity/mass transfer and exposing the active sites. The binding of CoNi alloy nanoparticles with NPC could improve the catalytic activity and stability. Jiang et al. reported that the CoNi nanoalloy in N-doped porous carbon frameworks (CoNi-NCF) was obtained by the pyrolysis of SiO2-modified CoNi-ZIF.117
 |
| Fig. 5 (a) Schematic illustration of the preparation of CoNi@NPC. Copyright 2022, Elsevier. (b) Illustration of the synthesis of CoFe/N-HCSs. Copyright 2021, Elsevier. (c) Schematic illustration of the preparation process of Co3O4/Mn3O4/CNx@CNFs. ORR (d) and OER (e) LSV curves of Co3O4/Mn3O4/CNx@CNFs. Copyright 2020, Elsevier. (f) Schematic illustration of the synthesis procedure of Ru–Cl-N SAC. (g) AC-HAADFSTEM image of Ru–Cl-N SAC. (h) ORR LSV curves of Ru–Cl-N SAC. Copyright 2021, Elsevier. (i) Specific capacity curves of Ru–Cl-N SAC. Copyright 2022, Elsevier. (j) Synthesis route of Co9S8/CeO2/Co–NC. (k) SEM image of Co9S8/CeO2/Co–NC. ORR (l) and OER (m) LSV curves of Co9S8/CeO2/Co–NC. Copyright 2021, The Royal Society of Chemistry. | |
3.2.3 Co–Fe-based ZIF derivatives.
As a common element with great catalytic potential, Fe is often used in various forms in many fields, where the Fe element is also employed in the field of oxygen electrocatalysis. Fe almost overlaps with Ni and Co in the volcano map, indicating its high electrocatalytic potential. Therefore, many reports combine Co and Fe to prepare CoFe-ZIFs and use novel methods to explore and improve their electrocatalytic performance.
Many reports used CoFe-ZIFs as a precursor to prepare oxides. For example, Chong et al. produced a catalyst with one- and two-dimensional structures grown on flexible carbon cloth (Co1−xFexO@NC).118 The TEM image of Co1−xFexO@NC showed the presence of two-dimensional nanosheets on one-dimensional nanowires, and this structure could increase the surface area of the catalyst and improve its catalytic performance. The unique structure of Co1−xFexO@NC also resulted in excellent catalytic performances with an ORR half-wave potential and OER overpotential of 0.81 VRHE and 260 mVRHE, respectively.
Some reports hybridized CoFe ZIFs with carbon materials and carbon sources (rGO,119 hollow polypyrrole spheres (h-PPy),120etc.) to prepare metal alloy catalysts. Liu et al. designed a reduced graphene oxide-wrapped Co9–xFexS8/Co,Fe–N–C composite (S-Co9–xFexS8@rGO) as a bifunctional oxygen electrocatalyst.119 This report explored the effects of one-step carbonization, simple semi-vulcanization followed by carbonization, and deep-vulcanization followed by carbonization on the catalytic performance. Li et al. designed a template-assisted method to reduce the crystal size of CoFe-ZIF via the in situ growth of hollow polypyrrole spheres (h-PPy) (Fig. 5b).120 As a result, well-dispersed ultrafine CoFe nanoalloy was obtained with N-doped hollow carbon microspheres (CoFe/N-HCSs) as the support. Owing to its unique alloy structure with ultrafine size and the uniform dispersion of the CoFe nanoalloy on the highly conductive N-HCS support via strong interactions, the proposed CoFe/N-HCSs exhibited high catalytic activity and stability toward ORR.
3.2.4 Others.
In addition to Zn, Ni, and Fe, many other metals, such as Mn,121 La,122 Ce,123 and Ru,124 can be combined with cobalt to prepare ZIF precursors. This chapter summarizes the CoM-ZIFs formed by co-binding with these metals and the methods employed to improve their electrocatalytic properties.
Li et al. prepared highly efficient electrocatalysts (Co3O4/Mn3O4/CNx@CNFs) using electrospinning, pyrolysis and oxidation processes (Fig. 5c).121 The XPS characterization results of Co3O4/Mn3O4/CNx@CNFs proved the presence of Co3O4, Mn3O4, Co–Nx and Mn–Nx. Many reports demonstrated the excellent electrocatalytic performance of Co3O4, Mn3O4, Co–Nx, and Mn–Nx. This study tested their ORR and OER, showing a high ORR half-wave potential of 0.85 VRHE and low OER overpotential of 1.63 VRHE (Fig. 5d and e), respectively. The potential difference between OER and ORR was only 0.78 VRHE, demonstrating that Co3O4/Mn3O4/CNx@CNFs had efficient ORR and OER bifunctional catalytic activity.
Zhou et al. designed a high-efficiency bifunctional oxygen electrocatalyst (La2O3–Co/NC) with CoLa-ZIFs as the precursor.122 The SEM images showed no significant change in the La2O3–Co/NC structure after CoLa-ZIFs/AB through pyrolysis at high temperatures. Meanwhile, the CoLa-ZIFs not attached to AB underwent significant structural changes after pyrolysis, indicating that AB benefited the structural maintenance of CoLa-ZIFs after exposure to high temperatures. La2O3–Co/NC had a high ORR half-wave potential of 0.86 VRHE.
Sun et al. designed a novel two-dimensional “senbei”-like nitrogen-doped carbon nanosheet (Co9S8/CeO2/Co–NC) (Fig. 5j and k).123 A phase transition from three-dimensional spherical cobalt-ZIFs to two-dimensional Co/Ce-ZIFs was achieved by introducing Ce ions. Co9S8/CeO2/Co–NC with a highly open two-dimensional structure and numerous distributed carbon nanotubes increased the specific surface area of the catalyst, thus providing more active sites. In addition, Co9S8/CeO2/Co–NC, with a unique Co9S8/CeO2 heterostructure and the synergy of the two components, showed excellent electrocatalytic performances in ORR and OER. The ORR half-wave potential and OER overpotential of Co9S8/CeO2/Co–NC were 0.875 and 1.60 VRHE (Fig. 5l and m), respectively. Furthermore, when used as a bifunctional air electrode, Co9S8/CeO2/Co–NC reached a peak power density of 164.24 mW cm−2, showing an outstanding cyclic charge–discharge stability (over 668 h) at a current density of 5 mA cm−2.
Chen et al. designed a Cl, N-coordinated Ru single-atom multifunctional electrocatalyst (Ru–Cl-NSAC) with CoRu-ZIFs as the precursor.124 The ORR half-wave potential of Ru–Cl-NSAC was 0.90 VRHE. The ZABs based on Ru–Cl-NSAC exhibited a specific capacity of 804.26 mA h g−1 at a current density of 10 mA cm−2. The excellent performance of the catalyst was mainly due to the following reasons: (1) the ZIF-67-derived three-dimensional porous carbon frame facilitated the charge and mass transport between the electrocatalyst and the electrolyte. (2) The doping of different types of N atoms could regulate the electronic structure of the carbon substrate. (3) The introduction of Cl and N coordinating with Ru atoms could adjust the electronic environment of the Ru single atom and reduce the adsorption free energy of oxygen-containing intermediates, thus significantly improving the OER and ORR performances.
Briefly, this chapter summarized the synthesis and modification methods of Co–M (M = Zn, Ni, Fe, Mo, etc.) ZIF derivatives. Among them, the electrocatalysts of Ru–Cl-NSAC showed the best difunctional oxygen electrocatalytic performance. The introduction of Cl and N coordinating with Ru atoms reduced the adsorption free energy of oxygen-containing intermediates. Therefore, the Ru–Cl-NSAC has excellent oxygen electrocatalytic performance. The relevant electrochemical properties of cobalt–M-based ZIF derivatives are shown in Table 2.
Table 2 A comparison of cobalt-based bimetallic ZIF derivatives for oxygen electrocatalysis and ZABs
Electrocatalysts |
ORR |
OER |
ΔE (VRHE) |
ZABs |
Ref. |
E
1/2 (VRHE) |
E
j=10 (VRHE) |
OCV (V) |
Peak power density (mW cm−2) |
Specific capacity (mA h g−1) |
Cycling stability |
Electrolyte |
ZnS/Co–NSCNTs |
0.871 |
— |
— |
1.52 |
182.6 |
750.1 |
10 h (5 mA cm−2) |
6 M KOH |
96
|
−0.2 M Zn(OAc)2 |
Co/Zn@NCF |
0.84 |
1.69 |
0.85 |
1.42 |
202 |
760 |
666 h (5 mA cm−2) |
Polyacrylate-KOH–ZnCl2 |
97
|
CoNC@N-CNF |
0.8 |
1.61 |
0.81 |
1.43 |
308 |
— |
63 h (5 mA cm−2) |
6 M KOH |
98
|
Co–N–PHCNTs |
0.89 |
1.62 |
0.73 |
1.4 |
125.14 |
— |
673 h (5 mA cm−2) |
6 M KOH |
99
|
−0.2 M Zn(Ac)2 |
CNT–Co3O4/NC |
−0.124 (AgCl) |
0.819 (AgCl) |
0.943 |
1.38 |
267 |
814 |
300 cycles (15 mA cm−2) |
6 M KOH |
100
|
NGPC@CoOx |
0.86 |
1.61 |
0.75 |
1.41 |
184.4 |
805.3 |
32 h (2 mA cm−2) |
6 M KOH |
101
|
−0.2 M Zn(Ac)2 |
Co-SAs/N–C/rGO |
0.84 |
— |
— |
1.52 |
104.91 |
671.94 |
26.31 h (1 mA cm−2) |
6 M KOH |
102
|
−0.2 M Zn(OAc)2 |
LDH@N–CoOx@C |
0.838 |
1.503 |
0.665 |
1.43 |
155.5 |
685.6 |
150 h (5 mA cm−2) |
6 M KOH |
103
|
−0.2 M Zn(Ac)2 |
Co–N/PCNs-2 |
0.88 |
— |
— |
1.49 |
135 |
— |
7 h (20 mA cm−2) |
6 M KOH |
104
|
−0.2 M Zn(Ac)2 |
Co@N-CNT-HC |
0.84 |
|
|
|
|
|
|
|
105
|
Co–N@CNT-C800 |
0.84 |
|
|
|
|
|
|
|
106
|
Co, N-PCL |
0.83 |
|
|
|
|
|
|
|
107
|
ZnCoFe–N–C |
0.878 |
— |
— |
1.52 |
350.2 |
794.7 |
10 h (5 mA cm−2) |
6 M KOH |
108
|
Co9S8/NSC-3 |
0.82 |
1.79 |
0.97 |
1.42 |
85 |
804 |
140 h (10 mA cm−2) |
6 M KOH |
109
|
−0.2 M Zn(Ac)2 |
Co/Co4N@N-CNTs/rGO |
— |
1.6 |
— |
1.33 |
200 |
783 |
440 h (5 mA cm−2) |
6 M KOH |
110
|
Co@N-HPCF-800 |
0.831 |
— |
— |
1.40 |
136.2 |
151 |
50 h (5 mA cm−2) |
6 M KOH |
111
|
−0.2 M Zn(Ac)2 |
Co–N–C@GNP |
0.86 |
1.64 |
0.78 |
1.60 |
236.2 |
— |
91 h (5 mA cm−2) |
gel |
112
|
Co/NHCNT-160 |
0.88 |
|
|
|
|
|
|
|
113
|
Ni/CNT/CoNi |
|
|
|
|
|
|
520 h (5 mA cm−2) |
6 M KOH |
114
|
−0.2 M Zn(Ac)2 |
Co1Ni1@N–C |
0.82 |
|
|
|
|
|
|
|
115
|
CoNi@NPC |
0.77 |
1.72 |
0.95 |
1.54 |
224 |
— |
40 h (5 mA cm−2) |
6 M KOH |
116
|
Co1−xFexO@NC |
0.81 |
1.49 |
0.68 |
1.30 |
52 |
673 |
150 h (5 mA cm−2) |
6 M KOH |
118
|
−0.2 M Zn(Ac)2 |
S-Co9−xFexS8@rGO |
0.84 |
1.52 |
0.68 |
|
|
|
|
|
119
|
CoFe/N-HCSs |
0.791 |
1.52 |
0.729 |
1.387 |
96.5 |
777.4 |
160 h (5 mA cm−2) |
6 M KOH |
120
|
−0.2 M Zn(Ac)2 |
Co3O4/Mn3O4/CNx@CNFs |
0.78 |
1.56 |
0.78 |
1.51 |
265 |
— |
50 h (5 mA cm−2) |
6 M KOH |
121
|
La2O3–Co/NC |
0.86 |
1.53 |
0.67 |
1.48 |
80.12 |
— |
170 h (2 mA cm−2) |
6 M KOH |
122
|
−0.2 M Zn(Ac)2 |
Co9S8/CeO2/Co–NC |
0.875 |
1.60 |
0.725 |
1.4 |
164.24 |
— |
668 h (5 mA cm−2) |
6 M KOH |
123
|
−0.2 M Zn(Ac)2 |
Ru–Cl-NSAC |
0.90 |
1.46 |
0.56 |
1.45 |
205 |
804.26 |
360 h (2 mA cm−2) |
6 M KOH |
124
|
−0.2 M Zn(Ac)2 |
3.3 Co-based trimetallic ZIF derivatives
The construction of bifunctional electrocatalysts with two or more different components is essential to provide abundant OER/ORR active sites, which show better OER and ORR performances than single-component catalysts. Through the interaction of trimetallic carbon and nitrogen compounds, they can reduce the binding energy of the reactants, intermediates, and active substances. Therefore, many reports employ polymetallic ZIF derivatives as bifunctional oxygen electrocatalysts.
Many reports have demonstrated that Co–NC and Fe–NC have excellent ORR and OER catalytic properties, and thus an increasing number of studies focus on compounds containing CoFe, among which CoZnFe-ZIFs have attracted much attention as single-atom catalyst precursors. For example, Zhong et al. used hydrothermal and pyrolysis processes to prepare cobalt-iron single atoms (CoFe–NC-1) anchored to a nitrogen-doped carbon substrate (Fig. 6a and b).125 The high ID/IG values in the Raman spectrum of FeCo–NC-1 indicated that this catalyst has many defects, favoring the exposure of the active sites. The BET characterization results for FeCo–NC-1 demonstrated that it possessed microporous and medium pores. The unique structure of FeCo–NC-1 also resulted in excellent catalytic properties, with an ORR half-wave potential of 0.84 VRHE (Fig. 6c). The FeCo–NC-1-based ZABs had a peak power density of 122.0 mV cm−2, specific capacity of 726.2 mA h g−1, and good stability, which exceeded that of Pt/C-based ZABs. Similar reports on a single atoms have also been reported. Chen et al. utilized a one-step pyrolysis method, where the precursor CoZnFe-ZIF was converted to a nitrogen-doped carbon material containing Fe and Co double single-atom electrocatalysts (FeCo-IA-NC) (Fig. 6d and e).126 The synergy between the atomic-dispersed Fe and Co in FeCo-IA-NC provided many active sites for the catalyst. Therefore, FeCo-IA-NC showed an excellent ORR performance with an ORR half-wave potential of 0.88 VRHE, reaction electron transfer number between 3.66 and 3.95, and reactive hydrogen peroxide yield of less than 8% (Fig. 6f and g).
 |
| Fig. 6 (a) Schematic illustration of the synthetic process of FeCo–NC. (b) HAADF-STEM image of FeCo–NC. (c) ORR LSV curves of FeCo–NC. Copyright 2019, Elsevier. (d) Synthesis route of FeCo-IA/NC. (e) AC-STEM image of FeCo-IA/NC. (f) ORR LSV curves FeCo-IA/NC. (g) K–L plots at different potentials FeCo-IA/NC. Copyright 2020, The Royal Society of Chemistry. (h) Synthesis route of Co9S8/MnS-USNC. (i) SEM image of Co9S8/MnS-USNC. (j) XRD patterns of Co9S8/MnS-USNC. (k) Cycling test of the all-solid-state zinc–air batteries based on Co9S8/MnS-USNC. Copyright 2021, The Royal Society of Chemistry. (l) Schematic illustration of the synthetic strategy of CoNi/N-CNN. (m) Typical Raman spectra image of CoNi/N-CNN. (n) ORR LSV curves of CoNi/N-CNN. (o) Polarization and power density plot of CoNi/N-CNN-based zinc–air batteries. Copyright 2022, Elsevier. | |
In addition to preparing single-atom catalysts through one-step pyrolysis of CoZnFe-ZIFs, there are also some reports that first hybridize CoZnFe-ZIFs with carbon materials, and then perform pyrolysis to prepare single-atom catalysts. Nguyen et al. used polystyrene ball (PS)-coated polypyrrole (PPy) as the core and CoZnFe-ZIFs as the shell to obtain a hollow structure catalyst (H–FeCo–NC).127 Given that the spherical surface of PS-PPy obtained after sodium sulfite treatment had a negative charge, it could attract Co and Zn ions, facilitating the growth of CoZn-ZIFs on its surface and avoiding the possible structural collapse of CoZn-ZIFs after pyrolysis at high temperature. The unique hollow structure of H–FeCo–NC and the synergistic interaction of the Fe single-atom and cobalt nanoparticles inferred its efficient catalytic performance. The ORR half-wave potential of H–FeCo–NC was 0.888 VRHE.
Some reports have also hybridized CoZnFe-ZIFs with carbon materials to prepare metal oxide catalysts. Sun et al. prepared self-supporting electrodes composed of two-dimensional nanosheets (Co3O4/Fe2O3NAs@CNFs) embedded with Co3O4 and Fe2O3 composed of carbon nanofibers by electrospinning and pyrolysis.128 The XPS characterization of Co3O4/Fe2O3NAs@CNFs illustrated the presence of Fe–Nx and Co–Nx, which were the active substances of ORR and provided the ORR catalytic performance of the catalyst. The ORR half-wave potential of Co3O4/Fe2O3NAs@CNFs was 0.80 VRHE, and its ZABs had a peak power density of 246 mV cm−2, exceeding that of 175 mV cm−2 for Pt/C + RuO2, and it also showed a better performance than Pt/C + RuO2 in the stability test.
In addition to compounds of CoZnFe-ZIFs, many reports used CoZnMn-ZIF derivatives as bifunctional electrocatalysts. Ye et al. designed catalysts (Co/Co2Mn3O8@NCFs) with a unique three-dimensional structure.129 The Co/Co2Mn3O8@NCFs had a unique three-dimensional structure, coupled with the rich active sites brought by Co2Mn3O8 and Co–Nx, which would have good bifunctional catalytic activity. The study tested their ORR and OER performances, and the test results showed that their ORR half-wave potential was 0.826 VRHE, and their OER catalytic performance was also better than that of RuO2. Li et al. used CoZnMn-ZIFs as the precursor to prepare a two-dimensional petal-thin Co9S8/MnS/S/N codoped carbon nanosheet (Co9S8/MnS-USNC) (Fig. 6h).130 The SEM image of Co9S8/MnS-USNC showed the presence of a petal-like structure and a large number of metal particles on its surface (Fig. 6i). The XPS and XRD characterization demonstrated the successful doping of S and N in the carbon substrate and the presence of two active species (Co9S8 and MnS) (Fig. 6j). The characterization of the ORR and OER performances of Co9S8/MnS-USNC showed an ORR half-wave potential of 0.90 VRHE and OER overpotential of 360 mVRHE. The ZABs based on Co9S8/MnS-USNC performed better than that based on Pt/C + Ir/C in the stability tests (Fig. 6k).
Ni–NC has also attracted significant attention as a substance with high-efficiency bifunctional electrocatalytic properties similar to that of Fe–NC and Mn–NC. Many reports used CoZnNi-ZIFs as precursors to prepare CoNi compounds as bifunctional electrocatalysts. For example, Li et al. preformed simple pyrolysis to produce a sea-like structural catalyst (CoNi/N-CNN) composed of carbon nanotubes embedded with CoNi nanoparticles (Fig. 6l).131 According to the Raman spectrum of CoNi/N-CNN, its ID/IG value was about 1.1, indicating a good balance between the degree of graphitization and the number of carbon defects, which could provide a lot of active sites and ensure good conductivity (Fig. 6m). The ORR and OER performance tests on CoNi/N-CNN showed that its ORR half-wave potential was 0.813 VRHE and its OER overpotential was 1.718 VRHE. The ZABs based on CoNi/N-CNN had a peak power density of 209 mV cm−2 and specific capacity of 810 mA h g−1 (Fig. 6n and o), respectively.
In short, this chapter summarized the methods for the synthesis and modification of cobalt-based trimetallic ZIF derivatives. The electrocatalysts of Co9S8/MnS-USNC have the best difunctional oxygen electrocatalytic performances. The synergistic effect of Co9S8 and MnS provides a good oxygen electrocatalytic performance. The relevant electrochemical properties of cobalt-based monometallic ZIF derivatives are shown in Table 3.
Table 3 A comparison of cobalt-based trimetallic ZIF derivatives for oxygen electrocatalysis and ZABs
Electrocatalysts |
ORR |
OER |
ΔE (VRHE) |
ZABs |
Ref. |
E
1/2 (VRHE) |
E
j=10 (VRHE) |
OCV (V) |
Peak power density (mW cm−2) |
Specific capacity (mAh g−1) |
Cycling stability |
Electrolyte |
CoFe–NC-1 |
0.84 |
— |
— |
1.50 |
122.0 |
726.2 |
— |
6 M KOH |
125
|
−0.2 M Zn(Ac)2 |
FeCo-IA-NC |
0.88 |
— |
— |
1.472 |
115.6 |
635.3 |
100 cycles (10 mA cm−2) |
6 M KOH |
126
|
−0.2 M ZnCl2 |
H–FeCo–NC |
0.888 |
|
|
|
|
|
|
|
127
|
Co3O4/Fe2O3NAs@CNFs |
0.80 |
1.6 |
0.8 |
1.533 |
246 |
— |
— |
6 M KOH |
128
|
−0.2 M Zn(Ac)2 |
Co/Co2Mn3O8@NCFs |
0.82 |
— |
— |
1.45 |
118.19 |
841.83 |
120 h (10 mA cm−2) |
6 M KOH |
129
|
−0.2 M Zn(Ac)2 |
Co9S8/MnS-USNC |
0.90 |
1.594 |
0.694 |
1.45 |
146 |
802 |
600 h (5 mA cm−2) |
18 M KOH |
130
|
−0.02 M Zn(Ac)2 |
CoNi/N-CNN |
0.819 |
1.718 |
0.899 |
1.42 |
209 |
810 |
90 h (2 mA cm−2) |
6 M KOH |
131
|
4. Summary and future perspective
In summary, the main problem associated with the current ZABs is the design of dual-functional electrocatalysts with high ORR and OER activity and cheap and safe materials that can replace the precious metal-based catalysts. ZIFs have received significant attention as a new type of crystalline porous materials with large surface area, high porosity, large pores, large volume, easy functionalization, and the production of highly conductive carbon-based electrocatalysts. In contrast to other non-noble metal ZIF derivatives, most of the cobalt-containing ZIF derivatives showed high ORR and OER activity, as described earlier herein. This review demonstrated that cobalt-containing ZIF derivatives can be used as various electrocatalytic materials to enhance ORR and OER and applied in ZABs. With the combination of different metal species, these materials were classified into cobalt-based ZIF derivatives, cobalt–M based ZIF derivatives, and cobalt-based trimetallic ZIF derivatives. This review highlighted the advantages of these cobalt-containing ZIF materials, their synthetic and modification parameters, various derivative compounds, and synergistic effects of metals. These cobalt-containing ZIF derivatives showed better ORR and OER activity and ZAB performances due to the synergistic generation of more active sites and tunable metal composition.
Despite the massive surge in the number of cobalt-containing ZIF derivatives in this area, various deficiencies still need attention.
(1) The main drawbacks in their electrochemical performance are well known: the original cobalt-ZIFs and cobalt-ZIF composites have poor electrical conductivity. A possible solution to this problem is synthesizing cobalt-containing ZIF derivatives with better electrical conductivity. In recent years, many researchers have attempted to design cobalt-containing ZIF derivatives with better electrical conductivity using the following methods: (i) combining better conductive materials (carbon nanotubes, graphene, PVP, etc.) with cobalt-containing ZIF derivatives and (ii) using the dual-ligand strategy to synthesize cobalt-containing ZIF derivatives. The typical organic ligands with good electrical conductivity are BIm, SMG, and GNP. However, most of these studies will affect the morphology and performance of cobalt-containing ZIF derivatives. Therefore, more studies are needed to improve the conductivity of cobalt-containing ZIF derivatives and make them ideal for electrocatalysis.
(2) To expose more active sites in cobalt-containing ZIF derivatives, pyrolysis is usually required to prepare cobalt-containing ZIF derivatives. This may cause problems such as the collapse of the structure of cobalt-containing ZIF derivatives and the agglomeration of metal particles. Thus, in recent years, many reports have adopted novel approaches to address these problems, as follows: (i) using a surfactant as a structure mediator (α-amylase, PVP, Pluronic F127, etc.) and organic ligand competitive coordination and regulating the nucleation and growth process of cobalt-containing ZIF precursors to construct more stable cobalt-containing ZIF derivatives. (ii) Using template-assisted strategies to synthetize, common template materials such as CNTs, PS and PAN, where these materials are wrapped on the surface of cobalt-containing ZIF derivatives to avoid their structural collapse during high-temperature pyrolysis. However, these methods usually require one-step or multi-step experimental steps, increasing the operation and time of material synthesis. In addition, these methods are followed by pyrolysis, which requires an additional energy-consuming step. Therefore, discovering other methods to expose more active sites is required.
(3) Cobalt-containing ZIF derivatives have achieved more significant development as the air cathode of ZABs. For example, cobalt-based ZIF derivatives, cobalt–M-based ZIF derivatives, and cobalt-based trimetallic ZIF derivatives with a sodalite (SOD) topology are widely used in the synthesis of cobalt-containing ZIF derivative-based air electrodes. However, the remaining topologies, such as rhodesite (RHO), Linde-type A (LTA), and gismondite (GIS), are rarely utilized. ZIFs with any topology structure have permanent porosity and high thermal and chemical stability, which is one reason why they have become attractive candidate materials for many applications, such as gas separation and storage. However, the practical application and development of ZIFs with different topologies still need to be improved. For example, LTA ZIFs with good ion exchange, adsorption, and catalytic performance are only used as water softeners in detergents and catalysts for petroleum refining. Therefore, the synthesis of cobalt-containing ZIF derivative-based air electrodes has not been fully developed, and researchers need to develop cobalt-containing ZIF derivatives with different topologies and link them to their electrocatalytic properties, design novel and efficient bifunctional electrocatalysts and apply them in ZABs.
(4) In the process of ORR and OER, it is difficult to determine the main catalytic active substances and the reactions/products during the reaction affecting these active substances. Therefore, advanced operation or in situ detection methods are required to determine the active substances, which will help researchers better understand the natural reaction mechanism of these materials and improve the catalytic performance and stability of catalysts.
Although many challenges remain, investigating the electrocatalytic applications of cobalt-containing ZIF derivatives, as well as a deeper understanding of their stability and reaction mechanism, will provide more information to guide researchers to design better cobalt-containing ZIF derivative-related electrocatalytic materials, which is expected to enable large-scale applications in industrial applications. The common methods for the industrial preparation of catalysts include solid-state method, hydrothermal method, and ball milling method. Thus, the application of laboratory-prepared cobalt-containing ZIF derivative catalysts in industry still has a long way to go.
Abbreviations
ZABs | Zinc air batteries |
ZIFs | Zeolitic imidazolate frameworks |
ORR | Oxygen reduction reaction |
OER | Oxygen evolution reaction |
MOFs | Metal–organic frameworks |
SMG | Supramolecular gels |
BIm | Benzimidazole |
HER | Hydrogen evolution reaction |
LDH | Layered double hydroxides |
ECSA | Electrochemical active surface area |
TOF | Turnover frequency |
3D | Three-dimensional |
2D | Two-dimensional |
NPs | Nitride nanoparticles |
CNT | Carbon nano-tube |
LIC | Laser-induced carbonization |
DTO | Dithiooxamide |
LED | Light-emitting diode |
PANI | Polyaniline |
NGC | Nitrogen-doped graphite carbon |
O2-TPD | Temperature-programmed desorption of oxygen |
h-PPy | Hollow polypyrrole |
SEM | Scanning electron microscopy |
BC | Bacterial cellulose |
rGO | Reduced graphene oxide |
OCV | Open circuit voltage |
ALD | Atomic layer deposition |
EPR | Electron paramagnetic resonance |
PS | Polystyrene |
PPy | Polypyrrole |
PVP | Polyvinyl pyrrolidone |
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
The work was supported by the PhD Research Startup Fund (BK202013) from Hubei University of Automotive Technology (HUAT), the Open Fund (QCCLSZK2021A02) from Hubei Key Laboratory of Critical Materials of New Energy Vehicles (HUAT), the Open Fund (ZDK1202201) from Hubei Key Laboratory of Automotive Power Train and Electronics (HUAT) and the Open Fund (ZDK22023A05) from Hubei Key Laboratory of Energy Storage and Power Battery (HUAT).
References
- Y. Arafat, M. R. Azhar, Y. Zhong, M. O. Tadé and Z. Shao, Metal-free carbon based air electrodes for Zn–air batteries: recent advances and perspective, Mater. Res. Bull., 2021, 140, 111315 CrossRef CAS
.
- W. Fang, J. Zhao, W. Zhang, P. Chen, Z. Bai and M. Wu, Recent progress and future perspectives of flexible Zn–Air batteries, J. Alloys Compd., 2021, 869, 158918 CrossRef CAS
.
- W. Xue, Q. Zhou, X. Cui, S. Jia, J. Zhang and Z. Lin, Metal–organic frameworks-derived heteroatom-doped carbon electrocatalysts for oxygen reduction reaction, Nano Energy, 2021, 86, 106073 CrossRef CAS
.
- X. Zhang and L. Wang, Research progress of carbon nanofiber-based precious-metal-free oxygen reaction catalysts synthesized by electrospinning for Zn–Air batteries, J. Power Sources, 2021, 507, 230280 CrossRef CAS
.
- W. Zhao, X. Ma, L. Gao, X. Wang, Y. Luo, Y. Wang, T. Li, B. Ying, D. Zheng and S. Sun, Hierarchical Architecture Engineering of Branch-Leaf-Shaped Cobalt Phosphosulfide Quantum Dots: Enabling Multi-Dimensional Ion-Transport Channels for High-Efficiency Sodium Storage, Adv. Mater., 2024, 36, 2305190 CrossRef CAS PubMed
.
- S. Hosseini, S. M. Soltani and Y.-Y. Li, Current status and technical challenges of electrolytes in zinc–air batteries: an in-depth review, Chem. Eng. J., 2021, 408, 127241 CrossRef CAS
.
- Y. Zhang, Y.-P. Deng, J. Wang, Y. Jiang, G. Cui, L. Shui, A. Yu, X. Wang and Z. Chen, Recent progress on flexible Zn–air batteries, Energy Storage Mater., 2021, 35, 538–549 CrossRef
.
- L. Yue, D. Wang, Z. Wu, W. Zhao, Y. Ren, L. Zhang, B. Zhong, N. Li, B. Tang and Q. Liu, Polyrrole-encapsulated Cu2Se nanosheets in situ grown on Cu mesh for high stability sodium-ion battery anode, Chem. Eng. J., 2022, 433, 134477 CrossRef CAS
.
- S. Wang, S. Chen, L. Ma and J. A. Zapien, Recent progress in cobalt-based carbon materials as oxygen electrocatalysts for zinc–air battery applications, Mater. Today Energy, 2021, 20, 100659 CrossRef CAS
.
- L. Peng, L. Shang, T. Zhang and G. I. Waterhouse, Recent advances in the development of single-atom catalysts for oxygen electrocatalysis and zinc–air batteries, Adv. Energy Mater., 2020, 10, 2003018 CrossRef CAS
.
- Q. Zhu, Y. Qu, D. Liu, K. W. Ng and H. Pan, Two-dimensional layered materials: high-efficient electrocatalysts for hydrogen evolution reaction, ACS Appl. Nano Mater., 2020, 3, 6270–6296 CrossRef CAS
.
- M. Luo, W. Sun, B. B. Xu, H. Pan and Y. Jiang, Interface engineering of air electrocatalysts for rechargeable zinc–air batteries, Adv. Energy Mater., 2021, 11, 2002762 CrossRef CAS
.
- E. J. Coleman, M. H. Chowdhury and A. C. Co, Insights into the oxygen reduction reaction activity of Pt/C and PtCu/C catalysts, ACS Catal., 2015, 5, 1245–1253 CrossRef CAS
.
- C. Wang, L. Jin, H. Shang, H. Xu, Y. Shiraishi and Y. Du, Advances in engineering RuO2 electrocatalysts towards oxygen evolution reaction, Chin. Chem. Lett., 2021, 32, 2108–2116 CrossRef CAS
.
- C. Ma, W. Sun, W. Qamar Zaman, Z. Zhou, H. Zhang, Q. Shen, L. Cao and J. Yang, Lanthanides regulated the amorphization-crystallization of IrO2 for outstanding OER performance, ACS Appl. Mater. Interfaces, 2020, 12, 34980–34989 CrossRef CAS PubMed
.
- M. Wu, X. Wu, Z. Wang, B. Hu, H. Guo, B. Zhang and L. Wang, Direct thermal annealing synthesis of FeO nanodots anchored on N-doped carbon nanosheet for long-term electrocatalytic oxygen reduction, Electrochim. Acta, 2021, 398, 139361 CrossRef CAS
.
- S. Guo, S. Zhang, L. Wu and S. Sun, Co/CoO nanoparticles assembled on graphene for electrochemical reduction of oxygen, Angew. Chem., Int. Ed., 2012, 51, 11770–11773 CrossRef CAS PubMed
.
- J. Huang, N. Zhu, T. Yang, T. Zhang, P. Wu and Z. Dang, Nickel oxide and carbon nanotube composite (NiO/CNT) as a novel cathode non-precious metal catalyst in microbial fuel cells, Biosens. Bioelectron., 2015, 72, 332–339 CrossRef CAS PubMed
.
- K. Min, S. Kim, E. Lee, G. Yoo, H. C. Ham, S. E. Shim, D. Lim and S.-H. Baeck, A hierarchical Co3O4/CoS microbox heterostructure as a highly efficient bifunctional electrocatalyst for rechargeable Zn–air batteries, J. Mater. Chem. A, 2021, 9, 17344–17352 RSC
.
- Z. Chen, R. Liu, S. Liu, J. Huang, L. Chen, R. Nadimicherla, D. Wu and R. Fu, FeS/FeNC decorated N,S-co-doped porous carbon for enhanced ORR activity in alkaline media, Chem. Commun., 2020, 56, 12921–12924 RSC
.
- T. Huang, H. Fang, L. Xu, Z. Wang, Y. Xin, J. Yu, S. Yao and Z. Zhang, Electrocatalytic performance of cubic NiS2 and hexagonal NiS for oxygen reduction reaction, J. Catal., 2018, 359, 223–232 CrossRef CAS
.
- Y. Lin, L. Yang, Y. Zhang, H. Jiang, Z. Xiao, C. Wu, G. Zhang, J. Jiang and L. Song, Defective carbon–CoP nanoparticles hybrids with interfacial charges polarization for efficient bifunctional oxygen electrocatalysis, Adv. Energy Mater., 2018, 8, 1703623 CrossRef
.
- B. Zhou, F. Yan, X. Li, J. Zhou and W. Zhang, An interpenetrating porous organic polymer as a precursor for FeP/Fe2P-embedded porous carbon toward a pH-universal ORR catalyst, ChemSusChem, 2019, 12, 915–923 CrossRef CAS PubMed
.
- X. Zhang, F. Yang, S. Sun, K. Wei, H. Liu, G. Li, Y. Sun, X. Li, J. Qian and S. Du, Boosting oxygen reduction via MnP nanoparticles encapsulated by N,P-doped carbon to Mn single atoms sites for Zn–air batteries, J. Colloid Interface Sci., 2024, 657, 240–249 CrossRef CAS PubMed
.
- M. E. Kreider, A. Gallo, S. Back, Y. Liu, S. Siahrostami, D. Nordlund, R. Sinclair, J. K. Nørskov, L. A. King and T. F. Jaramillo, Precious metal-free nickel nitride catalyst for the oxygen reduction reaction, ACS Appl. Mater. Interfaces, 2019, 11, 26863–26871 CrossRef CAS PubMed
.
- A. Miura, C. Rosero-Navarro, Y. Masubuchi, M. Higuchi, S. Kikkawa and K. Tadanaga, Nitrogen-Rich Manganese Oxynitrides with Enhanced Catalytic Activity in the Oxygen Reduction Reaction, Angew. Chem., Int. Ed., 2016, 55, 7963–7967 CrossRef CAS PubMed
.
- H. Abroshan, P. Bothra, S. Back, A. Kulkarni, J. K. Nørskov and S. Siahrostami, Ultrathin cobalt oxide overlayer promotes catalytic activity of cobalt nitride for the oxygen reduction reaction, J. Phys. Chem. C, 2018, 122, 4783–4791 CrossRef CAS
.
- H. Wang, C. Sun, Y. Cao, J. Zhu, Y. Chen, J. Guo, J. Zhao, Y. Sun and G. Zou, Molybdenum carbide nanoparticles embedded in nitrogen-doped porous carbon nanofibers as a dual catalyst for hydrogen evolution and oxygen reduction reactions, Carbon, 2017, 114, 628–634 CrossRef CAS
.
- C. Wang, S. Zhang, M. Zheng, R. Shu, S. Gu, J. Guo, C. Liu, J. Tang, J. Chen and X. Wang, Atomically thin titanium carbide used as high-efficient, low-cost and stable catalyst for oxygen reduction reaction, Int. J. Hydrogen Energy, 2020, 45, 6994–7004 CrossRef CAS
.
- W. Li, I. S. Amiinu, B. Zhang, C. Zhang, Z. Zhang, J. Zhu, J. Liu, Z. Pu, Z. Kou and S. Mu, Distorted niobium-self-doped graphene in situ grown from 2D niobium carbide for catalyzing oxygen reduction, Carbon, 2018, 139, 1144–1151 CrossRef CAS
.
- D. Xiang, X. Bo, X. Gao, C. Zhang, C. Du, F. Zheng, Z. Zhuang, P. Li, L. Zhu and W. Chen, Novel one-step synthesis of core@shell iron–nickel alloy nanoparticles coated by carbon layers for efficient oxygen evolution reaction electrocatalysis, J. Power Sources, 2019, 438, 226988 CrossRef CAS
.
- L. An, N. Jiang, B. Li, S. Hua, Y. Fu, J. Liu, W. Hao, D. Xia and Z. Sun, A highly active and durable iron/cobalt alloy catalyst encapsulated in N-doped graphitic carbon nanotubes for oxygen reduction reaction by a nanofibrous dicyandiamide template, J. Mater. Chem. A, 2018, 6, 5962–5970 RSC
.
- Z. Li, H. He, H. Cao, S. Sun, W. Diao, D. Gao, P. Lu, S. Zhang, Z. Guo and M. Li, Atomic Co/Ni dual sites and Co/Ni alloy nanoparticles in N-doped porous Janus-like carbon frameworks for bifunctional oxygen electrocatalysis, Appl. Catal., B, 2019, 240, 112–121 CrossRef CAS
.
- J. J. Barrios Cossio, P. Acevedo Pena, A. Hernández-Gordillo, L. F. Desdin Garcia, A. R. Puente Santiago, L. Echegoyen and E. Reguera, In situ aniline-polymerized interfaces on GO-PVA nanoplatforms as bifunctional supercapacitors and pH-universal ORR electrodes, ACS Appl. Energy Mater., 2020, 3, 4727–4737 CrossRef CAS
.
- Z. Ma, H. Tian, G. Meng, L. Peng, Y. Chen, C. Chen, Z. Chang, X. Cui, L. Wang and W. Jiang, Size effects of platinum particles@CNT on HER and ORR performance, Sci. China: Mater., 2020, 63, 2517–2529 CAS
.
- R. Du, N. Zhang, J. Zhu, Y. Wang, C. Xu, Y. Hu, N. Mao, H. Xu, W. Duan and L. Zhuang, Nitrogen-Doped Carbon Nanotube Aerogels for High-Performance ORR Catalysts, Small, 2015, 11, 3903–3908 CrossRef CAS PubMed
.
- J. Zhang, H. Yang and B. Liu, Coordination engineering of single-atom catalysts for the oxygen reduction reaction: a review, Adv. Energy Mater., 2021, 11, 2002473 CrossRef CAS
.
- W. Zhao, X. Ma, Y. Zheng, L. Yue, C. Xu, G. Wang, Y. Luo, D. Zheng, S. Sun and A. A. Alshehri, Hierarchical wormlike engineering: self-assembled SnS2 nanoflake arrays decorated on hexagonal FeS2@C nano-spindles enables stable and fast sodium storage, Chem. Eng. J., 2023, 459, 141629 CrossRef CAS
.
- X. Shu, M. Yang, D. Tan, K. San Hui, K. N. Hui and J. Zhang, Recent advances in the field of carbon-based cathode electrocatalysts for Zn–air batteries, Mater. Adv., 2021, 2, 96–114 RSC
.
- R. Tong, K. W. Ng, X. Wang, S. Wang, X. Wang and H. Pan, Two-dimensional materials as novel co-catalysts for efficient solar-driven hydrogen production, J. Mater. Chem. A, 2020, 8, 23202–23230 RSC
.
- B. Hu, A. Huang, X. Zhang, Z. Chen, R. Tu, W. Zhu, Z. Zhuang, C. Chen, Q. Peng and Y. Li, Atomic Co/Ni dual sites with N/P-coordination as bifunctional oxygen electrocatalyst for rechargeable zinc–air
batteries, Nano Res., 2021, 14, 3482–3488 CrossRef CAS
.
- W. Zhao, X. Ma, X. Wang, L. Yue, X. He, Y. Luo, D. Zheng, Y. Zheng, S. Sun and J. Zhang, A natural juncus-derived three-dimensional interconnected tubular carbon network decorated with tiny solid-solution metal sulfide nanoparticles achieves efficient sodium storage, J. Mater. Chem. A, 2023, 11, 2431–2442 RSC
.
- F. Dong, M. Wu, Z. Chen, X. Liu, G. Zhang, J. Qiao and S. Sun, Atomically dispersed transition metal–nitrogen–carbon bifunctional oxygen electrocatalysts for zinc–air batteries: recent advances and future perspectives, Nano-Micro Lett., 2022, 14, 36 CrossRef CAS PubMed
.
- X. Liu, G. Zhang, L. Wang and H. Fu, Structural design strategy and active site regulation of high-efficient bifunctional oxygen reaction electrocatalysts for Zn–air battery, Small, 2021, 17, 2006766 CrossRef CAS PubMed
.
- M. Yang, X. Shu, W. Pan and J. Zhang, Toward Flexible Zinc–Air Batteries with Self-Supported Air Electrodes, Small, 2021, 17, 2006773 CrossRef CAS PubMed
.
- F. Gao, J. He, H. Wang, J. Lin, R. Chen, K. Yi, F. Huang, Z. Lin and M. Wang, Te-mediated electro-driven oxygen evolution reaction, Nano Res. Energy, 2022, 1, 9120029 CrossRef
.
- H. Liu, W. Xie, Z. Huang, C. Yao, Y. Han and W. Huang, Recent Advances in Flexible Zn–Air Batteries: Materials for Electrodes and Electrolytes, Small Methods, 2022, 6, 2101116 CrossRef CAS PubMed
.
- X. Yan, Y. Ha and R. Wu, Binder-Free Air Electrodes for Rechargeable Zinc–Air Batteries: Recent Progress and Future Perspectives, Small Methods, 2021, 5, 2000827 CrossRef CAS PubMed
.
- X. Yang, H. Yang, J. Zhu, X. Cheng, Z. Chen, K. Zhang and C. Wang, Preparation and properties of flexible integrated cathode and electrolyte all-gel zinc–air batteries, J. Power Sources, 2023, 586, 233640 CrossRef CAS
.
- L. Yue, C. Ma, S. Yan, Z. Wu, W. Zhao, Q. Liu, Y. Luo, B. Zhong, F. Zhang and Y. Liu, Improving the intrinsic electronic conductivity of NiMoO4 anodes by phosphorous doping for high lithium storage, Nano Res., 2021, 15, 186–194 CrossRef
.
- M. Wu, G. Zhang, W. Wang, H. Yang, D. Rawach, M. Chen and S. Sun, Electronic Metal-Support Interaction Modulation of Single-Atom Electrocatalysts for Rechargeable Zinc–Air Batteries, Small Methods, 2022, 6, 2100947 CrossRef CAS PubMed
.
- Q. Wang, L. Shang, D. Sun-Waterhouse, T. Zhang and G. Waterhouse, Engineering local coordination environments and site densities for high-performance Fe–N–C oxygen reduction reaction electrocatalysis, SmartMat, 2021, 2, 154–175 CrossRef CAS
.
- Q. Wang, L. Shang, D. Sun-Waterhouse, T. Zhang and G. Waterhouse, Engineering local coordination environments and site densities for high-performance Fe–N–C oxygen reduction reaction electrocatalysis, SmartMat, 2021, 2, 154–175 CrossRef CAS
.
- Q. An, S. Bo, J. Jiang, C. Gong, H. Su, W. Cheng and Q. Liu, Atomic-Level Interface Engineering for Boosting Oxygen Electrocatalysis Performance of Single-Atom Catalysts: From Metal Active Center to the First Coordination Sphere, Adv. Sci., 2023, 10, 2205031 CrossRef CAS PubMed
.
- M. Karimpour, M. Dorri, A. Babaei, C. Zamani, M. Soleimani and M. Pourfath, New insights on the effects of surface facets and calcination temperature on electrochemical properties of MnCo2O4 as bifunctional oxygen electrocatalyst in zinc–air batteries, Electrochim. Acta, 2023, 467, 143022 CrossRef CAS
.
- D. H. Taffa, D. Balkenhohl, M. Amiri and M. Wark, Minireview: Ni–Fe and Ni–Co Metal–Organic Frameworks for Electrocatalytic Water-Splitting Reactions, Small Struct., 2023, 4, 2200263 CrossRef CAS
.
- S. Zhang, D. Lv, A. Zhou, D. Wang and D. Cai, Facile construction of CoNi-MOF nanosheet arrays and exploration of the active sites for oxygen evolution reaction with ultralow overpotential, J. Alloys Compd., 2023, 943, 169091 CrossRef CAS
.
- Y. Chen, S. He and Q. Rong, Stretchable, anti-drying, and self-healing hydrogel electrolytes for thermal adaptive zinc–air batteries with robust electrolyte/electrode interfaces, Mater. Today Chem., 2023, 33, 101726 CrossRef CAS
.
- K. W. Leong, Y. Wang, M. Ni, W. Pan, S. Luo and D. Y. Leung, Rechargeable Zn–air batteries: recent trends and future perspectives, Renewable Sustainable Energy Rev., 2022, 154, 111771 CrossRef CAS
.
- Y. Zhang, J. Wang, M. Alfred, P. Lv, F. Huang, Y. Cai, H. Qiao and Q. Wei, Recent advances of micro-nanofiber materials for rechargeable zinc–air batteries, Energy Storage Mater., 2022, 51, 181–211 CrossRef
.
- Z. Wang, D. Deng, H. Wang, S. Wu, L. Zhu, L. Xu and H. Li, Engineering Mn–Nx sites on porous carbon via molecular assembly strategy for long-life zinc–air batteries, J. Colloid Interface Sci., 2024, 653, 1348–1357 CrossRef CAS PubMed
.
- H. Liu, F. Yu, K. Wu, G. Xu, C. Wu, H. K. Liu and S. X. Dou, Recent progress on Fe-based single/dual-atom catalysts for Zn–air batteries, Small, 2022, 18, 2106635 CrossRef CAS PubMed
.
- Y. Deng, J. Du, Y. Zhu, L. Zhao, H. Wang, Y. Gong, J. Jin, B. He and R. Wang, Interface engineering of Ruddlesden–Popper perovskite/CeO2/carbon heterojunction for rechargeable zinc–air batteries, J. Colloid Interface Sci., 2024, 653, 1775–1784 CrossRef CAS PubMed
.
- W. Zhang, W. Pu, Y. Qu, H. Yang and Y. Liu, Facile synthesis of ultrathin SN co-doped carbon nanosheet as ORR electrocatalysts for application in sustainable zinc–air battery, Electrochim. Acta, 2023, 462, 142800 CrossRef CAS
.
- W. Wang, J. Gong, Q. Long, H. Wang, J. Huang, W. Dang, L. Chen, G. Li, Z. Hou and W. Xu, Fe–Fe3N composite nitrogen-doped carbon framework: multi-dimensional cross-linked structure boosting performance for the oxygen reduction reaction electrocatalysis and zinc–air batteries, Appl. Surf. Sci., 2023, 639, 158218 CrossRef CAS
.
- Q. Zeng, N. Deng, G. Wang, Y. Feng, W. Kang and B. Cheng, In situ growth of surface-reconstructed aluminum fluoride nanoparticles on N, F codoped hierarchical porous carbon nanofibers as efficient ORR/OER bifunctional electrocatalysts for rechargeable zinc–air batteries, J. Colloid Interface Sci., 2024, 654, 1063–1079 CrossRef CAS
.
- W. Chen, Y. Wang, W. Li, R. Liu, H. Zhang and Z. Zhang, Activity promotion of bifunctional catalysis by Se-doping and the application in rechargeable zinc–air batteries, Electrochim. Acta, 2023, 462, 142665 CrossRef CAS
.
- K. Li, X. Yang, J. Gao, Y. Guo, Y. Wang, Y. Zou, Z. Li and Z. Lü, Modulation of perovskite electronic configuration by cobalt doping for efficient catalysts in zinc–air battery electrodes, Chem. Eng. J., 2023, 475, 146301 CrossRef CAS
.
- S. Zhao, T. Liu, Y. Zuo, M. Wei, J. Wang, Z. Shao, D. Y. Leung, T. Zhao and M. Ni, High-Power-Density and High-Energy-Efficiency Zinc–Air Flow Battery System for Long-Duration Energy Storage, Chem. Eng. J., 2023, 470, 144091 CrossRef CAS
.
- M. Li, J. Shi, B. Xu, X. Yang, F. Gao, X. Zheng, Y. Liu, F. Cao, X. Guo and J. Zhang, Size-controlled Co/CoO heterogeneous nanoparticles confined in N-doped mesoporous carbon for efficient oxygen reduction in zinc–air batteries, J. Colloid Interface Sci., 2024, 653, 1317–1325 CrossRef CAS PubMed
.
- X. Ao, L. Li, S. Y. Yun, Y. Deng, W. Yoon, P. Wang, X. Jin, L. Dai, C. Wang and S.-J. Hwang, Highly accessible dual-metal atomic pairs for enhancing oxygen redox reaction in zinc–air batteries, Nano Energy, 2023, 118, 108952 CrossRef CAS
.
- X. Yang, H. Yang, J. Zhu, X. Cheng, Z. Chen, K. Zhang and C. Wang, Preparation and properties of flexible integrated cathode and electrolyte all-gel zinc–air batteries, J. Power Sources, 2023, 586, 233640 CrossRef CAS
.
- M. Karimpour, M. Dorri, A. Babaei, C. Zamani, M. Soleimani and M. Pourfath, New insights on the effects of surface facets and calcination temperature on electrochemical properties of MnCo2O4 as bifunctional oxygen electrocatalyst in zinc–air batteries, Electrochim. Acta, 2023, 467, 143022 CrossRef CAS
.
- Q. Liu, Q. Shi, Y. Ma, Z. Fang, Z. Zhou, G. Shao, H. Liu and W. Yang, ZIF-derived two-dimensional Co@Carbon hybrid: toward highly efficient trifunctional electrocatalysts, Chem. Eng. J., 2021, 423, 130313 CrossRef CAS
.
- C. Guan, A. Sumboja, W. Zang, Y. Qian, H. Zhang, X. Liu, Z. Liu, D. Zhao, S. J. Pennycook and J. Wang, Decorating Co/CoNx nanoparticles in nitrogen-doped carbon nanoarrays for flexible and rechargeable zinc–air batteries, Energy Storage Mater., 2019, 16, 243–250 CrossRef
.
- D. Chen, J. Yu, Z. Cui, Q. Zhang, X. Chen, J. Sui, H. Dong, L. Yu and L. Dong, Hierarchical architecture derived from two-dimensional zeolitic imidazolate frameworks as an efficient metal-based bifunctional oxygen electrocatalyst for rechargeable Zn–air batteries, Electrochim. Acta, 2020, 331, 135394 CrossRef CAS
.
- J. Zou, H. Dong, H. Wu, J. Huang, X. Zeng, Y. Dou, Y. Yao and Z. Li, Laser-induced rapid construction of Co/N-doped honeycomb-like carbon networks as oxygen electrocatalyst used in zinc–air batteries, Carbon, 2022, 200, 462–471 CrossRef CAS
.
- J. Sun, P. Leng, Y. Xie, X. Yu, K. Qu, L. Feng, H. Bao, F. Luo and Z. Yang, Co single atoms and Co nanoparticle relay electrocatalyst for rechargeable zinc air batteries, Appl. Catal., B, 2022, 319, 121905 CrossRef CAS
.
- Z. Zhang, Y. Tan, T. Zeng, L. Yu, R. Chen, N. Cheng, S. Mu and X. Sun, Tuning the dual-active sites of ZIF-67 derived porous nanomaterials for boosting oxygen catalysis and rechargeable Zn–air batteries, Nano Res., 2021, 14, 2353–2362 CrossRef CAS
.
- X. Wang, Z. Liao, Y. Fu, C. Neumann, A. Turchanin, G. Nam, E. Zschech, J. Cho, J. Zhang and X. Feng, Confined growth of porous nitrogen-doped cobalt oxide nanoarrays as bifunctional oxygen electrocatalysts for rechargeable zinc–air batteries, Energy Storage Mater., 2020, 26, 157–164 CrossRef
.
- L. Huang, L. Zuo, T. Yu, H. Wang, Z. He, H. Zhou, S. Su and T. Bian, Two-dimensional Co/Co9S8 nanoparticles decorated N, S dual-doped carbon composite as an efficient electrocatalyst for zinc–air battery, J. Alloys Compd., 2022, 897, 163108 CrossRef CAS
.
- A. Pendashteh, S. M. Vilela, I. Krivtsov, D. Ávila-Brande, J. Palma, P. Horcajada and R. Marcilla, Bimetal zeolitic imidazolate framework (ZIF-9) derived nitrogen-doped porous carbon as efficient oxygen electrocatalysts for rechargeable Zn–air batteries, J. Power Sources, 2019, 427, 299–308 CrossRef CAS
.
- L. Wei, L. Qiu, Y. Liu, J. Zhang, D. Yuan and L. Wang, Mn-doped Co–N–C dodecahedron as a bifunctional electrocatalyst for highly efficient Zn–air batteries, ACS Sustainable Chem. Eng., 2019, 7, 14180–14188 CrossRef CAS
.
- B. Lv, S. Zeng, W. Yang, J. Qiao, C. Zhang, C. Zhu, M. Chen, J. Di and Q. Li, In-situ embedding zeolitic imidazolate framework derived Co–N–C bifunctional catalysts in carbon nanotube networks for flexible Zn–air batteries, J. Energy Chem., 2019, 38, 1170–1176 Search PubMed
.
- C. Gu, J. Li, J.-P. Liu, H. Wang, Y. Peng and C.-S. Liu, Conferring supramolecular guanosine gel nanofiber with ZIF-67 for high-performance oxygen reduction catalysis in rechargeable zinc–air batteries, Appl. Catal., B, 2021, 286, 119888 CrossRef CAS
.
- K. Huang, Z. Yang, L. Liu, X. Yang, Y. Fan, M. Sun, C. Liu, W. Chen and J. Yang, A dual ligand coordination strategy for synthesizing drum-like Co, N co-doped porous carbon electrocatalyst towards superior oxygen reduction and zinc–air batteries, Int. J. Hydrogen Energy, 2021, 46, 24472–24483 CrossRef CAS
.
- F. Zhang, L. Chen, H. Yang, Y. Zhang, Y. Peng, X. Luo, A. Ahmad, N. Ramzan, Y. Xu and Y. Shi, Ultrafine Co nanoislands grafted on tailored interpenetrating N-doped carbon nanoleaves: an efficient bifunctional electrocatalyst for rechargeable Zn–air batteries, Chem. Eng. J., 2022, 431, 133734 CrossRef CAS
.
- H. Lei, S. Yang, Q. Wan, L. Ma, M. S. Javed, S. Tan, Z. Wang and W. Mai, Coordination and interface engineering to boost catalytic property of two-dimensional ZIFs for wearable Zn–air batteries, J. Energy Chem., 2022, 68, 78–86 CrossRef CAS
.
- H. Ge, G. Li, J. Shen, W. Ma, X. Meng and L. Xu, Co4N nanoparticles encapsulated in N-doped carbon box as tri-functional catalyst for Zn–air battery and overall water splitting, Appl. Catal., B, 2020, 275, 119104 CrossRef CAS
.
- J. Li, Y. Kang, D. Liu, Z. Lei and P. Liu, Nitrogen-doped graphitic carbon-supported ultrafine Co nanoparticles as an efficient multifunctional electrocatalyst for HER and rechargeable Zn–air batteries, ACS Appl. Mater. Interfaces, 2020, 12, 5717–5729 CrossRef CAS PubMed
.
- B. Zhu, J. Li, Z. Hou, C. Meng, G. Liu, X. Du and Y. Guan, MOF-derived nitrogen-doped carbon-based trimetallic bifunctional catalysts for rechargeable zinc–air batteries, Nanotechnology, 2022, 33, 405403 CrossRef CAS PubMed
.
- Z. Du, P. Yu, L. Wang, C. Tian, X. Liu, G. Zhang and H. Fu, Cubic imidazolate frameworks-derived CoFe alloy nanoparticles-embedded N-doped graphitic carbon for discharging reaction of Zn–air battery, Sci. China: Mater., 2020, 63, 327–338 CAS
.
- Z. Lei, Y. Tan, Z. Zhang, W. Wu, N. Cheng, R. Chen, S. Mu and X. Sun, Defects enriched hollow porous Co–N-doped carbons embedded with ultrafine CoFe/Co nanoparticles as bifunctional oxygen electrocatalyst for rechargeable flexible solid zinc–air batteries, Nano Res., 2021, 14, 868–878 CrossRef CAS
.
- M. Huo, B. Wang, C. Zhang, S. Ding, H. Yuan, Z. Liang, J. Qi, M. Chen, Y. Xu and W. Zhang, 2D Metal–Organic Framework Derived CuCo Alloy Nanoparticles Encapsulated by Nitrogen-Doped Carbonaceous Nanoleaves for Efficient Bifunctional Oxygen Electrocatalyst and Zinc–Air Batteries, Chem. – Eur. J., 2019, 25, 2780–12788 CrossRef PubMed
.
- X. Liu, Z. Yin, M. Cui, L. Gao, A. Liu, W.-N. Su, S. Chen, T. Ma and Y. Li, Double shelled hollow CoS2@MoS2@NiS2 polyhedron as advanced trifunctional electrocatalyst for zinc–air battery and self-powered overall water splitting, J. Colloid Interface Sci., 2022, 610, 653–662 CrossRef CAS PubMed
.
- K. Huang, W. Zhang, R. Devasenathipathy, Z. Yang, X. Zhang, X. Wang, D.-H. Chen, Y. Fan and W. Chen, Co nanoparticles and ZnS decorated N, S co-doped carbon nanotubes as an efficient oxygen reduction catalyst in zinc–air batteries, Int. J. Hydrogen Energy, 2021, 46, 30090–30100 CrossRef CAS
.
- H. Pan, C. Zhang, Z. Lu, J. Dou, X. Huang, J. Yu, J. Wu, H. Li and X. Chen, Self-standing electrospun Co/Zn@N-doped carbon nanofiber electrode for highly stable liquid and solid-state rechargeable zinc–air batteries and performance evaluated by scanning electrochemical microscopy at various temperatures, Chem. Eng. J., 2023, 477, 147022 CrossRef CAS
.
- W. Zhang, J. Chu, S. Li, Y. Li and L. Li, CoNxC active sites-rich three-dimensional porous carbon nanofibers network derived from bacterial cellulose and bimetal-ZIFs as efficient multifunctional electrocatalyst for rechargeable Zn–air batteries, J. Energy Chem., 2020, 51, 323–332 CrossRef
.
- Y. Guan, Y. Li, S. Luo, X. Ren, L. Deng, L. Sun, H. Mi, P. Zhang and J. Liu, Rational design of positive-hexagon-shaped two-dimensional ZIF-derived materials as improved bifunctional oxygen electrocatalysts for use as long-lasting rechargeable Zn–Air batteries, Appl. Catal., B, 2019, 256, 117871 CrossRef CAS
.
- C.-C. Wang, K.-Y. Hung, T.-E. Ko, S. Hosseini and Y.-Y. Li, Carbon-nanotube-grafted and nano-Co3O4-doped porous carbon derived from metal–organic framework as an excellent bifunctional catalyst for zinc–air battery, J. Power Sources, 2020, 452, 227841 CrossRef CAS
.
- Y. Zhao, X. Wang, X. Guo, N. Shi, D. Cheng, H. Zhou, N. Saito and T. Fan, Nitrogen-doped 3D porous graphene coupled with densely distributed CoOx nanoparticles for efficient multifunctional electrocatalysis and Zn–Air battery, Electrochim. Acta, 2022, 420, 140432 CrossRef CAS
.
- L. Li, N. Li, J. Xia, S. Zhou, X. Qian, F. Yin, G. He and H. Chen, Metal–organic framework-derived Co single atoms anchored on N-doped hierarchically porous carbon as a pH-universal ORR electrocatalyst for Zn–air batteries, J. Mater. Chem. A, 2023, 11, 2291–2301 RSC
.
- Y. Hao, Y. Kang, H. Kang, H. Xin, F. Liu, L. Li, W. Wang and Z. Lei, Self-grown layered double hydroxide nanosheets on bimetal–organic frameworks-derived N-doped CoOx carbon polyhedra for flexible all-solid-state rechargeable Zn–air batteries, J. Power Sources, 2022, 524, 231076 CrossRef CAS
.
- Y. Zhu, Z. Zhang, Z. Lei, Y. Tan, W. Wu, S. Mu and N. Cheng, Defect-enriched hollow porous Co–N-doped carbon for oxygen reduction reaction and Zn–Air batteries, Carbon, 2020, 167, 188–195 CrossRef CAS
.
- K. Li, Y. Zhang, P. Wang, X. Long, L. Zheng, G. Liu, X. He and J. Qiu, Core–Shell ZIF-67@ZIF-8-derived multi-dimensional cobalt-nitrogen doped hierarchical carbon nanomaterial for efficient oxygen reduction reaction, J. Alloys Compd., 2022, 903, 163701 CrossRef CAS
.
- Y.-T. Zhang, S.-Y. Li, N.-N. Zhang, G. Lin, R.-Q. Wang, M.-N. Yang and K.-K. Li, A carbon catalyst
doped with Co and N derived from the metal–organic framework hybrid (ZIF-8@ZIF-67) for efficient oxygen reduction reaction, New Carbon Mater., 2023, 38, 200–209 CrossRef CAS
.
- H. Park, S. Oh, S. Lee, S. Choi and M. Oh, Cobalt-and nitrogen-codoped porous carbon catalyst made from core–shell type hybrid metal–organic framework (ZIF-L@ZIF-67) and its efficient oxygen reduction reaction (ORR) activity, Appl. Catal., B, 2019, 246, 322–329 CrossRef CAS
.
- G. Li, W. Deng, L. He, J. Wu, J. Liu, T. Wu, Y. Wang and X. Wang, Zn, Co, and Fe tridoped N–C core–shell nanocages as the high-efficiency oxygen reduction reaction electrocatalyst in zinc–air batteries, ACS Appl. Mater. Interfaces, 2021, 13, 28324–28333 CrossRef CAS PubMed
.
- Y. Peng, F. Zhang, Y. Zhang, X. Luo, L. Chen, Y. Shi and N. S-doped, hollow carbon nanosheet-encapsulated Co9S8 nanoparticles as a highly efficient bifunctional electrocatalyst for rechargeable zinc–air batteries, Dalton Trans., 2022, 51, 12630–12640 RSC
.
- H. Qi, Y. Feng, Z. Chi, Y. Cui, M. Wang, J. Liu, Z. Guo, L. Wang and S. Feng, In situ encapsulation of Co-based nanoparticles into nitrogen-doped carbon nanotubes-modified reduced graphene oxide as an air cathode for high-performance Zn–air batteries, Nanoscale, 2019, 11, 21943–21952 RSC
.
- Y. Liu, Z. Li, S. Wang, J. Xuan, D. Xiong, L. Zhou, J. Zhou, J. Wang, Y. Yang and Y. Du, Hierarchical porous yolk-shell Co–NC nanocatalysts encaged ingraphene nanopockets for high-performance Zn–air battery, Nano Res., 2023, 16, 8893–8901 CrossRef CAS
.
- Y. Zhang, Q. He, Z. Chen, Y. Chi, J. Sun, D. Yuan and L. Zhang, Hierarchically porous Co@N-doped carbon fiber assembled by MOF-derived hollow polyhedrons enables effective electronic/mass transport: an advanced 1D oxygen reduction catalyst for Zn–air battery, J. Energy Chem., 2023, 76, 117–126 CrossRef CAS
.
- X. He, L. Chang, P. Han, K. Li, H. Wu, Y. Tang, F. Gao, Y. Zhang and A. Zhou, High-performance Co–NC catalyst derived from PS@ZIF-8@ZIF-67 for improved oxygen reduction reaction, Colloids Surf., A, 2023, 663, 130988 CrossRef CAS
.
- S. Ahmed, J. Shim, H.-J. Sun, H.-R. Rim, H.-K. Lee and G. Park, Nickel decorated bimetallic catalysts derived from metal–organic frameworks as cathode materials for rechargeable Zinc–Air batteries, Mater. Lett., 2021, 283, 128781 CrossRef CAS
.
- H. Ning, G. Li, Y. Chen, K. Zhang, Z. Gong, R. Nie, W. Hu and Q. Xia, Porous N-doped carbon-encapsulated CoNi alloy nanoparticles derived from MOFs as efficient bifunctional oxygen electrocatalysts, ACS Appl. Mater. Interfaces, 2018, 11, 1957–1968 CrossRef PubMed
.
- J. Li, H. Xue, N. Xu, X. Zhang, Y. Wang, R. He, H. Huang and J. Qiao, Co/Ni dual-metal embedded in heteroatom doped porous carbon core–shell bifunctional electrocatalyst for rechargeable Zn–air batteries, MRE, 2022, 2, 100090 CAS
.
- M. Jiang, J. Yang, J. Ju, W. Zhang, L. He, J. Zhang, C. Fu and B. Sun, Space-confined synthesis of CoNi nanoalloy in N-doped porous carbon frameworks as efficient oxygen reduction catalyst for neutral and alkaline aluminum-air batteries, Energy Storage Mater., 2020, 27, 96–108 CrossRef
.
- Y. Chong, Z. Pan, M. Su, X. Yang, D. Ye and Y. Qiu, 1D/2D hierarchical Co1−xFexO@N-doped carbon nanostructures for flexible zinc–air batteries, Electrochim. Acta, 2020, 363, 137264 CrossRef CAS
.
- J. Li, Y. Kang, W. Wei, X. Li, Z. Lei and P. Liu, Well-dispersed ultrafine CoFe nanoalloy decorated N-doped hollow carbon microspheres for rechargeable/flexible Zn–air batteries, Chem. Eng. J., 2021, 407, 127961 CrossRef CAS
.
- T. Liu, F. Yang, G. Cheng and W. Luo, Reduced Graphene Oxide-Wrapped Co9–xFexS8/Co,Fe–N–C Composite as Bifunctional Electrocatalyst for Oxygen Reduction and Evolution, Small, 2018, 14, 1703748 CrossRef PubMed
.
- L. Li, L. Fu, R. Wang, J. Sun, X. Li, C. Fu, L. Fang and W. Zhang, Cobalt, manganese zeolitic-imidazolate-framework-derived Co3O4/Mn3O4/CNx embedded in carbon nanofibers as an efficient bifunctional electrocatalyst for flexible Zn–air batteries, Electrochim. Acta, 2020, 344, 136145 CrossRef CAS
.
- Q. Zhou, L. Zhang and X. Wang, Bimetallic ZIFs derived 3D acetylene black loading La2O3/Co bifunctional ORR/OER catalysts, Appl. Surf. Sci., 2023, 610, 155551 CrossRef CAS
.
- Y. Sun, Y. Guan, X. Wu, W. Li, Y. Li, L. Sun, H. Mi, Q. Zhang, C. He and X. Ren, ZIF-derived “senbei”-like Co9S8/CeO2/Co heterostructural nitrogen-doped carbon nanosheets as bifunctional oxygen electrocatalysts for Zn–air batteries, Nanoscale, 2021, 13, 3227–3236 RSC
.
- J. Chen, J. Huang, R. Wang, W. Feng, H. Wang, T. Luo, Y. Hu, C. Yuan, L. Feng and L. Cao, Atomic ruthenium coordinated with chlorine and nitrogen as efficient and multifunctional electrocatalyst for overall water splitting and rechargeable zinc–air battery, Chem. Eng. J., 2022, 441, 136078 CrossRef CAS
.
- B. Zhong, L. Zhang, J. Yu and K. Fan, Ultrafine iron-cobalt nanoparticles embedded in nitrogen-doped porous carbon matrix for oxygen reduction reaction and zinc–air batteries, J. Colloid Interface Sci., 2019, 546, 113–121 CrossRef CAS PubMed
.
- L. Chen, Y. Zhang, L. Dong, W. Yang, X. Liu, L. Long, C. Liu, S. Dong and J. Jia, Synergistic effect between atomically dispersed Fe and Co metal sites for enhanced oxygen reduction reaction, J. Mater. Chem. A, 2020, 8, 4369–4375 RSC
.
- Q. H. Nguyen, K. Im and J. Kim, Synthesis of hollow Fe, Co, and N-doped carbon catalysts from conducting polymer-metal–organic-frameworks core–shell particles for their application in an oxygen reduction reaction, Int. J. Hydrogen Energy, 2022, 47, 24169–24178 CrossRef CAS
.
- L. Sun, Y. Qin, L. Fu, Y. Di, K. Hu, H. Li, L. Li and W. Zhang, A self-supported bifunctional air cathode composed of Co3O4/Fe2O3 nanoparticles embedded in nanosheet arrays grafted onto carbon nanofibers for secondary zinc–air batteries, J. Alloys Compd., 2022, 921, 166128 CrossRef CAS
.
- C. Ye, P. Fan, D. Wei, J. Wang and L. Xu, Nitrogen-doped carbon fibers loaded with Co/Co2Mn3O8 alloy nanoparticles as bifunctional oxygen electrocatalysts for rechargeable zinc–air batteries, J. Alloys Compd., 2023, 936, 168210 CrossRef CAS
.
- J. Li, W. Li, H. Mi, Y. Li, L. Deng, Q. Zhang, C. He and X. Ren, Bifunctional oxygen electrocatalysis on ultra-thin Co9S8/MnS carbon nanosheets for all-solid-state zinc–air batteries, J. Mater. Chem. A, 2021, 9, 22635–22642 RSC
.
- J. Li, Y. Wang, Z. Yin, R. He, Y. Wang and J. Qiao, In situ growth of CoNi bimetal anchored on carbon nanoparticle/nanotube hybrid for boosting rechargeable Zn–air battery, J. Energy Chem., 2022, 66, 348–355 CrossRef CAS
.
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.