Chen Liu*a and
Huahuan Wangb
aCollege of New Energy and Materials, China University of Petroleum, Beijing, PR China. E-mail: 13176230761@163.com
bState Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing, PR China
First published on 28th July 2025
Lithium–oxygen batteries (LOBs) are highly esteemed for their exceptional energy density (∼3500 Wh kg−1) and are regarded as one of the most promising battery technologies. However, several challenges hinder the commercialization and widespread adoption of LOBs, including side reactions occurring at the lithium anode, electrolyte decomposition, and growth of lithium dendrites. These issues contribute to reduced cycle life and increased overpotential, adversely affecting the performance of LOBs. Consequently, exploring effective cathode catalysts is crucial for advancing this field. Initially, this review provides background information on LOBs, including that on singlet oxygen formation, and then clearly and succinctly outlines their operational mechanisms. Subsequently, a detailed analysis of recent developments in cathode catalysts for aprotic LOBs is presented, notably spin engineering and amorphization strategy for spinel oxides and development of high-entropy alloys. Finally, innovative directions are proposed, especially new soluble catalysts and machine learning, for investigating cathode catalysts and enhancing the electrochemical performance of LOBs considering existing challenges.
By contrast, lithium–oxygen batteries (LOBs) can overcome the energy density deficiency with their extraordinarily high theoretical energy density of ∼3500 Wh kg−1 during discharging (based on Li2O2). Therefore, LOBs have the potential to extend the mile range of EVs to more than 500 miles in an ideal scenario.2,3 LOBs have emerged as the new star in the field of batteries since they were first reported in 1996 by Abraham and Jiang.4 Thus far, the major LOBs can be categorized into three types according to the differences in their electrolytes, which are aprotic, aqueous, and solid-state.4–8 Among them, aprotic LOBs are the most popular and promising, which are the topic of this review. Through the persistent efforts of numerous scientists and researchers, a large variety of potential catalysts and electrolytes have been found and studied to facilitate the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER). Thus, it appears that LOBs are advancing toward the stage of commercialization.
In terms of the anode, the Li anode is highly reactive and tends to undergo side reactions with CO2 and H2O.5,10,11 Furthermore, the formation of Li dendrites, which is a common issue in all Li-based batteries, severely impairs the performance of Li–O2.12 Dendrites can lead to short circuit, dead Li, low coulombic efficiency (CE), more serious adverse reactions and increased polarization, which eventually result in safety hazards, shortened cycle life and reduced energy density.12,13
In the case of the cathode, the high charging overpotential (ηOER) and short cycle life caused by the accumulation of insulating and insoluble Li2O2 in the cathode are also fatal to aprotic LOBs.9,16,17 Furthermore, the rate of O2 diffusion through the cathode materials is crucial for the Li–O2 performance; thus, improving the O2 diffusion kinetics becomes another priority.18,19 In cathodes composed of carbon, they also encounter erosion caused by LiO2/Li2O2.106
In addition, nucleophiles, bases and free radicals including LiO2, HOO−, HOO˙, O2−, and even Li2O2 are likely to attack the electrolyte.14,15,173 The reactions of electrolyte decomposition are diverse, such as nucleophilic attack, auto-oxidation, acid/base reactions and proton-mediated reactions, which seriously hinder the performance of Li–O2 cells.165 The decomposition of the electrolyte such as DME can result in its electrodeposition on the cathode, producing CO2 during charging.167 Therefore, it is necessary to reduce the negative effects cause by reactive oxygen species (ROSs).
The relationship between cathodes with carbon and the electrolyte is inseparable, given that it has been reported that their decomposition results in a synergistic effect.166 Bruce et al. found that in DMSO or TEGDME, the decomposition of the electrolyte is the main side reaction, which is accelerated by the catalysis of carbon, especially hydrophilic carbon.16 Thus, to avoid CO2 emission from side reactions, Kang et al. used LiNO3 in DME as the electrolyte and its solvate on the surface of the cathode deactivated carbon.168 The erosion of the carbon cathode and decomposition of the electrolyte catalyzed by carbon were both effectively controlled. Thus, these results demand a proper match between the cathodes and electrolyte in future research.
Furthermore, it seems that traditional catalysts have negative catalytic effects in Li–O2 cells. McCloskey et al. found that traditional OER catalysts on carbon (Au/XC72, MnO2/XC72 and Pt/XC72) had a lower Coulombic efficiency (OER/ORR) than pure carbon (XC72) in pure DME because these catalysts catalyzed the decomposition of DME into CO2 more than the oxidation of LiO2/Li2O2.167
It is universally acknowledged that singlet oxygen (1Δg or 1O2), a type of ROS, exhibits greater reactivity with organic matter than other ROSs. It is widely reported that singlet oxygen also contributes to parasitic reactions in Li–O2 cells,10,11,169 and water can lead to the generation of more 1O2.170 Thus far, several pathways for the formation of 1O2 have been proposed, mainly including the disproportionation of the superoxide anion,172 the oxidation of the LiO2 superoxide intermediate above 3.5 V,169,173 and the electrochemical oxidation of Li2CO3.174
However, the latest studies indicate that singlet oxygen is not the main source of electrolyte decomposition, and its production depends on the redox potential of the redox mediators (RMs).175,176 Bruce et al. found that instead of 1O2, 3O2 is the main generated oxygen molecule when RMs are added to the electrolyte.176
In conclusion, 1O2 may contribute to carbon erosion and electrolyte decomposition, but is not the major cause of the latter.
Besides the above-mentioned negative impacts, singlet oxygen has been reported to decompose RMs.171 However, RMs play an important role in the decomposition of the discharge products (mainly Li2O2). The gradual deactivation of RMs causes Li2O2 to be left on the cathode surface, and then the cathode catalyst cannot work normally.
Thus, to address these challenges, it is necessary to develop new effective, durable, and cost-effective cathode catalysts. Drawing on the understanding of the working principles of aprotic LOBs, this review systematically analyses the recent advances in cathode catalyst materials and several key modulation strategies. Additionally, it proposes several novel potential directions for enhancing the electrochemical performance of LOBs.
However, the charging and discharging reactions taking place at the cathode were controversial to some extent in the early stage of research on Li–O2 batteries.
In the early stage, the surface electrochemical reactions of the surface route were thought to be as follows:5,21,22
![]() | (1) |
![]() | (2) |
![]() | (3) |
The forward reactions occur during discharging, while the backward reactions take place during charging. However, it was unknown whether the specific steps of OER were just the reverse to the ORR at that time.
None of eqn (1)–(3) are solution electrochemical or chemical processes. Although there is clear evidence that Li2O2 can barely dissolve in aprotic electrolytes,5 some authors23 thought that the absorbed LiO2 species can dissolve in TEGDME, and then nucleate, form Li2O2 crystallites and grow into toroidal Li2O2 at a relatively low current density. They thought that the following reactions happened in the solution route:
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
Then, Johnson et al. proposed a unified mechanism of ORR (Fig. 2a), which explains that the solubility of LiO2 determines the Li2O2 form through either the surface route or the solution route.204 After subsequent studies by other groups,205–207 this improved mechanism has been widely accepted.
![]() | ||
Fig. 2 (a) Schematic of the first proposed unified O2 reduction mechanism in an aprotic solvent containing Li+, showing the surface pathway followed when ΔG° ≫ 0 (low DN) and the solution pathway followed when ΔG° ≪ 0 (high DN). ΔG° refers to the following equation: ![]() ![]() |
The solubility of LiO2 is thought to be determined mainly by the donor number (DN) and acceptor number (AN) of the solvent, with the former being more dominant.205 High-DN and high-AN solvents have been proven favorable for the solution route.206,207
To meet practical use, Li–O2 batteries must run at a low discharge overpotential and a high current. Under a low discharge overpotential, in low-DN solvents such as acetonitrile (ACN) and/or low-AN solvents such as anhydrous DME, only film-like Li2O2 is formed by the surface route, as follows.205
![]() | (9) |
![]() | (10) |
![]() | (11) |
By contrast, in high-DN solvents such as DMSO and/or high-AN solvents such as water and CH3OH,207 toroidal Li2O2, as the major product, is formed by the solution route.205 However, a second reduction still happens in high-DN and/or high-AN solvents to some extent, possibly due to its fast reaction speed.204
Li+ + e− + O2 = Li+ + O2−(cathode) | (12) |
2 (Li+ + O2−) = Li2O2 + O2(cathode) | (13) |
2Li+ + e− + O2− = Li2O2(cathode) | (14) |
It is widely known that insoluble LiOH can be formed in the presence of an appropriate amount of water. As mentioned above, even trace water can prompt the formation of toroidal Li2O2, which increases the battery capacity. With an increase in the content of water, it was reported that the size of Li2O2 also increases. Li et al. observed LiOH in Li–O2 cells with 0.5 M LiClO4 in DMSO with 120 ppm H2O.208 They proved that Li2O2 transformed to LiOH through a chemical process. Based on the experimental results, the overall ORR and OER mechanism for Li–O2 systems with trace water was proposed (Fig. 2b).
However, no LiOH was detected in Li–O2 batteries with 1 M LiTFSI in DME with varying water contents (up to 4000 ppm).207 However, the formation of H2O2 was indirectly proven. It was explained that the reason why no LiOH was detected in the former study207 may be that after a small amount of water was consumed, the existing H2O2 could not transform into H2O without catalysts such as MnO2, which led to the reserve reaction of step (i), and then Li2O2 was the major discharge product.208
However, it seems that there is no unified ORR mechanism for water-containing Li–O2 systems. With an increase in the water content, different mechanisms have been put forward.209–211 For example, in Li–O2 systems with added LiI redox mediator and 45000 ppm H2O, their discharge and charge reactions are shown in Fig. 2c.211 In this case, LiOH is produced from LiO2 and H2O via the action of LiI.
Wang et al. proposed an OER mechanism that relates the decomposition of Li2O2 to the donicity (donor number, namely) of the solvent (Fig. 2d).212 In low-donicity solvents, owing to their low ability to dissolve LiO2, Li2O2 first decomposes into solid Li2−xO2, as follows:
Li2O2 → Li2−xO2 + xLi+ + xe−![]() | (15) |
Li2–xO2 → O2 + (2 − x)Li+ + (2 – x)e−![]() | (16) |
Li2O2 → Li2−xO2 + xLi+ + xe−![]() | (17) |
Li2−xO2 → Li+ + O2− + xLi+ + xe−![]() | (18) |
2(Li+ + O2−) → Li2O2 + O2![]() | (19) |
In the case of rechargeable aprotic Li–O2 batteries, improving the diffusion of O2 is important, which further depends on the number and structure of pores in the catalysts. Then, the adsorption energies for O2 or intermediates especially LiO2, which are subject to change depending on the electronic structure of the catalyst, determine the amount of deposited Li2O2 and its formation route. They can influence the ORR/OER overpotentials, reversibility, capacity, and cycling life of Li–O2 cells. Therefore, the electronic structure of catalysts needs much attention. It is obvious and accepted that more Li2O2 can deposit on the unit surface area of the catalyst when its specific surface area is larger, especially in the case of 2D and hierarchical catalysts, leading to a more favorable capacity.
Considering the commercialization of LOBs, they need to satisfy the electricity demand of humans, who are concerned about their useful lives and maximum capacities. In addition, to make LOBs more endurable, their ORR/OER overpotentials must be low.
Therefore, in nearly all modulation strategies, the optimization of the structure of the pores, electronic structure and/or specific surface area is the method applied to achieve an improvement in cycling capability, increase in maximum capacity, and/or decrease in the ORR/OER overpotentials of Li–O2 cells.
To make this clear, several vital metrics are listed and stated briefly.
1. The structure of pores
Due to the importance of appropriate size and number of pores to achieve O2 diffusion, the structure of the pores of several catalysts, especially carbon-based materials and MOF-based materials are given attention.
2. The electronic structure
Modulation strategies such as doping and vacancy engineering can change the electronic structure of materials, which can be proven by the peak shifts in XPS, changes in the adsorption energies for O2, intermediates and the discharge products, AC impedance measurements, outcomes of DFT calculation, etc.
3. Specific surface area
Regarding catalyst materials, especially carbon-based and MOF-based catalysts, a higher specific surface area leads to the greater deposition of discharge products, which results in a larger capacity in batteries.
4. Cycling capability
The cycling number of a catalyst within a specific voltage window and/or within a set coulombic efficiency shows its potential as a practical catalyst in the future. More cycles under the same testing conditions indicate that the battery is more durable.
5. Maximum capacity
The maximum capacity determines the working duration of Li–O2 cells until electricity runs out. A larger maximum capacity at the same rate means a longer working time for batteries.
6. ORR/OER overpotentials
Besides good cycling capability and a large maximum capacity, an ideal catalyst also needs to show low ORR and OER overpotentials. Typically, the OER and ORR overpotentials increase as cycling proceeds. Therefore, in this review, we summarize the OER and ORR overpotentials in the first cycle in Table 1.
The first OER potentiala (V) | The first ORR potentialb (V) | Maximum capacityc (mAh g−1)/rated (mA g−1) | Cycle number/cut-off capacitye (mAh g−1)/rated (mA g−1) | Electrolyte | Ref. | |
---|---|---|---|---|---|---|
Default Li–O2 battery testing atmosphere is pure oxygen unless there is a special comment. The data in Ref. 47, 201 were obtained in (simulated) air atmosphere.a The first OER/ORR potential is the potential at half of the cut-off capacity in the fifth column of the table. If there is more than one cut-off capacity, the first OER/ORR potentials correspond to the first cut-off capacity. If there is no cut-off capacity, the first OER/ORR potential is the potential at half of the maximum capacity of the first cycle.b The first OER/ORR potential is the potential at half of the cut-off capacity in the fifth column of the table. If there is more than one cut-off capacity, the first OER/ORR potentials correspond to the first cut-off capacity. If there is no cut-off capacity, the first OER/ORR potential is the potential at half of the maximum capacity of the first cycle.c The default units of the maximum capacity, rate, and cut-off capacity are in their respective rear brackets. The special cases are labelled particularly.d The default units of the maximum capacity, rate, and cut-off capacity are in their respective rear brackets. The special cases are labelled particularly.e The default units of the maximum capacity, rate, and cut-off capacity are in their respective rear brackets. The special cases are labelled particularly.f Fe-UiO-66 is a photocathode in a photo-assisted Li–O2 battery; all the data in the row are obtained with light illumination. | ||||||
Transition metal compounds (spinel oxides and composites) | ||||||
Co3O4 films/Ni foams | ∼3.5 | ∼2.5 | 2460/200 | 35/1000/200 | 1.0 M LiTFSI in TEGDME | 18 |
Co–Mn–O | ∼4.0 | ∼2.75 | 7653/0.04 mA cm−2 | 100/500/0.16 mA cm−2 | 1.0 M LiCF3SO3 in TEGDME | 27 |
ZnCo2O4 nanoflakes | ∼4.0 | ∼2.6 | — | 30/500/0.1 mA cm−2 | 1.0 M LiTFSI in TEGDME | 28 |
NiFeO-600 | ∼3.5 | ∼2.75 | 23![]() |
193/1000/1000 300/2000/500 | 1.0 M LiTFSI in TEGDME | 29 |
Ni-Co3O4-Vo | 4.06 | 2.65 | 5275/200 | 371/500/200337/500/500 | 1.0 M LiTFSI in TEGDME | 56 |
Co3−xO4, Co-300 | ∼4.0 | ∼2.7 | 14![]() |
70/1000/100 | 1.0 M LiTFSI in TEGDME | 62 |
Co3O4@ND-CN | 4.10 | 2.78 | 9838.8/100 | 80/1000/100 | 1.0 M LiTFSI in TEGDME | 63 |
Co3O4-HCNFs | ∼4.3 | ∼2.6 | 14![]() |
180/600/100 | 1.0 M LiCF3SO3 in TEGDME | 64 |
Co3O4 NF/GNF | ∼4.1 | ∼2.7 | 10![]() |
80/1000/200 | 1.0 M LiTFSI in TEGDME | 65 |
Pt-Co3O4 NWs@CP | ∼3.8 | ∼2.8 | 17079.4/100 | 75/500/100 | 1.5 M LiTFSI in TEGDME | 66 |
Co2N/Co3O4–Ti3C2Tx | ∼3.2 | ∼2.6 | 14![]() |
>300/500/500 | 1.0 M LiTFSI in TEGDME | 67 |
NiSA–Co3O4/CC | ∼3.3 | ∼2.5 | 21![]() |
128/1000/200 | 1.0 M LiCF3SO3 in TEGDME | 68 |
{112} faceted Co3O4 plate | ∼3.9 | ∼2.7 | 9144/100 | 45/500/− | 1.0 M LiCF3SO3 in TEGDME | 159 |
Co3O4-H | ∼3.5 | ∼2.65 | 5200/200 | >100/500/200 | 1.0 M LITFSI in TEGDME | 160 |
Mesoporous ZnCo2O4 nanoflakes | ∼4.0 | ∼2.6 | ∼1322/0.1 mA cm−2 | 30/500/0.1 mA cm−2 | 1.0 M LITFSI in TEGDME | 177 |
3DOM ZnCo2O4 | ∼3.9 | ∼2.75 | 6024/100 | 27/500/100 | 1.0 M LiCF3SO3 in TEGDME | 178 |
Porous ZnCo2O4 nanofibers | ∼4.25 | ∼2.75 | ∼12![]() |
226/1000/500 | 1.0 M LiNO3 in DMAc | 180 |
ZnCo2O4/CC | ∼3.8 | ∼2.7 | 4.23 mAh cm−2/0.12 mA cm−2 | 80/0.3 mAh cm−2/0.2 mA cm−2 | 1.0 M LiTFSI in TETRAGLYME | 181 |
NiCo2O4/CC | ∼3.65 | ∼2.75 | 6.2 mAh cm−2/0.12 mA cm−2 | 200/0.3 mAh cm−2/0.2 mA cm−2 | ||
CuCo2O4/CC | ∼3.6 | ∼2.7 | 4.76 mAh cm−2/0.12 mA cm−2 | 107/0.3 mAh cm−2/0.2 mA cm−2 | ||
FeCo2O4/CC | ∼3.85 | ∼2.7 | 4.75 mAh cm−2/0.12 mA cm−2 | 107/0.3 mAh cm−2/0.2 mA cm−2 | ||
Cu1–ZnCo2O4(Cu0.15Zn0.85Co2O4) | ∼3.8 | ∼2.65 | 12984.9/100 | 400/1000/200 | 1.0 M LiTFSI in TEGDME | 182 |
ZCO-650 | ∼3.46 | ∼2.85 | — | 800 h/1000/500 | 1.0 M LiClO4 in DMSO | 183 |
NiCo2O4/SP | ∼3.9 | ∼2.8 | >8300/100 | 128/1000/200 | 1.0 M LiTFSI in TEGDME | 184 |
NiCo2O4 microspheres | ∼4.0 | ∼2.9 | 3163/0.08 mA cm−2 | 60/500/0.24 mA cm−2 | 0.1 M LiClO4 in DMSO | 185 |
Au/NiCo2O4/3D-G | ∼3.9 | ∼2.8 | 1275/42.5 | 40/510/42.5 | 0.1 M LiClO4 in DME | 186 |
NiCo2O4 NS@Ni | ∼3.4 | ∼2.8 | 7004/40 | 100/500/100 | 0.5 M LiClO4 in DMSO | 187 |
Flower-like NCO/N-rGO | ∼4.0 | ∼2.7 | 15![]() |
50/1000/200 | LiTFSI in TEGDME | 188 |
CeO2@NiCo2O4 NWAs | ∼3.1 | ∼2.95 | ∼3250/100 | 64/500/100 | 1.0 M LiCF3SO3 in TEGDME | 189 |
NiCo2O4@CNFs | ∼4.0 | ∼2.6 | 4179/100 | 350/1000/200 | 1.0 M LiTFSI in TEGDME | 190 |
NCO@CMs | ∼4.0 | ∼2.7 | 6489.5 | 90/1000/ | 1.0 M LiTFSI in TEGDME | 191 |
NCO-500 | ∼4.2 | ∼2.7 | 9231/100 | 80/600/100 | 1.0 M LiTFSI in TEGDME | 193 |
3D foam-like NiCo2O4 | ∼3.95 | ∼2.87 | 10![]() |
80/1000/200 | 1.0 M LiCF3SO3 in TEGDME | 194 |
Bowl-like NiCo2O4@CFPs | ∼3.55 | ∼2.7 | 9624.2/100 | 100/500/100 | 1.0 M LiTFSI in TEGDME | 195 |
NCO@BCNNT | ∼4.0 | ∼2.8 | 9823/100 | 320/1000/500 | 1.0 M LiCF3SO3 in TEGDME | 196 |
CuNiCo-5-5@Ni mesh | ∼3.9 | ∼2.8 | ∼7300/− | — | 1.0 M LiNO3 in DMSO | 197 |
CCNO | ∼3.49 | ∼2.9 | 9421.7/500 | 174/1000/500 | 1.0 M LiClO4 in DMSO | 198 |
TP-NCO/MO | ∼3.73 | ∼3.0 | ∼6 mAh cm−2/0.2 mA cm−2 | 480/0.5 mAh cm−2/0.5 mA cm−2800/0.2 mAh cm−2/0.2 mA cm−2 | 1.0 M LiTFSI in TEGDME | 199 |
NiCo2O4/NixP | ∼3.4 | ∼2.9 | 7.41 mAh cm−2/0.1 mA cm−2 | 563/0.2 mAh cm−2/0.2 mA cm−2 | 1.0 M LiTFSI in TEGDME | 200 |
NCO 120 | ∼3.6 | ∼2.85 | 13![]() |
173/500/100 | 1.0 M LiTFSI in TEGDME | 201 |
Transition metal compounds (other transition metals and their compounds and composites) | ||||||
TiC | ∼3.5 | ∼2.6 | — | 100/350/1000 mA cm−2 | 0.5 M LiClO4 in DMSO | 46 |
∼3.5 | ∼2.75 | — | 25/525/1000 mA cm−2 | 0.5 M LiPF6 in TEGDME | ||
Mo3P | 3.23 | 2.88 | — | 1200/500/500 | 0.3 M LiTFSI in a 72:25 volumetric ratio of DMSO: EMIM-BF4 mixture with 25 × 10−3 m of each of TEMPO and DBBQ RMs | 47 |
Co3O4–TiO2(B) | ∼3.9 | ∼2.75 | 11![]() |
200/1000/100 | 1.0 M LiTFSI in TEGDME | 57 |
V-TiO2/Ti3C2Tx | ∼2.9 | ∼2.75 | 11![]() |
200/1000/100 | 1.0 M LiTFSI in TEGDME | 58 |
V2−xO5@V2C MXene, V−400 | ∼3.25 | ∼2.9 | >5000/400 | 501/1000/100 | 1.0 Mol L−1 LiTFSI in 100 μL of TEGDME | 69 |
Ni2P/Ni12P5@NF | 3.63 | 2.74 | 13254.1/500 | 120/1000/500 | 1.0 M LiTFSI in TEGDME | 71 |
NiS2-NSs | ∼4.0 | ∼2.6 | 22![]() |
314/1000/1000 | 1.0 M of LiNO3 in DMAC with 0.1 M of TEMPO RM | 72 |
MoS2−x/hEG | ∼3.85 | ∼2.5 | 19000.3/500 | >500/1000/1000 | 1.0 M LiTFSI in TEGDME | 73 |
CoO/Ti3C2Tx | 3.26 | 2.72 | 16![]() |
>160/500/100 | 1.0 M LiTFSI in TEGDME | 78 |
CoP CPHs | ∼3.6 | ∼2.95 | 33![]() |
236/1000/300 | 1.0 M LiTFSI in TEGDME | 161 |
CoP NPs | ∼3.8 | ∼3.0 | 20![]() |
221/1000/300 | ||
Carbon-based materials | ||||||
N1-wdC-900 | ∼3.8 | ∼2.8 | 9.44 mAh cm−2/0.05 mA cm−2 | 113/0.05 mAh cm−2/0.05 mA cm−2 | 1.0 M LiTFSI in TEGDME | 30 |
Carbon-black-based self-standing membrane | ∼3.5 | ∼2.65 | 7000/0.4 mA cm−2 | 7/3 mAh cm−2/0.3 mA cm−2 | 0.5 M LiTFSI, 0.5 M LiNO3, and 0.5 M LiBr | 31 |
MACF | ∼3.5 | ∼2.6 | 11![]() |
110/1000/300 | 0.05 M LiI and 1.0 M LiCF3SO3 in TEGDME | 34 |
N-rGO | 4.21 | 2.64 | 10![]() |
40/1000/400 | 1.0 M LiTFSI in TEGDME | 70 |
Woven CNT | ∼4.7 | ∼2.6 | >2500/2000 | >60/1000/2000 | 1.0 M LiPF6 in 0.21 mL TEGDME | 83 |
SWNT/CNF buckypapers | — | — | >2500/0.1 mA cm−2 | — | 1.0 M LiPF6 in PC and THF | 84 |
GNSs | — | — | 8705.9/75 | — | 1.0 M LiPF6 in PC/EC (1![]() ![]() |
85 |
N-a-ex-G/KB | — | — | ∼11![]() |
— | 1.0 M LiPF6 in TEGDME | 86 |
3D HGNs | ∼4.0 | ∼2.75 | ∼3600/100 | 18/1000/100 | 1.0 M LiTFSI in TEGDME | 87 |
Mesocellular carbon foam | — | — | 2500/0.1 mA cm−2 | — | 1.0 M LiClO4 in PC | 88 |
MCC | ∼3.8 | ∼2.8 | 26![]() |
25/1000/400 | 0.1 M LiClO4 in DMSO | 92 |
Hierarchically woven CNT fibrils | ∼4.4 | ∼2.6 | — | 70/1000/2000 | 1.0 M LiPF6 in TEGDME | 95 |
N-CNTs | — | — | 866/75 | — | 1.0 M LiPF6 in PC | 99 |
1D C@LSSO NFs | ∼3.7 | ∼2.7 | 8058.3/200 | 54/500/200 | 1.0 M LiTFSI in TEGDME | 100 |
N-GNSs | — | — | 11![]() |
— | 1.0 M LiPF6 in TEGDME | 101 |
S-GNSs | — | — | ∼4300/75 | — | 1.0 M LiPF6 in TEGDME | 102 |
Co-SAs/N-C | ∼3.4 | ∼2.8 | 20![]() |
260/1000/400 | 1.0 M LiTFSI in TEGDME | 103 |
Porous graphene (PEG-2) | ∼4.1 | ∼2.75 | 29![]() |
20/500/200 | 0.1 M LiClO4 in DMSO | 104 |
CNT@RuO2 | 3.56 | 2.69 | 4350/385 | 100/500/100 | LiTFSI/TRIGLYME (1![]() ![]() |
108 |
TiC/MWNTs | ∼3.4 | ∼2.65 | 1800/− | 10/1000/250 | 0.5 mol L−1 LiClO4 in DMSO | 109 |
AAO/Ta/Fe/CNT | ∼4.2 | ∼2.7 | ∼4900/43 | >10/—/309 | 0.1 M LiClO4 in DME | 163 |
Ni-CNTs@Ni | — | — | 1814/0.05 mA cm−2 | 9/500/0.05 mA cm−2 | 1.0 M LiPF6 in DMSO | 164 |
Noble metals and their alloys, oxides, and composites | ||||||
Nanoporous gold (NPG) | ∼3.4 | ∼2.6 | 325/500 | 100/—/500 | 0.1 M LiClO4 in DMSO | 33 |
RuO2 decorated MACF (R-MACF) | ∼3.5 | ∼2.8 | 13![]() |
154/1000/300 | 0.05 M LiI and 1.0 M LiCF3SO3 in TEGDME | 34 |
PtRu | ∼4.0 | ∼2.75 | 3689/100 | 120/1000/500 | 1.0 M LiTFSI in TEGDME | 35 |
PtAu | ∼3.4 | ∼2.7 | 5049/100 | 220/1000/500 | 1.0 M LiCF3SO3 in TEGDME | |
Anisotropic Pt | ∼3.1 | ∼2.75 | 12![]() |
70/1000/1000 | 1.0 M LiCF3SO3 in TEGDME | 36 |
Pt3Co/KB | ∼3.2 | ∼2.7 | 5600/200 | 70/1000/200 | 1.0 M LiTFSI in TEGDME | 37 |
PtIr multipods | 3.29 | 2.85 | 8698/100 | 180/1000/100 | NMP | 38 |
Pd/Al2O3/C | ∼3.2 | ∼2.7 | 2750/100 | 15/500/100 | 1.0 M LiCF3SO3 in TEGDME | 40 |
TOH Au NCs@SP | 3.63 | ∼2.8 | ∼20![]() ![]() |
>30/1000/200 | 1.0 M LiNO3 in DMSO | 41 |
PdFe/N-rGO | 4.08 | 2.62 | ∼4800/200 | 400/1000/400 | 1.0 M LiTFSI in TEGDME | 70 |
Pd/N-rGO | 3.773 | 2.767 | ∼6700/200 | 100/1000/400 | ||
PdCo/N-rGO | 3.90 | 2.83 | ∼5800/200 | 120/1000/400 | ||
PdNi/N-rGO | 4.08 | 2.84 | ∼4900/200 | 200/1000/400 | ||
Ru-CB | ∼3.3 | ∼2.8 | 9800/200 | 150/1000/20040/4000/200 100/1000/1000 100/1000/400 |
0.1 M LiClO4 in DMSO | 82 |
Pt-HGNs | ∼4.0 | ∼2.75 | ∼5600/100 | 54/1000/100 | 1.0 M LiTFSI in TEGDME | 87 |
Ru@MCC | ∼3.5 | ∼2.8 | — | 120/1000/400 | 1.5 M LiNO3 in DMSO | 92 |
Pt/CNT | ∼3.8 | ∼2.6 | ∼2500/− | 130/1000/2000 | 1.0 M LiPF6 in TEGDME | 95 |
Ru@PEG-2 | ∼3.6 | ∼2.8 | 17![]() |
200/500/200100/1000/200 | 0.1 M LiClO4 in DMSO | 104 |
TiC/MWNTs-Ru | ∼3.5 | ∼2.9 | — | 90/1000/250 | 0.5 mol L−1 LiClO4 in DMSO | 109 |
PdMo NWs@SWCNTs | ∼4.2 | ∼2.75 | >10![]() |
243/500/300105/1000/300 | 1.0 M LiTFSI in TEGDME | 115 |
AgPd-3 NTs | ∼3.4 | ∼2.7 | 2650/0.2 mA cm−2 | 100/1000/0.2 mA cm−2 | 1.0 M LiCF3SO3 in TEGDME | 116 |
PdCu nanoparticles (PdCu NPs) | ∼3.7 | ∼2.7 | >12![]() |
50/1000/200 | 1.0 M LiCF3SO3 in TEGDME | 117 |
RuO2–Co3O4 | ∼3.7 | ∼2.9 | 19![]() |
>100/500/200 | 1.0 M LiTFSI in TEGDME | 120 |
FeSA–RuO2/HPCS | ∼3.4 | ∼2.8 | 23![]() |
232/1000/200 | 1.0 M LiSO3CF3 in TEGDME | 121 |
RuO2/HPCS | ∼3.75 | ∼2.8 | 19![]() |
134/1000/200 | ||
Fe2O3/HPCS | ∼3.8 | ∼2.75 | 17![]() |
70/1000/200 | ||
IrO2/MnO2 | ∼4.1 | ∼2.5 | 16![]() |
312/−/1600 | 1.0 M LiClO4 in TEGDME | 122 |
RuO2 hollow spheres | ∼3.6 | ∼2.7 | 1380/100 | 100/−/500 | 0.5 M LiClO4 in DMSO | 123 |
N-doped graphene with encapsulated RuO2 nanoparticles | ∼3.7 | ∼2.6 | 8700/200 | >110/2000/400 | 1.0 M LiTFSI in TEGDME | 124 |
IrO2/KB | ∼3.56 | ∼2.75 | 4500/0.1 mA cm−2 | 70/500/0.1 mA cm−2 | 1.0 M LiCF3SO3 in TEGDME | 125 |
FeCoNiMnPtIr | ∼3.34 | ∼2.75 | 39![]() |
∼150/1000/200 | 1.0 M LiTFSI in TEGDME | 218 |
PtPdIrRuAuAg SNRs | ∼3.4 | ∼2.8 | 5252/100 | 100/1000/500 | 1.0 M LiSO3CF3 in TEGDME | 219 |
Hollow RuIrFeCoNi nanoparticles | ∼3.6 | ∼2.8 | — | 80/4000/2000 | 1.0 M LiTFSI in DME | 220 |
PtRuFeCoNi HEA@Pt | ∼3.12 | ∼2.75 | 8400/100 | 210/1000/100 | 1.0 M LiTFSI in TEGDME | 221 |
PtFeCoNiCu@rGO | ∼3.57 | ∼2.8 | 13![]() |
148/500/100 | 1.0 M LiTFSI in TEGDME | 222 |
HEA@Pt-PtSAs-M | ∼3.2 | ∼2.5 | >12![]() |
470/1000/1000 | 1.0 M LiTFSI in TEGDME | 223 |
MOF-based materials | ||||||
N-Fe-MOF | — | — | 5300/50 | 50/∼1600/400 | 1.0 mol L−1 LiPF6 in TEGDME | 110 |
Mn-MOF-74 | ∼4.1 | ∼2.75 | 9420/50 | 30/1000/250 | TEGDME | 130 |
Nickel-hexaiminotriphenylene (Ni-HTP) | ∼3.9 | ∼2.7 | 18![]() |
150/1000/500 | 1.0 M LiTFSI in G4 | 132 |
NiRu-HTP | ∼3.63 | ∼2.75 | 15![]() |
200/1000/500 | ||
ZnO/ZnFe2O4/C nanocages (ZZFC) | 4.05 | 2.62 | 11![]() |
15/5000/300 | 1.0 mol L−1 LiTFSI in TEGDME | 133 |
Ni-MOFs | ∼4.0 | ∼2.8 | 13554.1/500 | 159/600/50093/600/1000 | 1 mol L−1 LiNO3 in DMSO | 134 |
CoNi-MOFs | ∼4.0 | ∼2.8 | 20370.13/20015784.2/500 13490.02/1000 |
186/600/500127/600/1000 | ||
3D printed Co-MOF (3DP-NC-Co) | ∼3.7 | ∼2.75 | 1124/0.05 mA cm−2525/0.8 mA cm−2 | 16/1 mAh/0.1 mA cm−2 | 0.5 M LiClO4 in DMSO | 135 |
Fe-UiO-66f | 3.56 | 3.42 | ∼2.5 mAh cm−2/0.02 mA cm−2 | 500/—/0.01 mA cm−2 | 1.0 M LiTFSI in TEGDME | 136 |
Considering the above-mentioned factors, the potential of developed materials as Li–O2 cathode catalysts can be in a systematically and comprehensively assessed.
In the case of transition-metal spinel oxides, morphological control and pore structure optimization are the key to improving their catalytic performance. Side reactions need to be suppressed for carbon-based materials. Thus, electronic structure adjustment, crystal plane effects, and formation of composites with noble metal-free materials are the common strategies applied for the preparation of noble-metal-based catalysts. In the case of MOF-based materials, despite their controllable pore structure, they have the drawbacks of poor electronic conductivity and complex synthesis. Thus, to achieve the practical application of Li–O2 batteries, the impeding problems must be addressed.
Due to their many advantages,5,49,51,52 including low cost, easy synthesis, controllable structure, high activity in ORR and OER, and good thermodynamic stability in alkaline media, spinels are widely used in LOBs. Moreover, to optimize their electronic structure, increase the number of active sites, and improve their electrical conductivity, many modification strategies9,20,26,53 have been applied, including doping modification, vacancy engineering, heterostructure engineering, crystal plane effects, synergistic effects, and dimension engineering. In this chapter, several common spinel oxides will be introduced.
![]() | ||
Fig. 3 (a) Schematic synthesis of Ni-Co3O4-Vo. (b) EPR curves of Ni-Co3O4-Vo, Ni-Co3O4, and Co3O4. (c) High-resolution XPS spectra of Co 2p. (d) High-resolution XPS spectra of O 1s. (e) Partial typical discharge/charge curves of Ni-Co3O4-Vo during cycling. (f) Cyclic performance of the three cathodes under fast charging and slow discharging conditions (the limited capacity of the battery was set to 500 mAh g−1, the charging current density was set to 1000 mA g−1, and the discharging current density was set to 100 mA g−1). (g) XPS analysis of the Ni-Co3O4-Vo electrode after discharge and recharge. (a)–(g) Reproduced with permission.54 Copyright 2025, Elsevier. |
Defect/vacancy engineering has been proven to be an effective modulation strategy in metal–air batteries, which is sometimes coupled with other strategies.5,26,59 Apart from anionic vacancies, cationic vacancies are also common in spinel oxides.
Liu et al. synthesized cobalt oxides with cationic vacancies (Co3−xO4) via the thermal treatment of the glycerolatocobalt (GlyCo) nanostructure.60 At a low temperature of 300 °C in air, the thermal decomposition of its surface functional groups (e.g., Co–O bonds) led to the formation of Co2+ cation vacancies. The discharge/charge specific capacities of the LOB using the synthesized Co3−xO4 were 14517/13254 mAh g−1 at a current density of 100 mA g−1. These excellent capacity values indicate that the change in the electronic structure caused by Co2+ vacancies effectively improved the deposition of Li2O2.
Due to their wide range of advantages, including high specific surface area, quantum size effect and easy heterostructure engineering with a support, nanomaterials are widely used in energy storage and electrochemistry.61–63 Zhai et al. synthesized Co3O4 nanoparticle evenly dotted hierarchical-assembled carbon nanosheet frameworks (Co3O4-HCNF) through a simple two-step carbonization–calcination process.64 Owing to its large BET specific surface area and unique internal structure, the HCNF significantly boosted the electron conduction. Thus, the Li–O2 battery assembled using the Co3O4-HCNF cathode showed outstanding ORR/OER capacities (14901/14
948 mAh g−1 at a current density of 100 mA g−1, respectively) and superior cycling stability and reversibility. Ryu et al. used Co3O4 nanofibers fixed on non-oxidized graphene nanoflakes as catalysts,65 which resulted in a high performance. Liu et al. synthesized Pt-doped Co3O4 nanowires hydrothermally on acid-treated carbon paper substrates.66 These nanowires were closely stacked and displayed an urchin-like morphology. It was found that the doped Pt changed the discharge products from solution-pathway-grown flake-like structures to surface-pathway-grown film-like structures. Then, the tight contact between the film-like discharge products and Co3O4 nanowires was favorable for the OER/ORR kinetics, which led to a low overpotential and high stability.
Heterostructure engineering and crystal plane effects are also widely used strategies for the synthesis of spinels and have been proven to be effective in improving the performance of Li–O2 batteries.48–52,67,68 For example, heterostructures such as the Co2N/Co3O4 structure optimized the electronic structure of the active sites, which led to an appropriate adsorption energy for LiOx, and simultaneously facilitated ORR and OER (Fig. 4b). Furthermore, the highly conductive MXene-based heterostructure accelerated the electron transfer.67 The XPS spectra showed a shift in the Co2+ peak to a higher binding energy due to the charge transfer caused by the electron-withdrawing effect (Fig. 4a). Thus, the LOB based on the heterostructure exhibited a higher reversible capacity (14271 mAh g−1), a much lower overpotential (0.65 V) during the first cycle, and more than 300 stable cycles compared to other cathodes (Fig. 4c and d).
![]() | ||
Fig. 4 (a) High-resolution XPS spectra of Co 2p for NCT, Co3O4–Ti3C2Tx and Ti3C2Tx materials. (b) Adsorption energy of LiO2 and Li2O2 on different slabs. (c) Discharge–charge curves in the first cycle at 500 mA g−1 for different cathodes tested in a pure oxygen atmosphere. (d) Cyclability at 500 mA g−1 for different cathodes tested in a pure oxygen atmosphere. (e) SEM image of the discharged electrode of a {001} faceted Co3O4 cube. (f) SEM image of the discharged electrode of a {112} faceted Co3O4 plate. (g and h) Atom states of the edge of Co3O4-H and Co3O4, respectively. (i) Co 2p XPS spectra of Co3O4-H and Co3O4. (j) O 1s XPS spectra of Co3O4-H and Co3O4. (a–d) Reproduced with permission.67 Copyright 2024, Elsevier B.V. All rights reserved. (e and f) Reproduced with permission.159 Copyright 2015, the American Chemical Society. (g–j) Reproduced with permission.160 Copyright 2019, the American Chemical Society. |
Crystal plane effects can cause great differences in the same catalyst material. Co3O4 plates dominantly enclosed by {112} facets were fully covered by the discharge product (Fig. 4e and f), whereas a small amount of discharge product was attached to only the edges of Co3O4 nanocubes with {001} facets as the dominant exposed planes, possibly due to the relatively higher surface energy of the edges.159 Therefore, Song et al. suggested that nucleation sites such as Co3+ and high surface energy atomic steps seem to be preferential for the growth of Li2O2. The LOB based on the {112}-faceted Co3O4 plate showed a larger maximum capacity of 9144 mAh g−1, a lower overpotential and a greater number of cycles than the {001}-faceted Co3O4 cube. In the case of Co3O4 nanosheets, the nanosheets with an edge-enriched {111} plane (marked as Co3O4-H) had more atomic steps and kink atoms, and exposed more Co2+ (the green spheres in Fig. 4g and h) than the poor-edge plane Co3O4, which introduced some oxygen vacancy sites, as proven by the XPS spectra Fig. 4i and j, respectively.160 The more exposed active sites enhanced the electrochemical performance of the Li–O2 battery based on Co3O4-H with a lower overpotential, a higher initial capacity and greater cycling stability.
ZnCo2O4 was first used as a cathode catalyst in Li–O2 batteries by Hung et al.177 The TEM image clearly showed the presence of mesopores in the ZnCo2O4 nanosheets (Fig. 5a and b). Li et al. synthesized 3D ordered mesoporous ZnCo2O4 and its TEM images showed clearly arranged mesopores (Fig. 5c and d).178 Song et al. fabricated ZnCo2O4 nanowires on carbon cloth, which possessed mesopores (Fig. 5e and f).181 These mesopores play an essential role in O2 diffusion and decide the upper limit of discharge product deposition.
![]() | ||
Fig. 5 (a) TEM micrograph and (b) lattice fringes of 2D mesoporous ZnCo2O4 NFs. The inset of (a) shows the corresponding SAED pattern. The scale bars of (a) and (b) are 200 nm and 5 nm, respectively. Reproduced with permission.177 Copyright 2013, RSC. (c and d) TEM images of as-prepared ordered 3D mesoporous ZnCo2O4. Reproduced with permission.178 Copyright 2015, Elsevier. (e and f) HRTEM images of NiCo2O4. Reproduced with permission.181 Copyright 2020, Elsevier. |
In general, spin engineering involves elevating the spin state of the B ion in AB2O4 to increase the number of unpaired electrons, which can make the material more active moderately. In the case of ZnCo2O4, the low spin state Co3+ without unpaired electrons (t2g6eg0) cannot trap and interact with the reactants or intermediates. Spin engineering can transform the low spin state Co3+ into high spin state Co3+ (t2g4eg2) with four unpaired electrons (Fig. 6a). Ren et al. obtained ZnCo2O4 with high spin state Co3+ by increasing the calcination temperature.183 They suggested that the Co–O–Co spin tunnel facilitated electron transfer (Fig. 6a). Pan et al. introduced Cu2+ in the tetrahedral sites to partially replace Zn2+, and constructed a Cu–O–Co spin tunnel (Fig. 6b).182
![]() | ||
Fig. 6 (a) Illustration of the different spin configurations of Co3+ and the formation of Co–O–Co spin channel with an increase in the Co3+ spin state. Reproduced with permission.183 Copyright 2023, Elsevier. (b) Illustration of the new co-top oxygen linkage Cu2+(Td)−O−Co3+(Oh) and the change in the spin configuration of Co3+. Reproduced with permission.182 Copyright 2023, Elsevier. |
A change in the spin state can adjust the ORR routes (the surface route and the solution route). Ren and coworkers observed a smaller amount of discharge products on ZnCo2O4 with high spin state Co3+ (Fig. 7a–f),183 which indicated that the ORR route changed from the solution route to the surface route. Pan and coworkers detected evenly distributed discharge products, a case between film-like Li2O2 by the surface route and large-sized Li2O2 by the solution route, as shown in Fig. 7g.182
![]() | ||
Fig. 7 SEM images of ZCO-350 electrodes in the (a) pristine state, (b) first discharged and (c) first recharged states. SEM images of ZCO-650 electrodes in (d) the pristine state, (e) first discharged and (f) first recharged states. Reproduced with permission.183 Copyright 2023, Elsevier. (g) Schematic of the ORR (discharge) and OER (charge) process for ZnCo2O4, Cu1–ZnCo2O4, and Cu2–ZnCo2O4. Reproduced with permission.182 Copyright 2023, Elsevier. |
The development of NiCo2O4 as an Li–O2 cathode catalyst has shifted from focusing on its structure, especially mesopores and porosity in previous studies, to changing its electronic structure through interface engineering,196 doping engineering,197,198 heterostructure engineering,199 amorphization strategy200 and stoichiometry optimization201 in recent research. These strategies can change the Ni3+/Ni2+ ratio and Co3+/Co2+ ratio, and obtained catalysts show a better performance, which indicates their improved electronic structure.
Li et al. emphasized the improved charge transfer at the interfaces between NiCo2O4 and boron carbon nitride (BCN) nanotubes.196 In the case of the composite of NiCo2O4 and BCN nanocubes (NCO@BCNNT), compared to NiCo2O4, a lower Ni3+/Ni2+ ratio (1.47 vs. 2.85) and a slightly lower Co3+/Co2+ ratio (0.55 vs. 0.62) were observed in its XPS spectra (Fig. 8a and b), respectively. The Nyquist plots showed that the diameter of the hemisphere in the high frequency region of NCO@BCNNT was the smallest among BCN, NCO and NCO@BCNNT (Fig. 8c), which revealed the accelerated electronic transfer in NCO@BCNNT. This was primarily attributed to the change in its electronic structure at the interfaces.196
![]() | ||
Fig. 8 High-resolution XPS spectra of (a) Ni 2p and (b) Co 2p. (c) Nyquist plots of batteries in the frequency range of 105–10−2 Hz. Reproduced with permission.196 Copyright 2021, Elsevier. (d) Magnetic field dependence of magnetization (M–H) profiles of CNO and CCNO at 300 K (the inset shows a partial enlarged image). (e) Co 2p and (f) Ni 2p spectra of CNO and CCNO. Reproduced with permission.198 Copyright 2022, Elsevier. |
Introducing heterogeneous spin states has been proven to be an effective method for improving the performance of catalysts, which can be detected based on the magnetic field dependence of magnetization (M–H) curves.202 These states are introduced by adding foreign metal ions with a similar atomic size and electronic structure to the crystal lattice, inevitably generating a regional Coulomb force imbalance. Then, this leads to a slight disorder in atomic arrangement in the lattice and exposure of more active sites. Ren et al. doped Ce atoms in NiCo2O4 nanowires and successfully introduced heterogeneous spin states (Fig. 8d).198 Similarly, they also observed a relatively obvious change in the Ni3+/Ni2+ ratio and nearly no change in the Co3+/Co2+ ratio (Fig. 8e and f, respectively), in accordance with the study by Li and coworkers.196 This may be a result of the smaller radius of Ni2+ and its position in the tetrahedrons, making changes in the Ni3+/Ni2+ electronic configuration easier.
It has been reported that amorphous materials provide rich active sites not only on their surfaces but also inside their structure because they possess abundant randomly oriented dangling bonds and unsaturated coordination sites.203 Xia et al. constructed a crystalline NiCo2O4 (NCO)/amorphous NixP (NP) heterostructure (c/a) catalyst, which could lower the adsorption of LiO2 and enhance the adsorption of Li (Fig. 9a).200 In the case of LiO2 on NCO/NP, Li+ loses more electrons and O receives less electrons (Fig. 9b), weakening the adsorption of LiO2. Therefore, Li2O2 was deposited on NCO/NP via the solution route (Fig. 9c).
![]() | ||
Fig. 9 (a) Adsorption energies of NiCo2O4 and NiCo2O4/Ni3P for discharge intermediates. (b) CDD and Bader charge for LiO2 adsorbed on NiCo2O4 and NiCo2O4/Ni3P, respectively. (c) Schematic reaction mechanism of Li2O2 generation on the surface of NCO/NP and NCO cathodes. Reproduced with permission.200 Copyright 2025, Elsevier. (d) Adsorption energy of the discharge intermediates for the 4e− and 2e− pathways. (e) CDD and Bader charge for LiO2 adsorbed on NiCo2O4 and NiCo2O4/MnO2. (f) TEM image of the NCO/MO cathode after deep discharge at 0.2 mA cm−2 (SAED in insert) with NCO/MO tested as a free-standing air cathode. (g) Schematic reaction mechanism of TP-NCO/MO. Reproduced with permission.199 Copyright 2024, Yongji Xia et al. |
Due to the electronic rearrangement at heterostructure interfaces that prompt charge transfer, the Motty–Schottky effect has been applied for the preparation of catalysts for use in Li–O2, Li–S, and Li–CO2 batteries. Recently, Xia et al. designed an NiCo2O4/MnO2 heterostructure.199 Similar to a previous study,200 this heterostructure exhibited stronger adsorption of Li and weaker adsorption of LiO2 (Fig. 9d), and Li+ also lost more electrons and O received less electrons (Fig. 9e). After deep discharge at 0.2 mA cm−2, chip-like Li2O2 was observed (Fig. 9f). It was explained that Li2O2 chips were obtained owing to the competition between the surface route and the solution route (Fig. 9g).199
The similarities between these studies reveal the reasons why the intrinsic catalytic activity of NiCo2O4 is not ideal, including its slow charge transfer, weak adsorption of Li+ and too strong adsorption of LiO2. Thus, future studies on NiCo2O4 can further focus on the exploration of other heterostructures.
Du et al. compared concave polyhedron CoP with exposed (211) crystal planes (CoP CPHs) and CoP polyhedron particles with exposed (011) and (111) crystal planes (commercial CoP).161 During the first cycle, Li2O2 on CoP CPHs was completely decomposed (Fig. 10a–c); however, film-like Li2O2 still remained on commercial CoP (Fig. 10d–f). This indicated that CoP CPHs had better reversibility than the commercial CoP. The CoP-CPH-based LOB exhibited a low overpotential (0.67 V), a maximum discharge capacity of 33743 mAh g−1 and long cycle life of 950 h due to its large density of atomic steps, edges, ledges, and kink atoms. Wang et al. used ultrafine Co3O4 nanocrystals to decorate atomic-thick TiO2(B) as a highly efficient catalyst,57 as shown in Fig. 11a. Due to the ion doping effect, oxygen vacancies were induced on TiO2 nanosheets by the decoration of Co3O4 nanocrystals. The existence of these vacancies was proven by XPS, as shown in Fig. 11b, which showed a dramatic shift in the peaks of Co3O4–TiO2(B) compared to TiO2(B). Compared to the LOBs with the TiO2(B) electrode with only 70 cycles (Fig. 11c), the LOBs with the Co3O4–TiO2(B) electrode showed 200 stable and reversible cycles (Fig. 11d). Based on the outcomes, Wang et al. proposed the possible mechanism (Fig. 11e), suggesting that the oxygen vacancies and the close contact between Li2O2 and Co3O4–TiO2 boost the ORR and OER activity. Zheng et al. synthesized oxygen vacancy-rich TiO2 nanoparticles on Ti3C2Tx MXene (V-TiO2/Ti3C2Tx) nanosheets through HF etching and ethanol-thermal treatment (Fig. 11f).58 The LOBs with the V-TiO2/Ti3C2Tx electrode showed a much lower overpotential (0.21 V) during the first discharge/charge circle in comparison to the Ti3C2Tx- (1.12 V) and TiO2-based (1.45 V) cells (Fig. 11g). The V-TiO2/Ti3C2Tx-based LOBs also showed a high specific capacity (11
487 mAh kg−1) and long cycle life (over 200 cycles). Thus, these results not only show the important role of vacancies but also the improved electronic conductivity brought by MXenes, which will be further discussed below.
![]() | ||
Fig. 10 SEM images of CoP CPHs (a–c) and commercial CoP (d–f) electrodes in different charging/discharging states; (a and d) initial electrodes, (b and e) after first discharging and (c and f) after first charging. (a–f) Reproduced with permission.161 Copyright 2021, Elsevier Inc. |
![]() | ||
Fig. 11 (a) Schematic depicting the synthesis of Co3O4–TiO2(B). Self-assembly involves a solvothermal process. Decoration of Co3O4 nanocrystals uses a hydrothermal method. (b) High-resolution XPS spectra of the Ti 2p for Co3O4–TiO2(B) and TiO2(B). (c) Cyclic performance of Li–O2 batteries with the TiO2(B) electrode. (d) Cyclic performance of Li–O2 batteries with the Co3O4–TiO2(B) electrode. (e) Schematic of the working mechanism of Co3O4–TiO2(B) catalysts. (f) Schematic of the procedure for the preparation of V-TiO2/Ti3C2Tx. HF etching: 4 h in 40% HF at room temperature. Ethanol-thermal: solvothermal synthesis at 140 °C for half of a day. (g) First discharge/charge curves of TiO2-, Ti3C2Tx- and V-TiO2/Ti3C2Tx-based Li–O2 batteries with a 1000 mAh g−1 cutoff capacity at 100 mA g−1. (a)–(e) Reproduced with permission.57 Copyright 2018, the American Chemical Society. (f) and (g) Reproduced with permission.58 Copyright 2019, the American Chemical Society. |
Single-transition-metal oxides in situ decorated on MXenes have been proven to be effective catalysts. Ti3C2Tx, a member of the MXene family, was first produced via the exfoliation of Ti3AlC2 in 2011.74 The MAX phases with the formula Mn+1AXn are layered hexagonal structures, where n can be 1, 2, or 3. M is an early transition metal including Ti, Zr, and V. A is mainly a group IIIA or IVA element, and X is C and/or N.58,74,75 By the etching the A layers from the MAX phases, early transition metal carbides and/or carbonitrides, denoted as MXenes, are produced. Therefore, MXenes are labeled as Mn+1XnTx, where Tx represents terminal groups such as –OH and –F.58 Considering the fact that there are more than 60 known MAX phases,74 the potential number of MXenes is huge. Due to their good electronic conductivity and high surface areas, MXenes have been widely used in the energy storage field in LIBs,76 supercapacitors,77 and LOBs.58,69,78 Recently, Xu's group synthesized cationic vanadium vacancy-enriched V2−xO5 on V2C MXene (V2−xO5@V2C MXene) as a bifunctional catalyst for LOBs.69 The Brunauer–Emmett–Teller (BET) measurement showed the presence of rich mesopores in this material, along with a high surface area, providing mass transfer pathways and active sites for electrochemical reactions. Li et al. produced CoO nanoparticle-decorated MXene nanosheets (CoO/Ti3C2Tx) as a cathode material for LOBs.78 Owing to its good electronic conductivity, MXene improved the interface electronic transfer rate. The N2 adsorption–desorption analysis showed that the prominent pores were mesopores. The BET results showed that the CoO nanoparticles further improved the specific surface area. Therefore, the decoration of transition metal oxide nanoparticles on MXenes can combine their individual advantages, showing potential as another effective cathode catalyst in lithium–oxygen batteries.
In conclusion, owing to their advantages such as relatively low cost and natural abundance, transition metals and their compounds have demonstrated significant potential as candidate catalysts for LOBs. These materials are often engineered through modulation strategies to enhance their electrochemical performances as cathode materials in LOBs, including doping engineering, defect/vacancy engineering, crystal plane effects and heterostructure engineering such as decorating them on supports such as MXenes. LOBs with these catalysts have shown improved stability and reversibility, excellent specific energy density, and extended cycle life. Thus, advances in this field will accelerate the commercialization of LOBs.
Studies have proven that the capacity of Li–O2 cells largely depends on the pore volume and number of mesopores in their carbon electrode.89,90 During discharge, oxygen moves to the carbon electrode through the electrolyte in the form of either O2 gas or dissolved oxygen.32 The cathode reactions take place at the interface among the solid electrode, the electrolyte and the O2 gas phase, where Li+ and O2 meet and eventually form Li2O2. Micropores are too small to allow the deposition of more Li2O2 because they become blocked and their small sizes make the diffusion of oxygen difficult. Conversely, macropores reduce the volume efficiency. The solid–liquid–gas tri-phase regions possess less macropores, which leads to the lower production of Li2O2.32
Besides mesopores, macropores with the appropriate size also play a role in carbon materials. Ding et al. found that at the discharge current of 0.1 mA, the cell capacity increased with an increase in pore size and peaked at the pore size of 80 nm.91 This indicates that small macropores also play an important role in oxygen diffusion and Li2O2 deposition.
Therefore, the synthesis of mesoporous and/or small macroporous carbon materials is the key to improving the capacity of LOBs. Yang et al. prepared mesocellular carbon foam (MCF-C) via nanocasting technology using a mesocellular foam (MCF) silica hard template.88 The MCF-C showed clear spherical pores of the same size, whose inner diameter was about 28 cm and outer diameter was about 35 cm, as shown in Fig. 12b. Sun et al. reported the preparation of mesoporous carbon nanocubes with numerous hierarchical mesopores and macropores.92 The TEM images (Fig. 12c and d) showed the presence of many interconnected mesopores. The N2 adsorption–desorption plots of MCCs showed that large-size pores accounted for most of the pore volume, and the inset indicated that the pore sizes were mainly about 50 and 100 nm (Fig. 12a). Guo's group reported the synthesis of three-dimensional ordered mesoporous and macroporous carbon sphere arrays (MMCSAs).93 These carbon spheres were hexagonal, with a diameter of about 200 nm, and possessed interstices of about 60 nm between them (Fig. 12e). Fig. 12f shows that the mesopores had a size of approximately 8 nm and a wall thickness of 3 nm. Sakaushi et al. produced mesoporous noble carbons that were N-doped and showed astonishing oxygen stability.94 The TEM images (Fig. 12g and h) and pore size distribution (Fig. 12i) graphs showed that this material had a mesoporous structure with an average pore diameter of about 8 nm. The mesopores and small macropores in this material not only promoted O2 diffusion and facilitated Li+ diffusion to the cathode immersed in electrolyte but also provided enough space for Li2O2 deposition.
![]() | ||
Fig. 12 (a) Nitrogen adsorption/desorption isotherms and pore-size distribution of the MCCs. Reproduced with permission.92 © 2015, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (b) TEM image of the MCF-C. Reproduced with permission.88 Copyright 2009, Elsevier. (c) Low-magnification and (d) high-magnification TEM images of the MCCs. Reproduced with permission.92 © 2015, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. Characteristics of the as-prepared MMCSAs: (e and f) TEM images at different magnifications. Reproduced with permission.93 © 2013, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (g and h) TEM images of the meso-NdCs. Reproduced with permission.94 © 2015, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (i) Pore size distributions of meso-NdCs. Reproduced with permission.94 © 2015, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. |
Besides possessing richly designed mesopores and/or small macropores, CNTs/CNFs have high surface areas, which can provide many more active sites. Through structure design by using chemical vapor deposition (CVD), Lim et al. synthesized hierarchical carbon electrodes with highly aligned CNT fibrils, as shown in Fig. 13a and b.95 The hierarchical porous structure avoided blocking by the discharge products and was strong enough to remain stable after 100 cycles. Mitchell et al. fabricated CNF electrodes via atmospheric pressure CVD on porous anodized aluminum oxide (AAO) substrates coated with thin layers of Ta and Fe (Fig. 13c and d).163 The hollow CNFs were almost vertical, while the substrate surface was horizontal. Due to their low carbon packing and void volume, a high gravimetric energy density of ∼2500 Wh kg−1 at a power of up to ∼100 W kg−1 was achieved.
![]() | ||
Fig. 13 (a) SEM images at various magnifications of the CNT electrode before Pt coating and TEM image of a single CNT. (b) SEM and TEM images at various magnifications of the Pt/CNT electrode after Pt coating. (c) Cross-sectional (70°-tilt) SEM micrograph of the porous anodized aluminium oxide (AAO) filter after thin film deposition using electron beam evaporation. Inset: Schematic representation of the electrode after the deposition of metal thin films (Ta 30 nm, Fe 2 nm) on one side of the AAO filter. (d) Cross-sectional (70°-tilt) SEM image of the AAO filter after nanofiber growth. Inset: Schematic representation of the electrode after the catalyzed growth of carbon nanofibers. (e) TEM image of N-CNTs. (a and b) Reproduced with permission.95 Copyright 2013, the Royal Society of Chemistry. (c and d) Reproduced with permission.163 Copyright 2011, the Royal Society of Chemistry. (e) Reproduced with permission.99 Copyright 2011, Elsevier B.V. All rights reserved. |
N-doping engineering is one of the most popular methods to promote the performance of CNTs. N atoms can form five main doping structures, including pyridinic N, amine N, pyrrolic N, quaternary N and oxidized N with binding energies of 398–399 eV, 399–400 eV, 400–401 eV, 401–402 eV and 402–405 eV, respectively.96,97 Among them, pyridinic N seems to play a vital role in changing the structure of CNTs.98 Chen's group revealed the effects caused by doped Ni in CNTs.97 They synthesized CNTs and N-doped CNTs (CNx) via the floating catalyst chemical vapor deposition (FCCVD) method. The TEM images showed that CNx possessed a bamboo-like structure, with kinks along the tubes and abundant surface defects. The XPS spectra showed that N-doping was successful and the N atoms were mainly pyridinic, graphitic, and oxidized nitrogen. By comparing the rotating ring-disk electrode (RRDE) tests for ORR using Pt/CNTs and Pt/CNx, they found that the electrochemical active area of Pt/CNx was 1.6 times greater than that of Pt/CNTs and the specific activity of Pt/CNx was 1.34 times higher than that of Pt/CNTs. Thus, N-doping increases the electrochemical active area and promotes the intrinsic electrocatalytic activity toward ORR.
Li et al. first used N-doped CNTs (N-CNTs) as a cathode material for Li–O2 cells.99 The TEM plots showed that N-CNTs possessed typical bamboo-like structures (Fig. 13e), in accordance with previous research.97 The N atoms improved the performance of N-CNTs, the specific discharge capacity of which was 1.5 times higher than that of CNTs. Further, Lin et al. synthesized binder-free nickel foam-supported nitrogen-doped carbon nanotubes (N-CNTs@Ni) using the FCCVD method to avoid the decomposition of the binder and consequent formation of LiF (Fig. 14a–d).164 The binder-free material improved the capacity (1814 mAh g−1 at 0.05 mA cm−1, more than 2 times the capacity of the N-CNT-based Li–O2 cells) and rate performance of the Li–O2 cells.
![]() | ||
Fig. 14 (a and b) SEM images and (c and d) TEM images of the N-CNT@Ni electrode. Reproduced with permission.164 Copyright 2013, Elsevier B.V. All rights reserved. (e) Typical SEM images of C@LSSO NFs. Reproduced with permission.100 Copyright 2023, Jong Guk Kim, Yuseong Noh, Youngmin Kim. Published by Elsevier Ltd. (f) HAADF-STEM and (g) BF-STEM images of a single CNT@RuO2 structure (space bar: 20 nm). (h) Schematic of a single CNT@RuO2 structure. (f–h) Reproduced with permission.108 Copyright 2014, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. |
Heterostructure engineering is another common method employed for the modulation of CNTs/CNFs. Kim et al. synthesized pyrochlore LaSrSn2O7 nanoparticles (LSSO NPs) anchored on CNFs (C@LSSO NFs) as bifunctional catalysts (Fig. 14e).100 Compared to LSSO NPs and KB, the C@LSSO-NFs-based Li–O2 batteries had a lower overpotential (1.01 V) and higher round-trip efficiency (72.6%), which could be attributed to the synergistic effect of the intrinsic activity of LSSO NPs and electronic conductivity of CNFs. To avoid parasitic reactions, Jian's group designed a core–shell-structured CNT@RuO2 catalyst to prevent direct contact between carbon and the electrolyte, as shown in Fig. 14f and h.108 The RuO2 shell had a thickness of about 4 nm and could effectively wrap CNTs (Fig. 14h). The CNT@RuO2-based Li–O2 cells showed a low overpotential of 0.71 V and a high round-trip efficiency of about 79%, which can be attributed to the electronic conductivity of CNFs and high OER catalytic activity of RuO2.
2D CNSs/GNSs are also promising materials for application in the energy storage field due to their high surface areas and unique structures. Li et al. first used GNSs as a cathode material for Li–O2 cells.85 Compared to two other carbon materials, BP-2000 and Vulcan XC-72, GNSs possessed much more unsaturated carbon, which was active in reacting with O2, and consequently led to higher ORR activity, and more mesopores with a size in the range of 2 nm and 20 nm. At a current density of 75 mA g−1, GNSs showed a much higher energy density (8705.9 mAh g−1) than BP-2000 (1909.1 mAh g−1) and Vulcan XC-72 (1053.8 mAh g−1). Then, Li et al. used N-doped GNSs (N-GNSs) that possessed defects and functional groups including C–O, CO, and O
C–O, which led to a better electrocatalytic performance for ORR and a higher discharge capacity of 11
660 mAh g−1 because the discharge products prefer to grow around the defective sites with functional groups according to density functional theory (DFT) calculation.101 This was verified by the Tafel plots of N-GNSs and GNSs, with N-GNSs showing a smaller Tafel tangent than GNSs and 0.99 and 0.80 electrons transferred in ORR, respectively. Compared to GNSs, smaller-sized discharge products were observed on N-GNSs (Fig. 15a and b, respectively). In addition, Li's group explored the effects of S doping on GNSs.102 However, the effects of S doping were not ideal because of the much lower discharge capacity (4300 mAh g−1) than GNSs and other ordinary performance indexes.
![]() | ||
Fig. 15 (a) GNS and (b) N-GNS electrodes discharged for 1 h. reproduced with permission.101 Copyright 2012, Elsevier B.V. All rights reserved. (c) Schematic of the working mechanism of the Co-SA/N-C electrodes. Reproduced with permission.103 Copyright 2022, Peng Wang et al. (d) TEM image of Ru@PGE-2 and its SAED pattern (inset). Reproduced with permission.104 Copyright 2014, the American Chemical Society. (e) SEM image of MnO2 NT/NExG composite electrode. Reproduced with permission.105 Copyright 2012, Electrochemical Society. |
N-doped GNS/CNS-based composites with transition metal or transition metal oxide nanostructures have been proven to be effective catalysts for Li–O2 cells. For example, Wang et al. embedded Co nanoparticles in N-doped CNSs and found that the Co-N4 active sites improved the OER performance, as illustrated in Fig. 15c.103 The Li–O2 cell with this material achieved a low overpotential of 0.40 V, a high-rate discharge capacity of 11098 mAh g−1 at 1 A g−1, and excellent cyclability of 260 cycles at 400 mA g−1. Sun et al. added Ru nanocrystals to porous graphene oxide (Ru@nanoporous graphene), as shown in Fig. 15d, and found that the Ru nanocrystals could accelerate OER.104 Consequently, the Ru@nanoporous-graphene-based Li–O2 cells showed a low overpotential of ca. 0.355 V, a high reversible capacity of 17
700 mAh g−1, and long cycling life of 200 cycles at a limited capacity of 1000 mAh g−1. Park et al. combined N-doped thermally exfoliated graphene (NExG) with α-MnO2 nanotubes (NTs) to form NExG/α-MnO2 NTs as a cathode material.105 Owing to the defective sites caused by N doping and the thin 2D structure of NexG (Fig. 15e), the ORR was improved dramatically, with a max power density 32% larger than that of the MnO2-NT/Vulcan-carbon-based Li–O2.
In summary, various types of carbon materials have been employed as cathode catalysts for Li–O2 batteries, primarily due to their low cost and good electronic conductivity. Nevertheless, the instability of carbon materials impedes their practical application in Li–O2 batteries. McCloskey et al. first reported that Li2O2 (or LiO2) could react with ether electrolyte or carbon cathodes to form Li2CO3 and other carbonates, which led to a high overpotential and poor cycle life.106 The research by Gallant's group's on the chemical and morphological changes in vertically aligned carbon nanotube electrodes also confirm this.107 Further, Thotiyl et al. found that above 3.5 V (vs. Li/Li+), carbon tended to oxidatively decompose to Li2CO3 and lithium carboxylates (HCO2Li and CH3CO2Li), while under 3.5 V (vs. Li/Li+), carbon was relatively stable. However, carbon catalyzing the degradation of the electrolyte is inevitable.16 In addition, as mentioned before, 1O2 may exacerbate carbon erosion. Faced with these problems, even many well-designed carbon materials have poor stable cycle lives (typically <100 cycles without severe capacity losses).108–110 Thus, increasing the stability of carbon and finding electrolytes that can coexist with carbon materials without degradation reactions are the key to carbon-based Li–O2 batteries.
It is acknowledged that the Sabatier principle is useful in heterogeneous catalysis related to noble metals and alloys, which suggests that the interactions between the intermediates and the catalyst should be moderate, not too strong or weak. Conversely, the d-band model suggests that electrons in the s band are important but not decisive to adsorption; however, the electronic states of the d-band of each metal are different and the widths of the d-band of diverse metals vary. Also, the position of the d-band center varies in different metals. In general, if the d-band center is high, which means it is close to the Fermi level, the adsorption of intermediates by noble metals will be strong; the adsorption will be weak if the d-band center is low. The position of d-band center determines the adsorption energy (sometimes also called the binding energy). Therefore, adsorption is mainly determined by the electronic states of the d-band. It has been proven that a universal volcano-like relationship exists between the adsorption energy and catalytic activity, regardless of the reactants.217 Thus, the d-band model can be used to explain the catalytic activity of catalysts based on the Sabatier principle.
In the case of OER activity, the d-band model has been proven to be effective in explaining heterogeneous catalysis and electrochemistry.112 Noble metals have controllable d-band conditions, which can manipulate the interactions between them and the intermediates during the OER.98 As mentioned before, the intensity of the adsorption also needs to be moderate for a better catalysis performance. Further, the changes in the surface structures of noble metals also play a vital role in their OER activity. Thus, these two main factors should be considered simultaneously. For example, the calculated overpotential for Pt{111} is reported to be 1.62 V,36 possibly due to the strong adsorption of intermediates by Pt{111}.37,38 Song et al. found that the high index {411} has a larger surface energy, and therefore stronger adsorption than {111}, showing a lower OER overpotential of 1.20 V,36 which seemed contradictory to the d-band model. However, the rough facet with atomic steps provides high-density atoms on the step edge, kink, and ledge as active sites (Fig. 16a–d). The overall influence of these two factors makes the {411} facets more active than other facets, with a much lower overpotential of 0.51 V (Fig. 16e).
![]() | ||
Fig. 16 (a) TEM and (b) HRTEM images of the anisotropic Pt catalysts. The inset of (b) is an atomic model of the {411} facets exposed on the surface of the anisotropic Pt catalysts. (c) Enlarged HRTEM image and corresponding FFT pattern. (d) Schematic of the atomic steps on the high-index facet. (f) Atomic-resolved HAADF-STEM image of TOH Au NCs taken along the [110] direction, showing the (441) surface, (110) terraces, and (001) step. Top-left corner inset of (f) is the corresponding indexed FFT pattern along the [110] zone axis. The middle inset of (f) is the simulated atomic arrangement of {110} facets, showing the projected {110} {111}, {221}, {331}, and {441} crystal planes along the [110] direction. (g) Charge–discharge curves of the cubic Au NCs@SP (red solid line), T-OCT Au NCs@SP (blue solid line), TOH Au NCs@SP (green solid line) and bare SP electrodes (black solid line) at 100 mA g−1 in the first cycle. (a–e) Reproduced with permission.36 Copyright 2018, the American Chemical Society. (f-g) Reproduced with permission.41 Copyright 2015, Dawei Su et al. |
Su et al. researched the variable index facets of Au and found that the {441} high-index facet has a high density of stepped surface atoms compared with the {111} facet, as shown in Fig. 16f.41 Cubic gold (Au) NCs enclosed by {100} facets, truncated octahedral Au NCs enclosed by {100} and {110} facets, and trisoctahedral (TOH) Au NCs enclosed by 24 high-index {441} facets loaded on Super-P (SP), respectively, were tested in Li–O2 cells (Fig. 16g). Among them, it was found that TOH Au NCs@SP showed the lowest overpotential of 0.95 V and the largest discharge capacity of 20298 mAh g−1. DFT calculations showed that the reaction energy between the Au, Li, and O atoms decreases as the surface energy increases. The high index {441} facet has the highest surface energy of 2.55 J m−2 and the lowest reaction energy with Li and O atoms (1.64 J m−2). The results of this work indicated that the high index facet of the same noble metal has stepped atoms with higher catalytic activity, in accordance with the research by Song's group.
Interestingly, an earlier work about Pt3Co reported that the introduction of Co atoms could change the electron states of Pt and reduce the ηOER of Pt{111} from 1.62 V to 1.13 V (Fig. 17a–c).37 The DFT calculations suggested that the reduction in ηOER was related to the decrease in the adsorption strength of LiO2 on the outermost Pt catalytic sites, with the LiO2 adsorption strength changing from 3.65 eV to 3.01 eV, which helped prove the reasonability of the former study.114
![]() | ||
Fig. 17 Calculated energy diagrams for Li-ORR/OER on (a) Pt(111), (b) Co(0001), and (c) Pt3Co(111), along with the optimized structures of LixO2 (x = 1, 2 and 4) adsorbed on the surfaces. (d) First discharge/charge profiles of carbon at 85 mA gcarbon−1 and of Au/C, Pt/C, and PtAu/C at 100 mA gcarbon−1 in an Li–O2 cell. Reproduced with permission.39 Copyright 2010, the American Chemical Society. (e) PDOS of Pt 5d orbitals for PtRu and PtAu catalysts. (f) Schematic of the bond formation between the adsorbate (LiO2) valence bands and the d states of PtM (M = Au and Ru). From PtRu to PtAu, the upshift in the d-band center elevates the corresponding antibonding orbits. (g) and (h) Free energy curves at various potentials for PtAu (g) and PtRu (h) cathodes. The embedded graphics in (g) and (h) show the refined crystal structures of the PtAu and PtRu catalysts with specific adsorbates at the corresponding steps. (a–c) Reproduced with permission.37 Copyright 2014, Byung Gon Kim et al. (e–h) Reproduced with permission.35 © 2022, Wiley-VCH GmbH. |
Due to its outstanding ORR catalytic activity among the noble metals, Pt is often alloyed with other noble or transition metals. PtAu nanoparticles were confirmed to be an efficient Li–O2 catalyst.39 PtAu/C had an average charge potential of 3.6 V, which is slightly lower than that of Pt/C and much lower than that of Au/C. Also, the discharge potential of PtAu/C was slightly lower than that of Au/C and much lower than that of Pt/C (Fig. 17d). Thus, the results suggested that the surface Au and Pt atoms of bifunctional PtAu are primarily responsible for its ORR and OER kinetics in Li−O2 cells, respectively. PtAu alloy has been further explored by designing the eg occupancy of Pt to alter its overpotentials.35 Compared to PtRu, PtAu had a smaller eg occupancy, which indicated that the number of eg electrons in PtAu (2.84) was less than that in PtRu (2.96) and led to an increase in the d-band center (Fig. 17e and f), thus enhancing its adsorption of LiO2. The Gibbs free energies for the two catalysts were calculated during the OER processes. The energy barrier of PtAu (0.842 eV) was less than that of PtRu (1.01 eV) because of the stronger adsorption of LiO2 on PtAu. Therefore, the calculated OER overpotential of PtAu was smaller than that of PtRu (Fig. 17g and h), respectively, in accordance with the experiment.
In addition, PtIr alloy has been researched for use in high-performance Li–O2 cells.38 Multipod-like PtIr was synthesized with a length of ca. 45 nm and a width of ca. 10 nm. At 1 mA g−1 when the capacity was set to 1000 mAh g−1, the PtIr multipods displayed much lower OER and ORR overpotentials of 0.33 V and 0.11 V, respectively, than that of Pt nanocrystals (1.02 V and 0.17 V). Due to the higher electronegativity of Pt than Ir, the charge state of Pt becomes more negative, leading to lower Lewis acidity. The Pt atoms with higher electron density and lower Lewis acidity showed weaker adsorption for LiO2 because of the downshift in the Pt d-band center. Other alloys have also been explored as Li–O2 catalysts, such as AgPd and PdCu.115–117
Recently, high-entropy materials (HEMs) have been used as Li–O2 catalysts, including HEAs, high-entropy oxides, and high-entropy sulfides.215 In this part, we only focus on HEAs.
Based on the Sabatier principle and the tunable d-band structure of HEAs due to the high entropy effect, Tian et al. used several HEAs to explore the detailed mechanism of their d-band structure and catalytic activity.218 The positions of the d-band center are shown in Fig. 18a, with the sequence from high to low being HEA > PtIr > HEAPtIr > HEAIr > HEAPt (here, HEA is FeCoNiMn alloy), which indicates that the order of the adsorption energies for the intermediates also follows this trend. Then, the five catalysts were tested in Li–O2 cells. As expected, the HEAPtIr-based LOB exhibited a super low OER overpotential of 0.38 V (Fig. 18b), which is much lower than that of the other four catalysts. Furthermore, though DFT calculations, the order of the binding energies for LiO2 and Li2O2 corresponds to the positions of the d-band centers and the conversion efficiency shows a volcano-like relationship with the binding energy (Fig. 18c and d). As shown in Fig. 18e and f, when the d-band center is too high, the energy of the antibonding orbital is higher than the Fermi level (EF), which no electrons fill. On the contrary, more electrons fill the bonding orbital, making adsorption too strong because of the stabilization effect of the bonding orbital. In the case of FeCoNiMn, the adsorption of LiO2 and Li2O2 is too strong on its surface. Thus, LiO2 cannot transform into Li2O2 sufficiently, and it is difficult for Li2O2 to decompose into O2. When the d-band center is too low, the energy of the antibonding orbital is lower than EF, and then the electrons fill both the antibonding and bonding orbitals; thus, the proportion of electrons in the antibonding orbital increases, which offsets the stabilization effect of the bonding orbital to some extent, making adsorption too weak. In the case of FeCoNiMnPt, the adsorption of LiO2 is too weak on its surface. Thus, the conversion of LiO2 to Li2O2 is sluggish, and the slow electronic transfer caused by the weak interaction between the catalyst and O makes it difficult for Li2O2 on its surface to decompose, leading to surface passivation. Therefore, when the energy of the d-band center is moderate, the energies of the antibonding and bonding orbitals and electrons filling them are suitable, making adsorption ideal.
![]() | ||
Fig. 18 (a) d-Band centers of HEA, PtIr, HEAPtIr, HEAIr, and HEAPt. (b) Corresponding overpotentials at 500 mAh g−1 of the five alloy materials at 200 mA g−1 with a limited capacity of 1000 mAh g−1. (c) Binding energies between LiO2 and as-designed catalysts (HEA, PtIr, HEAPtIr, HEAIr, and HEAPt), and the energy conversion efficiency of LOBs with different cathodes. (d) Binding energies of LiO2 and Li2O2 on as-designed catalysts. (e) Orbital interactions between LiO2/Li2O2 and catalysts with different d-band centers and (f) corresponding catalytic effects. Reproduced with permission.218 Copyright 2023, Wiley-VCH GmbH. |
Tao et al. developed a universal low-temperature method for synthesizing subnanometer ribbons (SNRs), with a single layer possesses a thickness of 0.8 nm and containing up to 8 elements, using Ag nanowires as a templete.219 PtPdIrRuAuAg SNRs/C (PtPdIrRuAuAg deposited on commercial carbon black) was used as an Li–O2 catalyst. At 100 mA g−1 with a limited capacity of 1000 mAh g−1, the Li–O2 cell with this catalyst showed a low OER overpotential of only 0.49 V. Even when the current increased to 1 A g−1, an OER overpotential of 0.75 V was observed. Also, the Li–O2 battery with this catalyst could run for 100 stable cycles at 500 mA g−1 with a limited capacity of 1000 mAh g−1, and its terminal charge voltage was lower than about 3.8 V.
In addition, HEA nanoparticles (NPs) have been well-designed recently. Wang et al. first developed a continuous “droplet-to-particle” method to fabricate hollow HEA NPs.220 The hollow structure can save 60% materials but provide the same number of active sites on its surface. Zhang et al. synthesized a core-satellite HEA@Pt heterogeneous catalyst with PtRuFeCoNi HEA NPs as the core and Pt dendrites on its surface.221 As mentioned before, changes in the electronic structure of heterointerfaces can improve the charge transfer and expose the active sites. In detail, exposed electron-rich and electron-deficient sites prompt ORR and OER kinetics, respectively. Due to the lower electronegativity of the other elements than Pt, the d-band center moved down, and then the adsorption of the adsorbates HEA@Pt was weakened compared to Pt. Owing to the moderate adsorption energy for LiO2 being between that of HEA and Pt, the HEA@Pt-based LOB showed the maximum capacity of 8400 mAh g−1 at 100 mA g−1 with a low overpotential of only 0.46 V, a small overpotential of only 0.37 V at 100 mA g−1 with a fixed capacity of 1000 mAh g−1, and 210 cycles, among which the terminal voltages of approximately the first 180 cycles were very stable. Wu et al. synthesized PtFeCoNiCu NPs loaded on reduced graphene oxide (PtFeCoNiCu@rGO) through a high-temperature annealing route.222 Due to the rich pores and large specific surface area of rGO as well as high electronic conductivity and lattice distortion effect of the HEA NPs, the PtFeCoNiCu@rGO-based LOB showed a high maximum capacity of 13949 mA g−1 at 100 mA g−1, a low overpotential of 0.77 V and 148 cycles at 100 mA g−1 with a limited capacity of 500 mAh g−1. By removing the nonnoble elements on the surface of PtFeCoNiCuMo HEA, causing some free Pt single atoms to be caught by defect-rich CNTs, Li et al. fabricated amorphous Pt-skin-coated HEA nanoparticles paired with Pt single atoms (HEA@Pt-PtSAs) on CNTs.223 Toroid-like Li2O2 tended to form on the pure Pt cathode and film-like Li2O2 tended to form on HEA. Due to the synergistic catalysis of Pt and HEA, needle-like Li2O2 was observed on HEA@Pt-PtSAs. Owing to its suitable adsorption energy, the HEA@Pt-PtSAs-based LOB showed an excellent performance including an ultralow total overpotential of 0.3 V at 200 mA g−1 with a fixed capacity of 1000 mAh g−1, 470 cycles at 1000 mA g−1 with a fixed capacity of 1000 mAh g−1, and a maximum capacity of 13
116 mAh g−1 at 200 mA g−1. Furthermore, no clear parasitic reactions were detected, revealing that HEA@Pt-PtSAs had an obvious inhibitory effect on them.
Compared to traditional alloys, HEAs can contain fewer noble metals by increasing the content of nonnoble metals, making HEAs more economical. In future studies, the development of more HEAs that show higher stability and catalytic activity than traditional alloys will be a great contribution to the commercialization of Li–O2 batteries.
Noble metals and their oxides have exhibited potential for accelerating ORR and OER in Li–O2 cells, respectively. Thus, the combination of noble metals and their oxides may be a possible way to find bifunctional catalysts. The mechanistic understanding of noble-metal-based catalysts has been systematically established through the d-band center theory combined with density functional theory (DFT) calculations.35,37,38,112,114,115,126 Nevertheless, the high cost of noble metals and their oxides has severely hindered their practical usage on a large scale. Alternatively, although noble-metal single-atom catalysts (SACs) have relatively lower cost, the agglomeration of their atoms can diminish their catalytic activity.158 Consequently, the design of composite catalysts with noble-metal-free catalysts, such as transition metals and carbon materials, as the main bodies decorated by a small amount of noble metals and/or their oxides, such as doping and core-shell structures, is one possible method to achieve suitable catalysts.
MOFs were first applied as Li–O2 cathode materials in 2014, and it was found that their rich and even pores could gather at most 18 times the O2 density than that in the atmosphere at 273 K.130 A high capacity of 9420 mAh g−1 was gained in Mn-MOF-74-based Li–O2 cells, which was more than 4-times larger than that of super-P-based Li–O2 cells.
After that, more MOF-derived materials, such as atom-doped MOFs and their composites (such as composites with MXenes), have been tested as Li–O2 cathodes because of the intrinsic poor electron conductivity and low stability of MOFs.27,42,129,131–134 To address these problems, an Fe(III)-MOF-5 carbonization-derived conductive carbon structure was fabricated through thermal annealing.133 Covered by a thin carbon layer, the hierarchical mesoporous ZnO/ZnFe2O4/C (ZZFC) nanocages obtained at the sintering temperature at 500 °C had a balance of good electronic conductivity and proper porosity with the average pore size of 12.2 nm (Fig. 19a–d). Therefore, the stability was enhanced, as proven by a series of LSV and NMR experiments, where the TEGDME-based electrolyte hardly decomposed in the presence of the ZZFC electrode (Fig. 19e–g). Furthermore, the ZZFC-based Li–O2 batteries could run for 15 cycles at the fixed capacity of 5000 mAh g−1, 12 more cycles than the super-P-based Li–O2 batteries (Fig. 19h and i).
![]() | ||
Fig. 19 (a) SEM images of ZZFC. (b) TEM images of ZZFC. (c and d) High-resolution TEM images of ZZFC. (e and f) LSV curves measured on a rotating glassy carbon electrode coated with various air-electrode materials (the sweep rate was set at 0.01 V s−1 without rotation; panel f is the magnified image of panel e). (g) Magnified 1H NMR spectra of the electrolytes with various cathodes after cycling (the cathode catalyst in each sample is (1) ZZFC, (2) Co3O4, (3) MnO2, and (4) 5% PtC; all spectra are presented after magnified). (5) Non-magnified 1H NMR spectra of 1.0 mol L–1 LiTFSI in TEGDME solution. (h) Fixed capacity galvanostatic discharge/charge curves of super P cathode (90% super P + 10% PVDF). (i) Galvanostatic discharge/charge curves of ZZFC cathode in a “deep discharge + fixed capacity” mode. The cell was first discharged to 8000 mAh g−1, and then cycled at a fixed capacity of 5000 mAh g−1. (a–i) Reproduced with permission.133 Copyright 2015, the American Chemical Society. |
Flower-like Co and Ni atom-doped MOFs (CoNi-MOFs) exhibited higher electrochemical catalytic performance than Ni-MOF due to the change in the adsorption of the LiO2 intermediate and the enlarged pore size of CoNi-MOFs (Fig. 20e and f).134 The pores with larger sizes permitted the smoother transportation of Li+ and O2 as well as the greater deposition of Li2O2, which accelerated the electron transfer and promoted the capacity. In comparison to Ni-MOFs, where only film-like Li2O2 was formed, film-like Li2O2 through the surface path and toroid-like Li2O2 through the solution path were formed on the surface of CoNi-MOFs after discharge to 600 mAh g−1 (Fig. 20a and b). As the discharge capacity increased to 2000 mAh g−1, toroid-like Li2O2 remained, which could better ease the congestion of the Li+ and O2 diffusion channels than the film-like Li2O2 (Fig. 20c). Also, after charging, Li2O2 completely disappeared (Fig. 20d), which indicated the excellent cyclic ability of CoNi-MOFs.
![]() | ||
Fig. 20 SEM images of the CoNi-MOF cathodes: (a) fresh cathode, (b) discharge to 600 mAh g−1, (c) discharged to 2000 mAh g−1 and (d) recharged to 2000 mAh g−1. (e) SEM images of CoNi-MOF (the inset is an enlarged SEM image). (f) Pore size distribution curve of Ni-MOF and CoNi-MOF. (a-f) Reproduced with permission.134 Copyright 2024, the American Chemical Society. |
Interestingly, a 3D printed MOF-derived hierarchically porous framework was used as an Li–O2 cathode material.135 Applying one-step calcination, Co nanoparticles assembled in N-doped mesoporous carbon flakes (3DP-NC-Co) were decomposed from the Co-MOF gained from 3D printing (3DP-Co-MOF), as shown in Fig. 21a. The 3DP-Co-MOF had large pores (ca. 200 nm), and its crystal blocks had micron-sized leaf structures (Fig. 21d–f). 3DP-NC-Co retained the framework of 3DP-Co-MOF (Fig. 21g–i). Compared to the Co-MOF-derived powder (NC-Co), 3DP-NC-Co showed a higher specific surface area (640 m2 g−1 vs. 433 m2 g−1) and a larger meso- and microporous volume (1.09 cm3 g−1 vs. 0.21 cm3 g−1), respectively, which illustrated the importance of the hierarchical porous network geometry. More Li2O2 nanoparticles could deposit in the micron-sized pores to improve the capacity and achieve good contact with the cathode to enable the easier decomposition of Li2O2 during charging, which led to lower overpotentials for 3DP-NC-Co than NC-Co randomly packed on carbon paper (RP-NC-Co) and NC-Co directly grown on carbon paper (CP-NC-Co), as shown in Fig. 21b.
![]() | ||
Fig. 21 (a) Diagram showing the preparation of a 3DP Co-MOF-derived framework. (b) Comparison of the charge overpotential with the marked cathodes at a limited capacity of 1 mAh and current density of 0.1 mA cm−2. (c) Discharge and charge profiles of the Li–O2 battery with and without illumination. (d–f) Typical characterization of 3DP-Co-MOF. (d and e) SEM images. (f) TEM image. (g–i) Typical characterization of 3DP-NC-Co. (g and h) SEM images. (i) TEM image. (a and b) and (d–i) reproduced with permission.135 © 2018, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. (c) Reproduced with permission.136 © 2024, Wiley-VCH GmbH. |
Photoactive MOFs have been successfully employed in photo-assisted Li–O2 batteries.137 Because of the impacts of photoelectrons and holes produced by photocatalysts, the discharge voltage can surpass the thermodynamic limit value (2.96 V),138 for example, 3.42 V in Fig. 21c. Recently, Fe3+ ions were doped in UiO-66 (Fe-UiO-66) as a bifunctional Li–O2 cathode catalyst.136 Fe3+ may act as an electron donor and oxidation site, and the charge was transferred from Fe3+ to Zr-oxo clusters, expanding the visible light absorption. Then, the photogenerated electrons accelerated the reduction of O2 to O2− radicals, expediting the ORR process. A thin Li2O2 film was formed after discharge in the presence of light, which is much more conducive to the charge process than bulk Li2O2 without light because of the larger contact area between Fe-UiO-66 and the Li2O2 film. In addition, there was an electronic energy input during exposure to light. Therefore, when employed in Li–O2 cells, Fe-UiO-66 in the presence of light exhibited a much lower overall overpotential of 0.14 V than Fe-UiO-66 in the dark (Fig. 21c) and 500 stable cycles.
As one of the emerging materials used as Li–O2 cathodes, MOF-based materials exhibit significant promise due to their well-defined porous structures, high specific surface areas and other inherent advantages. Furthermore, to improve their electronic conductivity and stability, the design of hierarchical structures, the optimization of their porosity, the coating of a carbon layer, metal atom doping, and formation of composites with high conductive carbon materials are widely applied. Research has explored the relationship between the quantitative electronic structure and the OER activity of bimetallic Ni-based MOFs in KOH electrolyte.139 However, analogous studies on MOF-based materials in the organic electrolytes or ionic liquids (ILs) commonly used in Li–O2 batteries remain limited. In addition, reducing the synthesis costs of MOF-based materials is also a critical challenge for their practical application.131
Doping engineering, vacancy/defect engineering, crystal plane effects, spin engineering and amorphization strategy have been proven to be some of the most common modulation methods for TMCs, as illustrated by the examples of Co3O4, ZnCo2O4 and NiCo2O4. Due to their wide range of advantages, including high specific surface area, quantum size effect, and easy heterostructure engineering with supports, nano-sized TMCs have also been widely applied in Li–O2 batteries. MXenes, as common supports for TMCs, were introduced in detail because of their excellent electronic conductivity and high surface areas.
Carbon-based materials play an important role in LOBs due to their rich resources, suitable prices, high electronic conductivity, etc. Much attention was paid to the importance of pores, especially mesopores, in terms of O2 diffusion and improving the Li–O2 capacity herein. Then, 1D CNTs/CNFs were reported through modulation strategies, among which N-doping was analyzed preferentially due to the great changes in nanotubes caused by N atoms, followed by a discussion on 2D CNSs/GNSs.
In general, noble metals exhibit superior ORR activity, while their oxides demonstrate higher activity in OER. The d-band model and DFT calculations serve as powerful tools for assessing the catalytic activities of noble metal-based materials for ORR and OER. Alloying has been demonstrated to be an effective strategy to improve the OER activity of noble metals. Recently, high-entropy alloys have gained much attention due to their peculiar effects not shown by traditional alloys. However, the high costs of noble metals hinder their practical use.
Recently, MOF-based materials with rich pores and high specific surface areas have begun to demonstrate potential as Li–O2 cathodes. Notably, photoactive MOFs exhibit the capability to dramatically reduce the overall overpotential, while maintaining long cycle lives.
However, there are common obstacles for all types of cathode catalysts, hindering the commercialization of LOBs, including Li dendrites, parasitic reactions, and unclear catalytic mechanisms.
Therefore, great efforts need to be devoted to solving these technological difficulties. The specific strategies are as follows:
1. Owing to the unclear catalytic mechanism, the unclear role of the electrolyte, and the controversial oxygen reduction reaction mechanism, the development of cathode catalysts is unsatisfactory. Also, the current analysis of the advantages of new catalysts mainly depends on theoretical calculations, such as DFT calculations.
In situ probing techniques, such as in situ electrochemical quartz microbalance (EQCM) and fluorescence microscopy, are helpful for detecting the changes in the composition and structure of the electrodes and electrolyte.140–142 For example, in situ TEM was used for detecting the decomposition of Li2O2.140 In situ electron paramagnetic resonance (EPR) can be used to monitor the formation of Li dendrites.143
However, most of the in situ techniques cannot combine the analysis of the evolution in electrochemical performance and chemical compositional and structural changes. Recently, in situ electrochemical impedance spectroscopy (EIS) combined with the distribution of relaxation times (DRT) and the distribution of capacitive times (DCT) was found to be a viable solution to this problem.144 Thus, the exploration of more effective in situ probing techniques is a direction to reveal the unclear mechanisms.
2. The research on universal descriptors for catalysts of the same category and the foundation of models are important because of their great roles in evaluating and forecasting new materials. For example, the eg occupancy, d-band center/model, and metal–O covalency have been widely used as descriptors for transition metal oxides.145–147 Recently, a new descriptor, WF, for 2D transition metal oxides has been reported to evaluate the LiO2 adsorption energy and overall catalytic properties.148 The descriptors and models can provide a better understanding of emerging cathode catalyst development.
3. Attention also needs to be paid to the control of the decomposition of the electrolyte and cathode materials in the presence of Li2O2, which means they should remain stable in the presence of each other. Generally, carbon-based materials can catalyze the degradation of the electrolyte, such as organic carbonate-based electrolytes and ethers.16,149 Thus, it is necessary to find stable and high-donor-number electrolytes that are not only resistant to side reactions but also able to achieve higher capacities.
4. The addition of soluble catalysts, including RMs in pairs, ligand mediators224–226 (LMs) that can change the solvation structure of ions or coordinate with them, multifunctional additives,227–230 soluble photocatalysts used for photos-assisted Li–O2 batteries,231 and anti-superoxide disproportionators,232,233 to the electrolyte has been attracting more attention recently. It has been reported that RMs can enhance the cycling ability and make carbon materials more stable.150,151 Ligand mediators form complexes with ions or molecules such as O2,224,226 Li+,225 and O2−,228,229 which prompt ORR and even make changes in the final discharge product from Li2O2 to LiOH possible.224 A low LiOH decomposition voltage effectively lowers the OER overpotential. In this case, multifunctional catalysts such as carbonized polymer dots (CPD)230 play two or three roles, including serving as RMs and/or LMs,227–229 protecting the Li anode,229,230 increasing the stability of the electrolyte,230 and accelerating the superoxide disproportionation reaction.227 Notably, to relieve the slow dynamics of Li2O2 decomposition, iridium (Ir) single-atom-based porous organic polymers (Ir/AP-POP), as anti-superoxide disproportionation, have been developed to make nonsolid LiO2 the final discharge product.232 An Ir/AP-POP-based Li–O2 battery with a carbon cloth cathode showed an ultralow OER overpotential of 0.03 V, an ultrahigh maximum capacity of 12.8 mAh and a long cyclic life of over 700 h. In short, the development of more stable soluble catalysts is favorable for the experimental exploration of Li–O2 batteries.
5. The Li anode should be protected from erosion due to H2O and CO2 and the growth of Li dendrites. Thus far, protective layers such as a film induced via an in situ electrochemical process,152 an organic–inorganic hybrid layer (OIHL),153 and a layer comprised of a conductive polymer and AlF3 particles154 have been used to protect the Li anode. The research on protective layers is meaningful for the suppression of the growth of Li dendrites, Li anode erosion, and electrolyte decomposition.
6. Machine learning can be a powerful tool for improving the performance of Li–O2 batteries. Thus far, in the field of Li–O2 batteries, machine learning has been used in developing a dual-solvent electrolyte possessing anion-induced ion-solvent-coordinated structures, which contribute to the formation of a stable inorganic-rich solid electrolyte interphase (SEI), and consequently a better performance in LOBs,155 finding factors that lead to a better performance based on thousands of experimental data,156 and predicting the chemical stability of organic materials including many electrolyte solvents and RMs.157
In the future, AI models that can anticipate new high-performance materials may be developed and make a great difference in this field based on big data on Li–O2 batteries including their working mechanism, basic systematic knowledge about common materials and experimental outcomes.
Through the implementation of these strategies and considering the urgent demand for high-energy-density batteries, lithium–oxygen batteries are anticipated to be promising candidates.
This journal is © The Royal Society of Chemistry 2025 |