Takahiro
Ami
a,
Kouki
Oka
*abc,
Hitoshi
Kasai
a and
Tatsuo
Kimura
*d
aInstitute of Multidisciplinary Research for Advanced Materials, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai, Miyagi 980-8577, Japan
bCarbon Recycling Energy Research Center, Ibaraki University, 4-12-1 Nakanarusawa, Hitachi, Ibaraki 316-8511, Japan
cDeuterium Science Research Unit, Center for the Promotion of Interdisciplinary Education and Research, Kyoto University, Yoshida, Sakyo-ku, Kyoto 606-8501, Japan
dNational Institute of Advanced Industrial Science and Technology (AIST), Sakurazaka, Moriyama-ku, Nagoya 463-8560, Japan. E-mail: t-kimura@aist.go.jp
First published on 3rd December 2024
The development of electrocatalysts for the oxygen evolution reaction (OER) is one of the most critical issues for improving the efficiency of electrochemical water-splitting, which can produce green hydrogen energy without CO2 emissions. This review outlines the advances in the precise design of inorganic- and organic-based porous electrocatalysts, which are designed by various strategies, to catalyze the OER in the electrolytic cycle for efficient water-splitting. For developing high-performance electrocatalysts with low overpotentials, it is important to design a chemical composition that optimizes binding energy for an intermediate in the OER and allows the easy access of reactants to active sites depending on the porosity of electrocatalysts. Porous structures give us the positive opportunity to increase the accessible surface of active sites and effective diffusion of targeting molecules, which is potentially one of the guidelines for developing active electrocatalysts. Further modification of the frameworks is also powerful for tailoring the function of pore surfaces and the environment of inner spaces. Designable organic molecules can also be embedded inside inorganic- and organic-based frameworks. According to chemical composition (inorganic and organic), nanostructure (crystalline and amorphous) and additional modification (metal doping and organic design) of porous electrocatalysts, the current status of resultant OER performance is surveyed with some problems that should be solved for improving the OER activity. The remarkable progress in OER electrocatalysts is also introduced for demonstrating the bifunctional hydrogen evolution reaction (HER) and for utilizing seawater.
Earth-abundant base metals, such as iron (Fe), cobalt (Co) and nickel (Ni), that are produced more than 1000 times than that of Ir and Ru per year,10,11 exhibit excellent OER activity. Such base metal derivative materials have been investigated aggressively for the development of inexpensive OER electrocatalysts with low overpotentials.12–14 In addition, electrochemical water-splitting involves high costs owing to the need for expensive equipment and large amount of freshwater to provide enough hydrogen. Hence, in recent years, bifunctional OER-HER and innovative OER electrocatalysts are strongly recommended to reduce equipment costs and/or to explore the possibility of utilizing unlimited seawater. Many studies have focused on the nanostructural design of base metal derivative electrocatalysts to maximize OER activity.15–17 (See Fig. 1) Highly porous structures are one of the most attractive structural features, having a great potential to enhance OER activity depending on many active sites and high diffusivity arising from high specific surface area, large pore volume and tunable pore shape/size.18–20 Further design of porous surfaces is possible for tailoring not only major physical properties but also functions of pore surfaces21,22 and the environment inside pores23,24 by modification with constituent molecules, feasibly enhancing uptake potential and unique interaction at the resultant pore surfaces against specific substances.
In our recent work, organic-based porous materials have been designed by controlling organic components, and surface environments are assessed by the introduction of substituents into constituent molecules.25–29 Basic and highly hydrophobic pore surfaces are designed with the fabrication of all-organic porous materials, enabling selective CO2 adsorption30 and improving the proton conductivity.31 Even in the application of OER electrocatalysts, porous materials can adjust properties (e.g., wettability, electron conductivity, and proton conductivity) against the reactants and products. Organic-based porous materials enable the precise control of their structures to confirm the effect of specific properties on OER activity and then help elucidate the OER mechanism on the microscopic scale. Inorganic-based porous materials can tune the composition of active metal species and then help elucidate the OER mechanism on the macroscopic scale. From this viewpoint, it is crucial to summarize the effects of compositions, nanostructures and properties on OER activities based on inorganic- and organic-based porous materials to systematically understand the OER mechanism as well as the material design of OER electrocatalysts.
Many reviews have been reported to date in the field of OER electrocatalysts and most of them focus on the material composition. This feature article focuses on a variety of porous electrocatalysts to maximize the OER performance in addition to the compositional design of the latest OER electrocatalysts for revealing the effects of the chemical composition of electrocatalysts, the design of porous structures and the additional modification of organic molecule. Besides, we survey the significant advances in OER-HER bifunctional electrocatalysts to catalyze not only OER at the anode but also HER at the cathode as well as those OER electrocatalysts that enable the use of seawater and thus provide the guidelines to design high-performance porous electrodes for the development of OER electrocatalysts (see Fig. 1).
The Gibbs free energy change (ΔG°) for the electrochemical water-splitting is +237.2 kJ mol−1 under the standard conditions (activities of all species; 1, ambient pressure; 1 atm, temperature; 298.15 K).32 ΔG° can be converted to the standard potential (E°, theoretical cell voltage of the electrolytic reaction) for the entire equilibrium reaction by eqn (1):
| ΔG° = −nFE° | (1) |
The electrochemical water-splitting reaction consists of HER at the cathode and OER at the anode, as described by reactions (2) and (3) in Fig. 2 under acidic conditions.33,34 Reaction (2) represents the electrode reaction at a standard hydrogen electrode (SHE) where E° is 0.00 V (vs.SHE). ΔG° of reaction (3) is −237.2 kJ mol−1,32 and E° is calculated to be +1.23 V (vs.SHE) by eqn (1).
HER produces hydrogen by the reduction of water, like reaction (4) in Fig. 2, not protons, under neutral and alkaline conditions.33,34 ΔG° of reaction (4) is +159.8 kJ mol−1,32 and E° is calculated to be −0.828 V (vs.SHE) by eqn (1). The OER corresponding to reaction (4) (reaction (5) in Fig. 2) is represented by the oxidation of water (reaction (3)) and the oxidation of the hydroxide anion produced by reaction (4). E° of reaction (5) is calculated to be +0.401 V (vs.SHE) based on the ΔG° of reaction (3) (−237.2 kJ mol−1). The reduction reaction (Ox + ne− ⇄ Red), in which the oxidized species (Ox) forms the reduced species (Red), is in electrochemical equilibrium and follows eqn (2) (Nernst's equation).
![]() | (2) |
E is the equilibrium potential, and R, T, and ai are the universal gas constant, the temperature, and the activity of species i. By applying eqn (2) to HER (reactions (2) and (4)), following eqn (3) is obtained in both cases.35
| E = (−0.059 × pH) V (vs.SHE) | (3) |
Furthermore, by applying eqn (2) to OER (reactions (3) and (5)), eqn (4) is obtained in both cases.35
| E = (+1.23 − 0.059 × pH) V (vs.SHE) | (4) |
Therefore, the theoretical (minimum) cell voltage for electrochemical water-splitting is 1.23 V under all pH conditions.
Practical water electrolyzers require a working voltage higher than the theoretical cell voltage for electrochemical water-splitting (1.23 V), which is attributed to the overpotentials on both HER and OER electrodes.36 In particular, the OER, through a complex four-electron reaction with multiple intermediates, proceeds with a much larger overpotential than the two-electron HER.3–7,37,38 To accelerate the overall electrochemical water-splitting reaction, highly active OER electrocatalysts should be required to minimize the overpotential. The theoretical overpotential of OER is confirmed by following the oxygen adsorption energy of the catalyst surface.36 A strong interaction between the reaction intermediates and the catalyst surface needs excess energy to desorb oxygen (O2) molecules. If such an interaction is too weak, the OER does not proceed efficiently. To minimize the overpotential for OER, appropriate control of the oxygen adsorption energy on the catalyst surface is quite important.39
It is rational to consider that OER proceeds through two possible reaction mechanisms (see Fig. 3) called the adsorbate evolution mechanism and lattice-oxygen oxidation mechanism.14 Both mechanisms involve four proton–electron transfer reactions and lead to the production of gaseous O2 from water (H2O) on the active site and through lattice oxygen, respectively. Especially, the adsorbate evolution mechanism is completed with four steps, such as reactions (6)–(9) under acidic conditions (Fig. 3a) and reactions (10)–(13) under neutral/alkaline conditions (Fig. 3b). In this mechanism, OER activity is limited by the scaling relationship of the adsorption energy between *OH/*OOH intermediates.40 The theoretical minimum overpotential is considered to be 370 mV for the adsorbate evolution mechanism.41 Besides, two mechanisms via different active centers have been proposed for the lattice-oxygen oxidation mechanism.42 The lattice oxygen is the active center in this mechanism (Fig. 3c). The lattice oxygen accepts the OH− directly through a nucleophilic attack to form the *OOH intermediates. Then, the release of O2 creates oxygen vacancy sites (□), which then adsorb OH− to form *OH intermediates. The metal site acts as an active center to adsorb OH− and causes a deprotonation reaction (Fig. 3d). Surface reconstruction leads to the formation of *OOH intermediates through the combination of *O species and activated lattice oxygen, releasing O2. In both mechanisms, the lattice-oxygen oxidation mechanism can effectively avoid the adsorption of *OH/*OOH, and thus this mechanism possibly proceeds with smaller overpotentials than the theoretical minimum overpotential of 370 mV for the adsorbate evolution mechanism.43 The OER mechanism mainly depends on the composition, structure, and crystallinity of electrocatalysts, but the reaction conditions predominant by the lattice-oxygen oxidation mechanism rather than the adsorbate evolution mechanism are still open for discussion.
To reduce the activation energy barrier during the OER cycle, the surface property of electrocatalysts, depending on the composition and structure, has been designed using unique methods. Dai et al. succeeded in fabricating an electrochemically active Co3O4 nanosheet with a high specific surface area by a plasma-engraving strategy.44 The nanosheet showed 10 times higher OER activity than conventional Co3O4 materials. Cao et al. reported that a hierarchically porous Co(OH)F material was very useful for the efficient diffusion of reactants and catalyzed OER with a very small overpotential of 313 mV.45 The OER activity should still be improved by designing the number of active sites at the surfaces of electrocatalysts as well as the enhancement of diffusion in the nanopores of electrocatalysts. Accordingly, the precise design of nanoporous structures is one of the recent research trends in optimizing OER electrocatalysts.
Tafel slope (b) is used to evaluate the reaction kinetics as well as reaction mechanisms. The Butler–Volmer equation is defined by eqn (5) as the most fundamental relationship in electrochemical kinetics.
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
![]() | (9) |
Here, λ is reorganization energy. If the Tafel slope of the electrocatalyst is 120 mV dec−1, RDS follows a single-electron reaction. According to Bockris and Reddy, the transfer coefficients in multi-electron reactions are formulated as,
![]() | (10) |
Innovative strategies have been developed in the research fields of metal–organic framework (MOF) and covalent-organic framework (COF) type porous materials with designable organic linkers.52–54 The synthetic strategies to tune the pore shape and surface function have been triggered with an elegant proposal as the reticular chemistry by Yaghi et al. since 1995.55 Further functionalizations of organic linkers have been challenged by the precise adjustment of physical properties and careful design of nanoporous structures, attracting much attention as candidates for OER electrocatalysts. The utilization of MOF and COF would be beneficial for structural determination because of their highly crystalline, uniform, and designable frameworks. To maximize the OER activity of such porous materials showing the functions arising from organic linkers, we should understand the physical and chemical properties of the clear frameworks during the water-splitting reaction.
| OER electrocatalyst | Substrate | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | BET surface area (m2 g−1) | Pore size (nm) | Pore volume (cm3 g−1) | Ref. |
|---|---|---|---|---|---|---|---|---|
| Co3O4 | Carbon fiber paper | 1 M KOH | 409 | 62 | 37.7 | — | — | 62 |
| Co3O4 | Carbon black | 1 M KOH | 368 | 59 | 47.3 | 26.2 | — | 63 |
| Co3O4 | Ni foam | 1 M KOH | 311(η20) | 76 | — | — | — | 64 |
| Co3O4 NPs | Glassy carbon | 1 M KOH | 392 | 44 | 23.2 | 3.0 | 0.0082 | 65 |
| Co3O4 NWs | Glassy carbon | 1 M KOH | 392 | 53 | 22.1 | 15.0 | 0.077 | 65 |
| Co3O4 NSs | Glassy carbon | 1 M KOH | 394 | 52 | 24.2 | 4.5 | 0.019 | 65 |
| Co3O4 NCs | Glassy carbon | 1 M KOH | 405 | 58 | 15.7 | 3.9 | 0.0040 | 65 |
| DL-Co3O4 | Ni foam | 1 M KOH | 256(η50) | 61 | 57.4 | 12.6 | — | 66 |
| Cl–Co3O4-h | Ni foam | 3 M KOH | 257 | 70 | 202 | 3.3 | 0.48 | 67 |
| CoOx | Glassy carbon | 1 M KOH | 306 | 65 | 30.4 | — | — | 68 |
| CoOx | Carbon fiber paper | 1 M KOH | 300 | 40 | — | — | — | 69 |
| NiO | Ni foam | 1 M KOH | 310 | 54 | — | — | — | 70 |
| NiO/NiFe2O4 | ITO | 0.1 M KOH | 328 | 42 | 107 | 3.0 | — | 71 |
| NixMn1−xOy | Glassy carbon | 1 M KOH | 297 | 91 | 152 | 7.9 | — | 72 |
| NixCo3−xO4 | Ni Foam | 1 M KOH | 287 | 88 | 118 | 2.0–4.0 | — | 73 |
| NixCo3−xO4 | Ti foil | 1 M NaOH | — | 64 | 112 | — | — | 74 |
| Ni–FeOx | Glassy carbon | 1 M KOH | 284 | 48 | — | — | — | 75 |
| NiCo2O4 | Ni Foam | 1 M KOH | 271 | 172 | 60.7 | ≈3.0 | 0.33 | 76 |
| MnCo2O4 | Ni Foam | 1 M KOH | 289 | 182 | 26.4 | ≈3.0 | 0.11 | 76 |
| ZnCo2O4 | Ni Foam | 1 M KOH | 340 | 183 | 34.7 | ≈3.0 | 0.17 | 76 |
| (Fe0.2Co0.2Ni0.2Cr0.2Mn0.2)3O4 | Carbon paper | 1 M KOH | 275 | 50 | 118 | 1.7–5.7 | — | 77 |
| Fe–Co3O4 | Ni foam | 1 M KOH | 204 | 38 | 199 | 0.95 | — | 78 |
| Fe–Co3O4 | Glassy carbon | 0.1 M KOH | 486 | — | 110 | 3.0, 11.0 | 0.16 | 79 |
| Co2–Ni1–O | Ni foam | 1 M KOH | 310 | 57 | 181 | 2.5 | — | 80 |
| NiCo2O4 | FTO | 1 M KOH | 565 | 292 | 71.9 | — | — | 81 |
| NiCo2O4 | Carbon fiber paper | 0.1 M KOH | 340 | 63 | — | — | — | 82 |
| NiCo2O4 | Glassy carbon | 1 M KOH | 230 | 85 | — | — | — | 83 |
| CoFe2O4 | Glassy carbon | 1 M KOH | 342 | 57 | 163.3 | — | 0.79 | 84 |
| CoFe2O4 | Glassy carbon | 0.1 M KOH | 408 | 82 | 61.5 | 4.0 | — | 85 |
| ZnxCo3−xO4 | Ti foil | 1 M KOH | 320 | 51 | 78.5 | 3.0, 6.0 | — | 86 |
| ZnCo2O4 | Glassy carbon | 1 M KOH | 389 | 60 | 65.9 | 6.3 | — | 87 |
| WCoO-NP | Glassy carbon | 1 M KOH | 270 | 92 | 142 | 8.3 | 0.24 | 88 |
| NFC@CNSs | Glassy carbon | 1 M KOH | 256 | 61 | 145 | 11.7 | 0.50 | 89 |
A monometallic cobalt oxide (Co3O4), one of the 3d transition metal oxides, has been widely investigated as an OER electrocatalyst. Transition metals, such as cobalt, nickel and iron, exhibit redox activity and are suitable as catalysts for electrochemical reactions. Xie et al. succeeded in fabricating a porous Co3O4 nanosheet at the atomic scale by fast-heating strategy using an intermediate precursor of an atomically-thick CoO nanosheet (see Fig. 4a).90 The ultrathin thickness of 0.43 nm afforded 5-coordinated Co3+ and 4-/3-coordinated pore-surrounding Co3+ atoms. The high pore occupancy facilitated easy electrolyte infiltration and ensured a large contact area with the electrolyte, thus enlarging the reaction space. This Co3O4 porous nanosheet yielded a highly dense current of up to 341.7 mA cm−2 at 1.0 V vs. Ag/AgCl. Paik et al. synthesized a porous Co3O4 nanosheet by a graphene-templated method (see Fig. 4b). This Co3O4 nanosheet showed an excellent OER overpotential of 368 mV at 10 mA cm−2, comparable to the benchmark catalyst (RuO2).63 Yin et al. converted Co(OH)2 nanoplates to CoOx with the formation of controllable oxygen vacancies (see Fig. 4c).68 The Co(OH)2 nanoplate was prepared by the hydrolysis of CoCl2 and modified with polyacrylic acid as a reducing agent through the coordination of the carboxyl group to Co2+. The resultant CoOx nanoplate having surface oxygen vacancies showed an extremely small overpotential of 306 mV at 10 mA cm−2. Lu et al. enhanced OER activity through the precise design of a highly porous Co3O4 replicated by using a KIT-6 type mesoporous silica with a doping of metal species (e.g., Pd).91 Electrons in Co3+ migrated due to the presence of Pd nanoparticles (NPs) and changed into Co4+ as the higher oxidation state, accelerating the formation of *OOH species. Accordingly, the Pd-doped mesoporous Co3O4 type material showed an onset OER overpotential of 415 mV smaller than that without Pd (480 mV).
![]() | ||
| Fig. 4 Base metal oxide-based OER electrocatalysts: (a) characteristics of a porous Co3O4 nanosheet fabricated using a fast-heating strategy,90 (b) schematic of the preparation of a porous Co3O4 nanosheet and linear sweep voltammetry (LSV) curves and Tafel slopes,63 (c) schematic of the preparation of a porous CoOx nanoplate with LSV curves,68 (d) schematic of the preparation of a 3D porous NiO with LSV curves and Tafel slopes,70 (e) representative structures of a normal spinel (MgAl2O4), an inverse spinel (NiFe2O4) and a complex spinel (CuAl2O4) in different styles and views,92 (f) schematic of the preparation of CoFe2O4 through replication using a SBA-15 type mesoporous silica,84 and (g) schematic of the preparation of a porous CoFe2O4 nanosheet with LSV curves and Tafel slopes.93 Reproduced with permissions from ref. 90 and 92, Copyright 2014 and 2017 Royal Society of Chemistry and ref. 63, 68, 70, 84 and 93, Copyright 2016, 2018, 2018, 2019 and 2024 Elsevier. | ||
Pawar et al. designed a 3D porous NiO layer, which showed enough mechanical adhesion at the nickel (Ni) foam substrate surface by annealing at high temperatures (see Fig. 4d).70 The thickness of the resultant nanoporous wall can be controlled by varying the annealing temperature. The parent porous structure of the Ni foam is important for the formation of a high-surface-area nanoporous structure. For example, the electrocatalyst constructed by the thinnest NiO porous layer on the Ni form showed an OER overpotential of 310 mV at 10 mA cm−2.
The crystal structures of mixed metal oxides (spinel structure of AB2O4; A = Li, Mn, Zn, Cd, Co, Cu, Ni, Mg, Fe, Ca, Ge, Ba, etc. B = Al, Cr, Mn, Fe, Co, Ni, Ga, In, Mo, etc.) are determined as outlined by Bragg and Nishikawa since 1915.94,95 Oxygen anions are arranged in a cubic close-packed lattice and metal ions fill the gaps between the tetrahedral and octahedral units (see Fig. 4e). The spinel is classified into three types, such as like normal, inverse and complex of crystal structures, depending on the cation distribution.92 Such spinel-type mixed metal oxides are promising as high-performance OER electrocatalysts based on good conductivity, structural stability and high catalytic activity.96 Fortunelli et al. demonstrated the importance of iron (Fe) species through theoretical calculation and experimental analysis, revealing that the lattice-oxygen oxidation mechanism proceeded in preference to the adsorbate evolution for OER using CoFe2O4.97 Cuenya et al. also compared the OER activity based on epitaxially grown thin films of Co3O4(111), CoFe2O4(111) and Fe3O4(111).98 CoFe2O4(111) showed up to about four and nine times activity than Co3O4(111) and Fe3O4(111), respectively. Hao et al. succeeded in fabricating an ordered mesoporous CoFe2O4 type material through hard-templating using an SBA-15 type ordered mesoporous silica (Fig. 4f).84 This unique mesoporous CoFe2O4 type material resulted in not only an excellent OER overpotential of 342 mV at 10 mA cm−2 but also a high tolerance during OER even under alkaline conditions. Zhao et al. also prepared a porous CoFe2O4 nanosheet over Fe foam through a spontaneous redox reaction between the Fe foam and Co2+ (see Fig. 4g).93 The resultant porous CoFe2O4 nanosheet showed a small OER overpotential of 429 mV at 10 mA cm−2 under neutral and alkaline conditions.
![]() | ||
| Fig. 5 Base metal hydroxide-based OER electrocatalysts: (a) schematic of layered double hydroxide,105 (b) cyclic voltammetry at 50 mV s−1 in 1.0 M KOH solution of bulk-Co–OH and meso-Co–OH,106 (c) schematic of the phase transition of Co(OH)2 in the OER with LSV curves at a scan rate of 2 mV s−1 after the correction (iR compensation: 100%),107 (d) volcano plot of the intrinsic activities of transition metal (oxy)hydroxides vs. M–OH bond strength where the green dotted lines indicate a hypothetical, perfect volcano,108 and (e) schematic of the NiFe-LDH nanoplate with LSV curves.109 Reproduced with permissions from ref. 105 and 106, Copyright 2019 and 2016 Elsevier, ref. 107 and 108, Copyright 2022 and 2016 American Chemical Society and ref. 109, Copyright 2014 Royal Society of Chemistry. | ||
| OER electrocatalyst | Substrate | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | BET surface area (m2 g−1) | Pore size (nm) | Pore volume (cm3 g−1) | Ref. |
|---|---|---|---|---|---|---|---|---|
| α-Co(OH)2 | Glassy carbon | 1 M KOH | 320 (η20) | — | 457 | 4.0 | 0.43 | 106 |
| β-Co(OH)2 | Co plate | 1 M KOH | 332 | 68 | — | — | — | 110 |
| β-Ni(OH)2 | Ni foam | 1 M KOH | 250 | 61 | — | — | — | 111 |
| Co(OH)2@NCNTs | Ni foam | 1 M KOH | 270 | 72 | — | — | — | 112 |
| Co-LDH | Glassy carbon | 1 M KOH | 218 | 88 | 103.82 | 13.9 | 113 | |
| CoFe-LDH | Glassy carbon | 1 M KOH | 410 | 111 | 127 | <150 | 0.44 | 114 |
| CoFe-LDH/BUGC | Biordered ultramicroporous graphitic carbon | 1 M KOH | 370 | 199 | 177 | <16 | 0.33 | 114 |
| CoFe-LDH/IMC | Immense microporous carbon | 1 M KOH | 305 | 85 | 251 | <16 | 0.71 | 114 |
| CoFe-LDH/MMC | Micro-/mesoporous carbon | 1 M KOH | 285 | 69 | 234 | <20 | 0.83 | 114 |
| NiFe LDH | Ni foam | 0.1 M KOH | 280 (η30) | 50 | 0.24 (m2 cm−2) | — | — | 109 |
| Ni0.33Fe0.66 DH | Glassy carbon | 1 M KOH | 320 | 59 | 429 | 2.0–10.0 | 0.54 | 115 |
| Ni0.5Fe0.5 LDH | Glassy carbon | 1 M KOH | 284 | 48 | 414 | 2.0–10.0 | 0.72 | 115 |
| Ni0.66Fe0.33 LDH | Glassy carbon | 1 M KOH | 248 | 46 | 333 | 2.0–4.0, 5.0–50 | 0.99 | 115 |
| Ni0.8Fe0.2 LDH | Glassy carbon | 1 M KOH | 283 | 71 | 185 | 2.0–10.0 | 0.20 | 115 |
| FeNi–OH LDH | Ni foam | 1 M KOH | 244 (η50) | 47 | — | — | — | 116 |
| NiCo LDH | Ni foam | 0.1 M KOH | 420 | 113 | — | — | — | 117 |
| NiCo-LDH | Glassy carbon | 1 M KOH | 203 | 81 | 146.3 | 7.8 | — | 113 |
| NiAl LDH | Ni foam | 1 M KOH | 180 | 93 | — | — | — | 118 |
| NiFe LDH@NiCoP | Ni foam | 1 M KOH | 220 | 49 | — | — | — | 119 |
| NiTe@FeNi LDH | Ni foam | 1 M KOH | 218 | 32 | — | — | — | 120 |
| NiFeV LDH | Ni foam | 1 M KOH | 195 (η20) | 42 | — | — | — | 121 |
| NiFeCo LDH | Ni foam | 1 M KOH | 210 | 39 | — | — | — | 122 |
| Ni3S2–NiFe LDH | Ni foam | 1 M KOH | 230 (η50) | 61 | — | — | — | 123 |
| NiFeCo-LDH@MXene | Ni foam | 1 M KOH | 220 | 52 | — | — | — | 124 |
| FeOOH/Ni/NF | Ni foam | 1 M KOH | 190 | 26 | 30.4 | 200 | — | 125 |
| Co(OH)2/La(OH)3 | Carbon fiber paper | 1 M KOH | 273 (η100) | 89 | — | 2.0 | — | 126 |
| CuCoNi–OH | Carbon cloth | 3 M KOH | 290 | 58 | 33.1 | >10 | — | 127 |
| P–Ag–Co(OH)2 | Glassy carbon | 1 M KOH | 235 | 50 | — | — | — | 128 |
| Ni0.65Mn0.35 LDH | Ni foam | 1 M KOH | 253 (η50) | 130 | — | — | — | 129 |
| Mo-NiO@NiFe LDH | Ni foam | 1 M KOH | 253 (η1000) | 30 | — | — | — | 130 |
| CoFe-LDH/MoS2 @NDCDs | Ni foam | 1 M KOH | 258 | 93 | 12.4 | 39 | 0.13 | 131 |
Abiti et al. synthesized a mesoporous cobalt hydroxide (meso-Co–OH) with a high specific surface area of 457 m2 g−1 by a direct liquid crystal template method.106 X-ray diffraction (XRD) measurement revealed that meso-Co–OH powder had the α-Co(OH)2 structure and higher crystallinity than bulk Co–OH. The open mesoporous framework of meso-Co–OH significantly accelerated the mass property. A small overpotential of 320 mV at 25 mA cm−2 was observed by using meso-Co–OH, being 10 times lower than the current value of bulk Co(OH)2 (see Fig. 5b). Lin et al. also reported the synthesis of α-Co(OH)2 through the intercalation of NO3− into Co(OH)2 and mixed β/γ-CoOOH through dehydration and dehydrogenation, and subsequent conversion to OER-active γ-CoOOHx and β-CoOOHy (see Fig. 5c).107In situ Raman spectra of α-Co(OH)2, recorded as a function of the applied voltage, revealed that the metal species in the layered cobalt hydroxides were changed into those having a higher oxidation state during the OER reaction.
The chemical compositions of LDHs can be designed, and multiple metals have been evaluated to improve the OER performance. The combinations of trivalent (e.g., Fe3+) and divalent cations (e.g., Co2+ and Ni2+) significantly enhanced the OER activity by synergistic catalytic effects. Because Ni and Fe are on opposite sides of the volcano plot (see Fig. 5d),108 the strength of chemical bonds between the active metal sites and reaction intermediates can be varied by mixing the metals. Accordingly, various types of NiFe-based OER catalysts have been developed so far. Duan et al. successfully deposited vertically aligned nanoplates of 3D porous NiFe-LDH over Ni foam by hydrothermal synthesis and accelerated the diffusion of electrolytes with a small OER overpotential of 280 mV at 30 mA cm−2 (Fig. 5e).109 Additional doping of metals to NiFe-LDHs was also significant for enhancing the electronic conductivity related to OER activity.132,133 Kim et al. reported a one-step electrochemical deposition of amorphous and 2D porous NiFeCo-LDH on Ni foam.122 The elemental mapping images taken during observations by using transmission electron microscopy (TEM) revealed that NiFeCo LDH was almost fabricated with a stoichiometric atomic ratio of Ni
:
Fe
:
Co = 1
:
1
:
0.10 and the metals were distributed uniformly on the NiFeCo LDH nanosheet. The porosity of NiFeCo LDH also accelerated the electron transport and electrolyte diffusion. The amorphous structure of NiFeCo LDH is also interesting for intercalating OH− and forming many edge sites and/or defects. The introduction of Co would cause a synergistic effect to show an excellent OER overpotential of 210 mV at 10 mA cm−2.
| OER electrocatalyst | Substrate | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | BET surface area (m2 g−1) | Pore size (nm) | Pore volume (cc g−1) | Ref. |
|---|---|---|---|---|---|---|---|---|
| CoPi film | Pt | 0.1 M KOH | 447 | — | — | — | — | 142 |
| CoPi | Glassy carbon | 1 M KOH | 380 | 59 | 210 | 2.9 | 0.18 | 143 |
| CoPi/Ti | Ti mesh | 0.1 M PBS | 450 | 187 | — | — | — | 144 |
| (Co0.5Ni0.5)3(PO4)2 | Ni foam | 1 M KOH | 273 | 59 | — | — | — | 145 |
| CoPi-HSNPC | Glassy carbon | 1 M KOH | 320 | 85 | 69 | — | — | 146 |
| NixCo3−x(PO4)2 | Ni foam | 1 M KOH | 336 | 35 | 206 | 3.6 | — | 147 |
| FeP–FePxOy | Glassy carbon | 1 M KOH | 280 | 48 | — | — | — | 148 |
| NFPy | Glassy carbon | 1 M KOH | 210 | 47 | 21.3 | — | 0.33 | 149 |
| NiGLy | Carbon paper | 1 M KOH | 349 | 78 | 274 | 1.3 | 0.52 | 150 |
| CoNTO-1–3 | Carbon paper | 1 M KOH | 312 | 61 | 204 | 1.0 | 0.23 | 151 |
| CoPIm | Carbon paper | 1 M KOH | 334 | 59 | 291 | 1.2 | 0.31 | 152 |
| NiPIm | Carbon paper | 1 M KOH | 363 | 105 | 204 | 1.1 | 0.23 | 152 |
| NiCoPIm | Carbon paper | 1 M KOH | 351 | 67 | 267 | 1.3 | 0.28 | 152 |
| CoFePi | Glassy carbon | 1 M KOH | 277 | 31 | 170 | 12.3 | 0.70 | 153 |
| CoFeNiPi | Glassy carbon | 0.1 M KOH | 309 | 51 | — | — | — | 153 |
| CoNiPi | Glassy carbon | 0.1 M KOH | 402 | 87 | 80 | 2.0–50 | — | 153 |
| FePi | Glassy carbon | 0.1 M KOH | 554 | 75 | — | — | — | 153 |
Yamauchi et al. prepared a crystalline CoPi-type material with a typical 2D hexagonal mesoporous structure using cetyltrimethylammonium bromide (CTAB, see Fig. 6a).143 The abundant active sites were helpful for realizing an excellent OER overpotential of 380 mV at 10 mA cm−2, being higher than that reported for the precious metal, base metal, and non-metal-based electrocatalysts and much higher than that observed for non-porous CoPi bulk (OER activity was 560 mV at 10 mA cm−2). Accordingly, they demonstrated that the design of porous materials is quite important for enhancing OER activity. Besides, Zeng et al. focused on synergistic catalytic effects of trivalent Fe3+ and divalent Co2+ and Ni2+ for improving the electrochemical activity by adjusting the electronic environment of the active site.153 Amorphous base metal phosphates, such as CoPi, FePi and binary CoFePi, were prepared by using a bi-template coprecipitation method and the resultant OER activities were compared (see Fig. 6b). By the synergistic effects, the OER overpotential CoFePi (315 mV at 10 mA cm−2) was quite superior to those observed for (388 mV and 554 mV). To demonstrate a new guideline for the design of base metal phosphates to improve OER performance, Yamauchi et al. also synthesized three porous metal phosphonates such as cobalt (CoPIm), nickel (NiPIm) and nickel–cobalt phosphonates (NiCoPIm) by utilizing iminodi(methylphosphonic acid) that can interact with Co2+ and Ni2+ with the introduction of organic functions (see Fig. 6c).152 An electron donation from nitrogen filled the M–OH antibonding orbital and caused modulation of the interactions between metal and oxygen, leading to excellent OER overpotentials of 334 mV (CoPIm), 363 mV (NiPIm) and 351 mV (NiCoPIm).
![]() | ||
| Fig. 6 Base metal phosphate-based OER electrocatalysts: (a) schematic of the preparation of CTAB templated CoPi with crystallized frameworks and the corresponding OER overpotentials,143 (b) schematic formation mechanism of bi-templated CoPi and other transition metal containing ions and the resultant LSV curves at a scan rate of 10 mV s−1,153 and (c) characteristics of organic–inorganic transition metal-based phosphonates synthesized through a simple hydrothermal reaction, and their LSV curves.152 Reproduced with permissions from ref. 143 and 153, Copyright 2016 and 2018 Wiley-VCH and ref. 152, Copyright 2020 Elsevier. | ||
The morphological design of porous metal phosphates has also been achieved by various methods. The strong proton donation property from the phosphate group is helpful for overcoming the low conductivity that has been a problem in the cases of base metal oxides. In addition, considering the synthesis pathway of metal phosphonates, a wide variety of metal phosphonates, where metal phosphate units and organic groups are located at the molecular scale in the inorganic–organic hybrid frameworks, are prepared with the selection of metal species and the design of organic functions.154,155 This unique structural feature is different from those of inorganic-only materials. The introduction of organic groups gives us a good opportunity to tune the electronic structure of the active metal species as an environment optimized for OER.153
In base metal-based inorganic materials, forming porous nanostructures is the most efficient and universal strategy to enhance OER activity. As summarized in Fig. 7, a certain correlation between the specific surface area and OER performance was confirmed for oxide-based materials, though such a tendency has hardly been found for hydroxide- and phosphate-based ones. However, it has been reported that porous materials of hydroxides and phosphates show higher OER performance than non-porous ones.106,143 Accordingly, making inorganic-based electrocatalysts more porous and promoting mass transport have been effective ways to improve their OER performance. The precise control of porous structures (e.g., templating to control the uniformity of pore) is also very important for understanding the effects of pore size on inorganic-based porous OER electrocatalysts.
![]() | ||
| Fig. 7 Plots of the overpotential and BET-surface area in (a) oxide-based materials, (b) hydroxide-based materials, and (c) phosphate-based materials. | ||
In recent years, further optimization of OER activity has been investigated not only by improving the physical properties but also by adjusting the electronic environment of metal species through the combination of multiple metal species and the doping of heteroatoms such as nitrogen and sulfur. Synergistic catalytic effects between trivalent Fe ions and divalent Co or Ni ions also play an important role in enhancing the OER activity. Metal composite materials containing multiple metal species are up-and-coming candidates for designing highly active electrocatalysts (see Fig. 8). The lack of precise elucidation of the active site structure and OER mechanism is a common and major problem in the case of base metal-based electrocatalysts. Recent developments in operando and in situ spectroscopic technologies have revealed that the OER mechanism varies according to the catalyst and changes in the reaction environment.156 The diversity of the OER mechanism is an obstacle to designing electrode materials for scale-up to industrial applications. More detailed and accurate elucidation of the OER mechanism is urgently needed to systematically understand the influences of the electronic states in the structures and metal species on the OER performance.
| OER electrocatalyst | Substrate | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | BET surface area (m2 g−1) | Pore size (nm) | Pore volume (cc g−1) | Ref. |
|---|---|---|---|---|---|---|---|---|
| NiCoFe-MOF-400 °C-1 h | Glassy carbon | 1 M KOH | 238 | 29 | 558 | 1 | 0.19 | 157 |
| NiCoFe-MOF-74-600 °C-2 h | Glassy carbon | 1 M KOH | 366 | 117 | 197 | — | 0.060 | 157 |
| (Co,0.3Ni)-HMT | Glassy carbon | 1 M KOH | 330 | 66 | 100.2 | 18.4 | 0.54 | 158 |
| Ni(Fe)-MOF-74 | Ketjenblack | 1 M KOH | 274 | 40.4 | 684 | 1.0, 5.0 | 0.34 | 159 |
| NiFe-MOF/G | Glassy carbon | 1 M KOH | 258 | 49 | 437 | — | — | 160 |
| FeNi-DOBDC-3 | Carbon | 1 M KOH | 270 (η50) | 62.7 | 77.9 | 11.0 | 0.27 | 161 |
| MFN@GA/NF | Ni foam | 1 M KOH | 270 (η20) | 66 | 77.8 | 12.4 | — | 162 |
| FCN-BTC-MOF | Ni foam | 1 M KOH | 218 | 29.3 | 233.9 | 3.3 | — | 163 |
| NiCo-MOF | Glassy carbon | 1 M KOH | 270 | — | 27.5 | 1.5–60 | — | 164 |
| CoxZn3−x(HITP)2 | Glassy carbon | 1 M KOH | 210 | 50 | 520.5 | 0.71 | 0.60 | 165 |
| CoxMn3−x(HITP)2 | Glassy carbon | 1 M KOH | 250 | 54 | 371.5 | 0.53 | 0.45 | 165 |
| CoxNi3−x(HITP)2 | Glassy carbon | 1 M KOH | 280 | 56 | 403.0 | 0.74 | 0.50 | 165 |
| NiPc-MOF | Quartz | 1 M KOH | 250 (ηonset) | 74 | 593 | — | — | 166 |
| Fe1Ni4–HHTP | Carbon cloth | 1 M KOH | 213 | 96 | — | 1.8 | — | 167 |
| dye@MOF | Ni foam | 1 M KOH | 194 | 199 | — | — | — | 168 |
| Co3(HITP)2 | Carbon cloth | 1 M KOH | 254 | 86.5 | 281 | — | 0.62 | 169 |
| Am-FeCo(OH)x-30 | Glassy carbon | 1 M KOH | 257 | 47 | 838 | 2.0 | — | 170 |
| RuNC/Ni-M-SH | Glassy carbon | 1 M KOH | 248 | 57.3 | 420 | 1.41 | — | 171 |
| MOF-H | Ni foam | 1 M KOH | 263 | 63.4 | — | — | — | 172 |
| MOF-F | Ni foam | 1 M KOH | 262 | 52.7 | — | — | — | 172 |
| MOF-Cl | Ni foam | 1 M KOH | 258 | 57.5 | — | — | — | 172 |
| MOF-Br | Ni foam | 1 M KOH | 251 | 44.5 | — | — | — | 172 |
Su et al. fabricated MOFs with different conductivities due to a modulation of the electronic active site by substituting metal species and linkers in the conductive MOF.173 The conductivity was correlated with the OER activity because an enhancement in the conductivity promotes the transfer and the supply of electrons during the reactions. Du et al. synthesized a nickel phthalocyanine-based 2D MOF (NiPc-MOF) on fluorine-doped tin oxide (FTO) substrate (see Fig. 9a).166 NiPc-MOF promoted an electron transfer within the structure via in-plane π-delocalization and weak out-of-plane π–π stacking of phthalocyanine, resulting in a small OER overpotential of 250 mV at 1.0 mA cm−2. The triphenylene derivative organic linker is very interesting for adjusting the electronic conductivity in the range of 10−6 to 101 S cm−1 by designing the organic linker and the valuation of the metal centers.174–178 Wang et al. succeeded in synthesizing a nanowire array of triphenylene derivative 2D MOF (Fe1Ni4–HHTP, HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene) on a carbon cloth (CC) substrate (see Fig. 9b).167 Calculations based on the density functional theory (DFT) revealed that self-adaptive structural adjustment of the Fe site was generated through the hydrogen bond between the OER intermediate and the adjacent layer with the reduction of the free energy in the OER process. The resultant Fe1Ni4–HHTP showed small OER overpotentials of 213 mV and 300 mV at 10 mA cm−2 and 150 mA cm−2, respectively. Zhang et al. synthesized two isostructural MOFs with different proton conductivities (FJU-82-Co; 7.40 × 10−5 S cm−1, FJU-82-Zn; 5.80 × 10−7 S cm−1 at 60 °C under 98% RH) to elucidate a direct correlation between proton conductivity and OER activity (see Fig. 9c).179 FJU-82-Co showed a much lower overpotential of 570 mV than that observed for FJU-82-Zn (1170 mV) at 1.0 mA cm−2. Jiao et al. encapsulated 3-hydroxy-2,5,6,8,9-pentanitro-pyrene-1-sulfonate (HPTS) as a dye molecule into the pores of a MOF (dye@MOF) (see Fig. 9d).168 The hydroxyl and sulfonic groups of the HPTS molecule improved the proton conductivity of dye@MOF. The resultant dye@MOF showed an extremely small OER overpotential of 194 mV at 10 mA cm−2 and higher OER activity.
![]() | ||
| Fig. 9 Conductive site introduced MOF-based OER electrocatalysts: (a) schematic of the preparation of NiPc-MOF and LSV curves in various electrolytes,166 (b) space-filling drawings of Ni-HHTP along the c axis with resultant LSV curves in 1 M KOH,167 (c) perspective view of the four-layer stacking diagram along the b-axis of FJU-82-Co and FJU-82-Zn,179 and (d) LSV curves of dye@MOF.168 Reproduced with permissions from ref. 166 and 167, Copyright 2018 and 2019 Royal Society of Chemistry, ref. 179, Copyright 2020 American Chemical Society and ref. 168, Elsevier, Copyright 2023. | ||
Jia et al. succeeded in fabricating a defect-rich hierarchical Am–FeCo(OH)x hybrid architecture by using amino-functionalized iron-based MOF (MIL-101) as a sacrificing modification agent (see Fig. 10a).170 The resultant hierarchical 3D architecture with abundant active sites/defects and coordinated unsaturated sites reduced the reaction energy barrier and accelerated the reaction kinetics, leading to a small OER overpotential of 257 mV at 10 mA cm−2. The enhanced surface hydrophilicity also promotes the release of oxygen (O2) bubbles and proton-coupled electron transfer. Li et al. synthesized MIL-101(Fe)-X with the different amounts of the amino (–NH2) group (X = H
:
NH2 = 1
:
0, 1
:
0.5, 1
:
1, 1
:
2, 1
:
4, 0
:
1, specific surface area; 380 m2 g−1 at the highest) by the solvothermal synthesis (see Fig. 10b).180 The generated electron–hole pairs were effectively separated by the incorporation of the electron-donating –NH2 group, and the oxygen evolution rate was maximized at the ratio of H
:
NH2 = 1
:
2 under simulated sunlight irradiation of 11.7 mmol h−1 g−1. An excessive introduction of the –NH2 group resulted in the reduction of the OER activity with a decrease in the porosity. Tripathi et al. synthesized a thiol (–SH) functionalized MOF (Ni–M–SH) by the solvothermal reaction of Ni nodes and thiolated 2-aminoterephthalic acid (TPA-SH), followed by the adsorption of ruthenium (Ru) ions to obtain RuNC/Ni–M–SH (see Fig. 10c).171 Because the –SH group was helpful for electron transfer and oxidation of Ni offered electrons, Ru3+ was stabilized in the RuNC/Ni–M–SH based electrocatalyst showing a small overpotential of 248 mV at 10 mA cm−2.
![]() | ||
Fig. 10 Function modified MOF based OER electrocatalysts: (a) schematic of the preparation of Am–FeCo(OH)x and the comparison of overpotentials at 10 mA cm−2 with different reaction times,170 (b) photocatalytic oxygen evolution mechanism of MIL-101(Fe)-X with the comparison of O2 evolution rates with the increase in amino group content,180 and (c) schematic of the preparation of RuNC/Ni–M–SH and LSV curves at a scan rate of 0.001 mV s−1 in O2-saturated 1 M KOH.171 Reproduced with permissions from ref. 170 and 180, Copyright 2021 and 2024 Elsevier and ref. 171, Copyright 2024 American Chemical Society. | ||
The OER performance is not explained systematically by using the porosity of MOF type materials only. However, by preparing a series of porous electrodes with different surface areas from a MOF-type material, a higher specific surface area is useful for improving the OER activity.157 Thus, increasing the surface area as much as possible is an important factor for maximizing the OER performance. In terms of the pore size, although the resultant data are confusing while using the difference in the composition of MOF-type materials, small pores having a diameter of around 1–3 nm seem to be suitable for OER because reactants/products are very small (<0.5 nm) such as water, hydroxide ions and oxygen atoms.
Organic–inorganic hybrid MOF-type materials have attracted much attention in the field of electrocatalysis with the outstanding development of synthetic strategies (see Fig. 11). Conductive MOFs, including conductive organic components, have been developed and several studies have revealed that the electron/proton conductivity of the frameworks is helpful for improving the OER performance. The functional design of organic linkers is an important strategy to improve the OER activity. To maximize the OER activity of MOF-based electrocatalysts, a comprehensive study should be conducted for understanding the effects of porous structure, active metal center and organic linker. The development of a scalable and cost-effective synthesis method is also essential for their industrial use.
| OER electrocatalyst | Substrate | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | BET surface area (m2 g−1) | Pore size (nm) | Pore volume (cm3 g−1) | Ref. |
|---|---|---|---|---|---|---|---|---|
| Co-TpBpy | Glassy carbon | 0.1 M phosphate buffer (pH 7) | 400 (η1) | 59 | 450 | — | — | 182 |
| Co-PDY | Cu foam | 1 M KOH | 270 | 99 | — | 2.3 | — | 183 |
| Co@COF-Pyr | Glassy carbon | 1 M KOH | 450 | 100 | 392 | 1.6 | 0.37 | 184 |
| COF-TpDb-TZ-Co | Glassy carbon | 1 M KOH | 390 | 82 | 241 | 1.4 | — | 185 |
| Ni-COF | Carbon cloth | 1 M KOH | 302 | 56 | — | — | — | 186 |
![]() | ||
| Fig. 12 Active site designed COF-based OER electrocatalysts: Schematic routes to synthesize (a) Co-TpBpy via proton tautomerized Schiff base condensation and Co2+ impregnation with cyclic voltammogram and Tafel plot in a phosphate buffer at pH = 7.0182 and (b) Co-PDY with LSV curves.183 Reproduced with permissions from ref. 182 and 183, Copyright 2016 and 2019, with permission from the Royal Society of Chemistry. | ||
Even in COF, the specific surface area is not the same as the number of active sites related to the OER performance. The design of coordination sites and the pore volume is important for increasing active sites because the active metal species are coordinated in the frameworks of COF-based electrocatalysts after the formation of periodic porous structures. In addition, the pore size of COF-based electrocatalysts seems to be slightly larger than that of MOF-based ones. This is due to the fact that reactants and products should have access to active metal centers.
In the application of COF for OER, the recent trend is the introduction of active metal species with the design of organic monomers (see Fig. 13), providing a flexible control of active metal species. In combination with the clear structure of COF-based porous materials, it is expected to make a significant contribution to the elucidation of the OER mechanism at the active site. The limited number of monomer candidates and the small variation in porous structure are major problems. The field of applications of COF as electrochemical catalysis is just beginning to attract attention, and further development is desirable.
mA cm−2 by a small cell voltage of 1.42 V, being lower than that observed for the benchmark catalysts (1.57
V for IrO2–Pt).
| Electrocatalyst | Substrate | Electrolyte | HER | OER | Cell voltage (V at 10 mA cm−2) | Long-term stability | Ref. | ||
|---|---|---|---|---|---|---|---|---|---|
| η (mV) | Tafel slope (mV dec−1) | η (mV) | Tafel slope (mV dec−1) | ||||||
| PCPTF | Glass slide | 1 M KOH | 430 (η30) | 53 | 330 (η30) | 65 | — | 13.8 h (at 10 mA cm−2) | 191 |
| NSP-Co3FeNx | Ni foam | 1 M KOH | 23 (η10) | 94 | 222 (η20) | 46 | 1.54 | 2000 CV cycle | 192 |
| Cu-CMP | Glassy carbon | 1 M KOH | 350 (η10) | 135 | 450 (η10) | 62 | — | 12 h (at 1 mA cm−2) | 193 |
| Ni/NiP | — | 1 M KOH | 130 (η10) | 59 | 270 (η30) | 73 | 1.61 | 4 h (at 10 mA cm−2) | 194 |
| MoO2 | Ni foam | 1 M KOH | 27 (η10) | 41 | 260 (η10) | 54 | 1.53 | 24 h (at 10 mA cm−2) | 195 |
| Ni2P/Ni | Ni foam | 1 M KOH | 98 (η10) | 72 | 200 (η10) | — | 1.49 | 20 h (at 20 mA cm−2) | 196 |
| NiFe/NiCo2O4 | Ni foam | 1 M KOH | 105 (η10) | 88 | 340 (η1200) | 39 | 1.67 | 20 h (at 10 mA cm−2) | 197 |
| FeP/Ni2P | Ni foam | 1 M KOH | 14 (η10) | 24 | 154 (η10) | 23 | 1.42 | 40 h (at 500 mA cm−2) | 190 |
| Fe-H2cat | Fe foam | 1 M KOH | 243 (η10) | 77 | 238 (η10) | 83 | 1.65 | 50 h (at 10 mA cm−2) | 198 |
| Ni0.75Fe0.125V0.125-LDHs/NF | Ni foam | 1 M KOH | 125 (η10) | 64 | 231 (η10) | 39 | 1.59 | 15 h (at 30 mA cm−2) | 199 |
| CoNi/CoFe2O4 | Ni foam | 1 M KOH | 82 (η10) | 96 | 230 (η10) | 45 | 1.57 | 48 h (at 100 mA cm−2) | 189 |
| NiFeCo LDH | Ni foam | 1 M KOH | 108 (η10) | 73 | 210 (η10) | 39 | 1.57 | 50 h (at 10 mA cm−2) | 122 |
| CoP–N | Co foam | 1 M KOH | 100 (η50) | 82 | 260 (η50) | 51 | 1.61 | 50 h (at 50 mA cm−2) | 200 |
| Ni@NiFe LDH | Ni foam | 1 M KOH | 92 (η10) | 72 | 218 (η10) | 66 | 1.53 | 24 h (at 10 mA cm−2) | 201 |
| Co3(OH)2(HPO4)2 | Ni foam | 1 M KOH | 87 (η10) | 91 | 182 (η10) | 69 | 1.54 | 240 h (at 30 mA cm−2) | 202 |
| Ni–Co–Fe–P NBs | Ni foam | 1 M KOH | 35 (η10) | 65 | 187 (η10) | 29 | 1.46 | 100 h (at 100 mA cm−2) | 203 |
| FeNiZn/FeNi3 @NiFe | Ni foam | 1 M FeKOH | 245 (η1000) | 45 | 367 (η1000) | 46 | 1.58 (at 100 mA cm−2) | 100 h (at 150 mA cm−2) | 204 |
| Ni–Fe–Mn–P/NC | Ni foam | 1 M KOH | 72 (η10) | 80 | 274 (η30) | 57 | 1.52 | 35 h (at 50 mA cm−2) | 205 |
| NaOH-ABC | Carbon cloth | 0.5 M H2SO4 | 231 (η10) | 168 | 155 (η10) | 144 | — | — | 206 |
| KOH-I ABC | Carbon cloth | 0.5 M H2SO4 | 213 (η10) | 146 | 126 (η10) | 137 | — | — | 206 |
| KOH-II ABC | Carbon cloth | 0.5 M H2SO4 | 188 (η10) | 137 | 90 (η10) | 127 | 1.49 | — | 206 |
| NiCo2O4-B-CC | Carbon cloth | Carbon cloth | 26 (η10) | 106 | 215 (η10) | 95 | 400 (at 400 mA cm−2) | 20 h (at 10 mA cm−2) | 207 |
![]() | ||
| Fig. 14 HER-OER bifunctional electrocatalysts: Schematics of the preparation of (a) porous CoNi/CoFe2O4/NF and LSV curves,189 (b) FeNiZn/FeNi3@NiFe and comparison of the overpotentials204 and (c) N-self-doped defect-rich porous carbon nanosheets and LSV curves.206 Reproduced with permissions from ref. 189, Copyright 2018 Royal Society of Chemistry, ref. 204, Copyright 2023 Springer Nature, ref. 206 Copyright 2024 Elsevier. | ||
Liu et al. succeeded in fabricating an outstanding bifunctional electrocatalyst FeNiZn/FeNi3@NiFe (see Fig. 14b).204 The interfaces between FeNiZn and FeNi3 were active for the reactions, which were promoted due to fast mass transport by the multimodal porous structure. The excellent electrocatalytic activity was then achieved even at the high current density (1000 mA cm−2) with small overpotentials of 367 mV and 245 mV in 1 M KOH for both HER and OER, respectively. Besides, FeNiZn/FeNi3@NiFe could operate for 400 hours at 1000 mA cm−1 without significant deactivation. Kim et al. developed N-self-doped defect-rich porous carbon nanosheets derived from Platycladus orientalis tree-cone biomass as a carbon-based bifunctional electrocatalyst without metals (see Fig. 14c).206 This nanosheet-type material exhibited an exceptional specific surface area of 3369 m2 g−1, pore volume of 2.1 cm3, and electric conductivity of 12.69 S cm−1 and thus showed small overpotentials of 188 mV and 90 mV at 10 mA cm−2 in 0.5 M H2SO4 for HER and OER, respectively. In the overall water-splitting reaction using the metal-free nanosheet as both electrodes, a current density of 10 mA cm−2 was obtained at a low voltage of 1.49 V.
| OER electrocatalyst | Substrate | Electrolyte | η (mV) | Tafel slope (mV dec−1) | Long-term stability | Ref. |
|---|---|---|---|---|---|---|
| S–NCFO | Ni foam | 0.5 M NaCl + 1 M KOH | 290 (η50) | 32 | 48 h (at 100 mA cm−2) | 216 |
| NixCryO | Carbon paper | Seawater + 1 M KOH | 270 (η100) | — | 280 h (at 500 mA cm−2) | 217 |
| Fe(Cr)OOH/Fe3O4 | Ni foam | Seawater + 1 M KOH | 278 (η500) | 34 | 100 h (at 400 mA cm−2) | 218 |
| S–(Ni,Fe)OOH | Ni foam | Seawater + 1 M KOH | 300 (η100) | 49 | 100 h (at 100 mA cm−2) | 219 |
| CoFeZr | Ni foam | 0.5 M NaCl + 1 M KOH | 303 (η100) | 41 | 20 h (at 100 mA cm−2) | 220 |
| NF/NiFe LDH | Ni foam | Seawater + 1 M KOH | 247 (η100) | 33 | 96 h (at 500 mA cm−2) | 221 |
| NiFe-LDH | Ni foam | 0.5 M NaCl + 1 M KOH | 257 (η500) | — | 24 h (at 500 mA cm−2) | 222 |
| NixFeyN@C | Ni foam | Seawater + 1 M KOH | 283 (η100) | 45 | 100 h (at 500 mA cm−2) | 223 |
| Ni2P–Fe2P | Ni foam | Seawater + 1 M KOH | 305 (η100) | 58 | 38 h (at 500 mA cm−2) | 224 |
| S–NiFe–Pi | NiFe foam | Seawater + 1 M KOH | 241 (η100) | 52 | 100 h (at 500 mA cm−2) | 225 |
| ZnFe-BDC-0.75 | Ni foam | Seawater | 308 (η100) | 48 | 50 h (at 10 mA cm−2) | 226 |
| Ni3Fe-BDC | Flake graphite | 1 M NaCl + 1 M KOH | 295 (η10) | 95 | 25 h (at 10 mA cm−2) | 227 |
| NH2–NiCoFe–MIL-101 | Ni foam | Seawater + 1 M KOH | 390 (η1000) | 48 | 90 h (at 60 mA cm−2) | 228 |
| MIL-88(FeCoNi) (HMIL-88@PPy-TA) | Ni foam | 0.5 M NaCl + 1 M KOH | 244 (η100) | 46 | 100 h (at 100 mA cm−2) | 229 |
| NiFe-LDH/MOF | Ni foam | Seawater + 1 M KOH | 235 (η20) | 61 | 100 h (at 20 mA cm−2) | 230 |
| NiFe-MOF@Ni2P/Ni(OH)2 | Ni foam | 0.5 M NaCl + 1 M KOH | 302 (η100) | 43 | 120 h (at 500 mA cm−2) | 231 |
| Ni MOFs/FeS | Fe foam | Seawater + 1 M KOH | 280 (η100) | 55 | 25 h (at 100 mA cm−2) | 232 |
The design of a protective layer over an electrocatalyst is the most direct method to reduce the OER selectivity and Cl− corrosion. The access of Cl− to the active site is prevented by the presence of a protective layer. Feng et al. succeeded in fabricating an electrocatalyst (NixFeyN@C/NF) for seawater-splitting by growing an array of nickel–iron nitride micro-sheets coated with a carbon layer on Ni foam (see Fig. 15a).223 In addition to the existence of many active sites, the mass and charge transport was accelerated by the superhydrophilic and superhydrophobic surfaces and the synergistic effect of the Ni3FeN and carbon layers. NixFeyN@C/NF showed excellent OER overpotentials of 283 mV and 351 mV in alkaline simulated seawater at current densities of 100 mA cm−2 and 500 mA cm−2, respectively, being very stable during the operation for up to 100 hours at a current density of 100 mA cm−2. Surface poisoning by Cl− as well as the surface deposition of insoluble species during seawater-splitting is prevented by enhancing hydrophilicity and wettability with surface modification of metal oxide derivative electrocatalysts.233 Yang et al. succeeded in fabricating a hybrid electrocatalyst (Fe(Cr)OOH/Fe3O4/NF) with the heterointerface of Fe(Cr)OOH and Fe3O4 on Ni foam (see Fig. 15b).218 The super hydrophilic surface and high permeability of the electrolyte based on the introduced Cr (a first-row transition element with strong oxidizing properties) promoted electron transfer in the OER process, resulting in a small OER overpotential of 278 mV at a high current density of 500 mA cm−2 in simulated seawater. The electrolyzer using Fe(Cr)OOH/Fe3O4/NF as the anode achieved a current density of 10 mA cm−2 with a small cell voltage of 1.49 V and maintained a high current density of 400 mA cm−2 for 100 hours.
![]() | ||
| Fig. 15 Seawater-splitting OER electrocatalysts: Schematic of the preparation of (a) NixFeyN@C/NF with a comparison of the overpotentials for HER and OER at 100 mA cm−2 and 500 mA cm−2 in different electrolytes and durability tests of the electrolyzer at 100 mA cm−2 and 500 mA cm−2 in 1 M KOH and alkaline seawater in 1 M KOH223 and (b) Fe(Cr)OOH/Fe3O4/NF with LSV curves for OER at high current densities measured in different electrolytes.218 Reproduced with permissions from ref. 223, Copyright 2021 Royal Society of Chemistry, ref. 218, Copyright 2022 Elsevier. | ||
HER-OER bifunctional electrocatalysts are also very important in reducing the cost of green hydrogen production of seawater-splitting. The catalytic activity of porous nanostructured electrocatalysts can be adjusted in specific conditions by varying the electronic and surface environments. Several electrocatalysts have been developed so far to promote water electrolysis more efficiently on a lab scale than the typical benchmark IrO2–Pt catalyst. There are several challenges to be dealt with to advance bifunctional electrocatalysts further for seawater-splitting. Theoretical predictions and in situ spectroscopic studies help in the detailed understanding of the reaction mechanisms and real-time structural changes of electrocatalysts during electrolysis. DFT calculation is also very useful for identifying the active species and optimizing electrocatalysts at the molecular level.234 For the industrial use of seawater-splitting, in addition to designing electrocatalysts with high activity (η10 < 300 mV) and stability (operating for >1000 hours under high-density current of >500 mA cm−2), a simple and low-cost pre-treatment process of natural seawater is necessary to alleviate corrosion and poisoning as much as possible, as well as smooth and selective OER even by using natural seawater.
Exceptional qualities of small-sized transition metal oxide clusters are promising for utilizing metal cation defects and oxygen vacancies due to the presence of low-coordinated active sites, high-surface-energy amorphous structures, and so on.241 Fundamentals on the mechanism (e.g., lattice-oxygen oxidation mechanism and adsorbate evolution mechanism) and kinetics of OER have been disclosed further for the precise design of efficient electrocatalysts.242 Further optimization of OER activity has also been investigated by using the synergistic catalytic effect, combining the property arising from the presence of multiple metals in the crystal structure of spinel oxides and LDHs, and adjusting the electronic environment of the metals by doping heteroatoms such as nitrogen and sulfur. In recent years, there have been many reports of base metal derivative electrocatalysts with OER activity comparable to those observed for conventional precious metal electrocatalysts such as IrO2 and RuO2.
| This journal is © The Royal Society of Chemistry 2025 |