Next Article in Journal
CO2 Hydrogenation to Methanol on CuO-ZnO/SiO2 and CuO-ZnO/CeO2-SiO2 Catalysts Synthesized with β-Cyclodextrin Template
Previous Article in Journal
Catalytic Transfer Hydrogenation Performance of Magnesium-Doped ZrO2 Solid Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Synergistic Modulation of Transition-Metal-Based Electrocatalysts for Water Oxidation: A Mini Review

Department of Applied Biology and Chemical Technology and Research Institute for Smart Energy, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong SAR, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(9), 1230; https://doi.org/10.3390/catal13091230
Submission received: 18 July 2023 / Revised: 14 August 2023 / Accepted: 15 August 2023 / Published: 22 August 2023
(This article belongs to the Special Issue Materials and Phenomenology for Electrocatalysis and Redox Reactions)

Abstract

:
Synergistic modulation has been extensively explored to develop highly efficient transition-metal-based electrocatalysts for oxygen evolution reaction (OER) because coupling effects among intrinsic activity, conductivity, mass transfer, mass diffusion, and intermediates adsorption can further promote catalytic activity. In this review, recent progress in both experimental and theoretical research on synergistic modulation for transition-metal-based alkaline OER electrocatalysts is focused. Specifically, synergistic effects will be presented in the following aspects: (1) metal reactive sites and heterogeneous atoms; (2) heterogeneous atoms and crystallographic structure; (3) electronic structure and morphology; (4) elementary reaction steps; and (5) external fields. Finally, the remaining challenges and prospects of synergistic modulation for efficient OER are further proposed.

1. Introduction

Efficient electrochemical energy conversion techniques, such as water splitting, CO2 reduction, and Zn-air battery, play a crucial role in mitigating energy and environmental challenges by replacing fossil fuels and decarbonizing the power and transportation sectors [1,2]. Oxygen evolution reaction (OER) can couple with reduction reactions (hydrogen reaction and CO2 reduction) to constitute a redox reaction circle. However, the effectiveness of these techniques is impeded by the thermodynamic uphill nature and sluggish kinetics of OER occurring on the anode [3,4]. As a result, it is often necessary to use precious metal (Ru/Ir)-based electrocatalysts to facilitate OER, yet the scarcity and high cost hamper their wide application [5]. Therefore, the pursuit of affordable and efficient electrocatalysts remains an important research focus. Typically, high-performance electrocatalysts require the following features: (1) high intrinsic activity to minimize the overpotential required for catalyzing OER; (2) abundant active sites for rapid reaction kinetics; (3) favorable conductivity for efficient electron transfer; (4) sufficient channels for effective mass transfer and gas evolution; (5) robust structural and chemical stability to ensure long-term durability; and (6) low cost for scalable fabrication.
Transition-metal-based OER catalysts have been extensively explored as alternatives to precious metals because of their abundant reserves, cost-effectiveness, and favorable theoretical electrochemical activity. Transition metals are characterized by their d-orbital valence electronic structure [6,7,8]. The interaction between oxygen-containing species and the transition-metal surface leads to electron transfer at the interfaces, driven by the difference in their electrochemical potentials. Specifically, the O 2p orbital of oxygen-containing species hybridizes with the d orbitals of transition metals to split into two energy levels [9]. The catalytic activity of OER catalysts, such as transition-metal oxides, hydroxides, sulfides, and phosphides, is closely associated with the electronic number of metal d band because eg orbit can bind surface anions, thus affecting the combination of oxygen intermediates [1,10,11,12,13,14,15,16]. Although transition-metal-based catalysts have demonstrated appealing OER performances, they still exhibit shortcomings such as poor conductivity, less accessible active sites, and high activation energy barriers [17,18,19]. Numerous strategies have been adopted to modify their electronic structure, enhance electron transfer, improve mass diffusion, and optimize intermediates’ adsorption and desorption. However, most efforts focusing on a single individual aspect were unable to fulfill all the requirements for high-performance catalysts. To overcome this limitation, synergistic modulation, which has exhibited significant effects in the fields of CO2 reduction, N2 reduction, and H2 evolution reactions [20,21,22,23,24], for simultaneously targeting multiple aspects has been suggested to realize advanced OER catalysts based on transition metals.
In this review, we first summarize the recent research progress of several transition-metal-based OER catalysts (oxides, phosphides, sulfides, and borides). Then, synergistic modulations on heterogeneous atoms, defects, morphology, elementary reactions, and external fields for high-performance OER catalysts in alkaline environment are discussed. Finally, we point out the challenges associated with synergistic modulation and offer an outlook on the future directions in this emerging field.

2. Development of Transition-Metal-Based OER Electrocatalysts

The cost and availability of materials are crucial factors to consider in the production of electrocatalysts. By taking advantage of their easy and cost-effective accessibility, earth-abundant first-row transition metals such as Fe, Co, and Ni have gained significant attention (Figure 1) [25,26,27,28,29]. This is mainly because of their variable valence states and unique electronic structures, which enable them to exhibit catalytic performance comparable to that of noble-metal-based catalysts. Some significant works published in recent two years are summarized in Table 1. The most widely investigated transition metal catalysts can be roughly categorized into oxides, phosphides, sulfides, and borides.
Transition metal oxides, including perovskites, spinels, and layered double hydroxides (LDHs), have been extensively studied for their potential as electrocatalysts for the OER. The use of perovskite materials in the OER was first demonstrated by Bockris and Otagawa in 1984 [51], and subsequently, Suntivich et al. proposed them as promising candidates for OER applications [52]. Metal oxide perovskites (ABO3) are structurally stable and have a wide compositional range. Generally, substitutions at the A- and B-sites primarily influence the ability to adsorb oxygen and the reactivity of adsorbed oxygen, respectively [53]. Matienzo et al. synthesized six perovskites compounds (LaFeO3, LaCoO3, LaNiO3, PrCoO3, Pr0.8Sr0.2CoO3, and Pr0.8Ba0.2CoO3) by co-precipitation and thermal annealing methods [54]. Among these perovskites, LaNiO3 exhibited the best OER performance within La-based compounds, while Pr0.8Sr0.2CoO3 had the optimal OER activity within Pr-based perovskites.
Spinel oxides have a structural formula of AB2O4, where A and B typically represent divalent and trivalent cations, respectively, located in the oxygen pores of tetrahedrons and octahedrons, respectively. Transition-metal spinel materials, such as Co3O4 and its derivatives, have shown excellent OER activity thanks to their three-dimensional (3D) structures [55,56]. Qiao’s group developed a 3D hybrid paper by loading spinel NiCo2O4 nanosheets on N-doped graphene film [57]. The edges of NiCo2O4 and N(O)-metal (Ni or Co) bonds were identified as active sites which were responsible for the excellent OER performance. Additionally, the incorporation of metals such as Mn, Zn, Cu, and Fe into spinel Co3O4 was shown to enhance its OER activity [58].
LDHs typically adopt a brucite-type structure, consisting of metal hydroxide layers with intercalated anions or water molecules. The metal layer comprises octahedral MO6 units with shared edges, where two types of metal ions occupy the layer. One valence state is M2+ (M = Mg, Ca, Mn, Fe, Co, Ni, Cu, and Zn), and the other valence state is M3+ (M = Al, Mn, Fe, Co, and Ni). As a result, the metal layers are positively charged [59]. To neutralize this positive charge, anions such as Cl, Br, NO3, and SO42− are inserted into the interlayer spacing to increase the interlayer distance [60]. Due to their tunable composition, facile tailoring of structure, and easy functionalization with other materials, LDHs have shown significant advantages in the field of electrocatalysis. The thickness of LDH catalysts significantly affects their OER activity. Zhang et al. fabricated porous monolayer NiFe-LDH nanosheets with a thickness of 0.8 nm through a facile one-step strategy [61]. This NiFe-LDH exhibited a remarkably low overpotential of 230 mV and a Tafel slope of only 47 mV dec−1. Besides NiFe-LDH, other LDH materials, such as NiCo-LDH and CoFe-LDH, have also emerged as potential OER catalysts [62,63]. Sun et al. synthesized hierarchical NiCo-LDH hollow nanopolyhedra by a facile self-templated method [64]. Density functional theory (DFT) calculations revealed that Co3+ hollow sites served as the adsorption and active sites in the OER process. Pei et al. grew CoFe-LDH nanosheets on nickel foam (NF) to form a 3D hierarchical structure via electrodeposition. Such CoFe-LDH exhibited excellent OER activity with an overpotential of 250 mV and a Tafel slope of 35 mV dec−1 [65].
In recent years, transition metal sulfides (NixSy, CoxSy, and FexSy), phosphides (NixPy, CoxPy, and FexPy), and borides have also shown excellent catalytic activity in the field of electrocatalytic OER [66,67,68,69,70,71]. Hexagonal 2D CoS nanosheets with varying thicknesses were synthesized by chemical vapor deposition [72]. The thinnest 5 nm-CoS nanosheets showed optimal OER performance (290 mV at 10 mA cm−2) due to more exposed catalytic active sites, stronger intermediate adsorption, and promoted electron transfer. Nitrogen-doped carbon-decorated CoP@FeCoP yolk-shelled micro-polyhedra were prepared via phosphorization of ZIF-67@Co-Fe Prussian blue analogues, which only required 238 mV overpotential to reach 10 mA cm−2 and displayed outstanding long-term stability [73]. Such excellent OER activity could be ascribed to the increased specific surface area and active sites resulting from the unique yolk-shell structure and carbon matrix. Sun et al. used the chemical reduction method to in situ deposit tungsten-iron-nickel-boron (W-Fe-Ni-B) nanoparticles on the surface of a NF substrate, creating WFeNiB/NF catalyst [74]. A low overpotential of only 223 mV could afford a current density of 10 mA cm−2, and the corresponding Tafel slope was 38.8 mV dec−1. Furthermore, it exhibited excellent long-term stability, as indicated by a mere 6 mV shift after 10 h of continuous operation, which could be attributed to the enhanced corrosion resistance of borate species in the catalyst.

3. Synergistic Modulation on Transition-Metal-Based Electrocatalysts for OER

Despite significant efforts dedicated to the development of various transition-metal OER electrocatalysts, achieving industrial-level OER performance remains a formidable challenge. Applying modification to the electronic structure, morphology, crystalline, elementary reactions, and external fields to synergistically modulate intrinsic activity, active sites number, conductivity, mass diffusion, the free energy of intermediates adsorption, and external forces has been accepted as an effective approach to optimizing OER catalytic performance.

3.1. Metal Active Sites and Heterogeneous Atoms

The electrocatalytic OER performance is known to be influenced by intrinsic activity, which is determined by the energy barriers associated with the adsorption and desorption of oxygen-containing intermediates [75]. Nørskov’s theory proposes that the difference in adsorption energy between *O to *OH (∆G*O − ∆G*OH) serves as a descriptor for OER activity (Figure 2a), which can be balanced by electronic modulation [76]. The introduction of heteroatoms, including both metal and non-metal atoms, is a highly effective strategy for tuning the electronic structure [77,78,79]. To effectively adjust the electronic structure of the reactive sites, the heteroatom must possess a relatively low electronegativity or induce an abundance of electrons.

3.1.1. Metal Active Sites and Heterogeneous Metal Atoms

Sargent’s group demonstrated that introducing metallic dopants (W, Mo, Nb, Ta, Re, and MoW) with high-valence charges can lower the energy barriers for valence charge transition in 3d metals, such as Fe, Co, and Ni, thereby improving catalytic OER performance (Figure 2b) [80]. Furthermore, the adsorption energy of NiFe-LDH towards oxygen intermediates can also be optimized after the introduction of those high-valence metals. Zhao’s group investigated the in situ structural reconstruction from V-doped Ni2P to NiV oxyhydroxides, where the synergistic interaction between Ni hosts and V dopants can modulate the electronic structure of NiV oxyhydroxides, facilitating the adsorption of *OH and deprotonation of *OOH intermediates (Figure 2c,d) [81]. Wang et al. incorporated high-valence state tantalum (Ta) into the pristine NiFe-LDH through the hydrothermal method [82]. Structural characterizations and DFT results revealed that Ta doping induced electronic structure modulation around Ni, Fe, and Ta, and the eg orbital of Ta, resulting from charge transfer, promoted the adsorption of OH species on Ta sites and improve the conductivity of NiFe-LDH. It is worth noting that surface reconstruction can occur easily before OER. Rare earth metals containing unique 4f sub-shell electrons have also attracted significant attention [83,84]. Sun et al. synthesized Ce-doped LaNiO3 and found that low-concentration Ce doping at the A-site can promote surface reconstruction into a highly active NiOOH phase by optimizing the O 2p level [85].
In addition to doping, loading single atoms on catalyst surfaces has emerged as a promising strategy to achieve outstanding catalytic properties by utilizing low-coordination and unsaturated active sites [86]. Wang et al. uniformly anchored single Ir atoms on the outermost surface of the NiO matrix (Figure 3a,b). As shown in Figure 3c, DFT calculations indicated that the substituted single Ir atom served as an OER active site and activated nearby surface Ni sites because of the excess electrons contributed by the Ir atom, which synergistically enhanced the OER activity of NiO [87]. They also reported that the Ir-single-atom decorated Ni2P catalyst exhibited an ultralow overpotential of 149 mV to achieve a current density of 10 mA cm−2 for OER, which could be ascribed to the optimized adsorption and desorption of OH intermediate species on Ir–O–P/Ni–O–P coordination sites [69]. Additionally, Chen et al. anchored Pt single atoms inside interlayers of NiFe-LDHs. During the OER process, intercalated Pt single atoms not only promoted surface reconstruction but also optimized the intrinsic activity of Ni2+δ–O–Fe3+ζOxHy in the active phase. The interaction between Pt single atoms and Ni2+δ/Fe3+ζ species further promoted OER activity [88].

3.1.2. Metal Active Sites and Heterogeneous Non-Metal Atoms

Opposite to metal elements, highly electronegative non-metal atoms can attract electrons from metals to form adsorption sites for oxygen-containing intermediates during the OER process. Li et al. demonstrated the OER performance of NiFeP catalysts can be improved by partially replacing P with S [89]. The formation of metal-sulfur bonds modulated the electronic structure of the catalysts, leading to a decrease in the energy barrier during the adsorption process and reaction pathway of OER. Moreover, S doping facilitated the generation of *OOH and the release of O2 during the OER process. N-doped NiS2 exhibited enhanced OER activity because of its well-defined morphology, fast charge transfer, and enriched N doping [90] (Figure 4a). Specifically, the presence of N atoms adjacent to the active sites of Ni shifted the position of Ni d-states closer to the Fermi level, and the strong electron-withdrawing property of N atoms endowed adjacent Ni atoms with a higher oxidation state (Figure 4b). Moreover, the introduction of N atoms also promoted the value of (∆G*OH − ∆G*OOH) close to the volcano center, which indicated optimized adsorption energy towards oxygen-containing intermediates during the OER process (Figure 4c,d). Additionally, halogen atoms (F, Cl, and Br) have been proven to effectively modulate the electronic structure of the matrix to improve OER performance [91,92,93,94,95].

3.2. Heterogeneous Atoms and Crystallographic Structure

In addition to the conventional method of incorporating heterogeneous atoms, manipulating the crystalline nature of catalysts provides an alternative approach to modulating their electronic properties. By synergistically combining heteroatoms with the creation of vacancies, lattice distortion, and grain boundaries, researchers have demonstrated the effectiveness of this strategy in precisely adjusting the electronic properties of catalysts, leading to significant improvements in catalytic performance or even the mechanism of OER.

3.2.1. Synergistic Modulation on Heterogeneous Atoms and Cation/Anion Vacancies

The construction of oxygen vacancies is a prevalent strategy in the design of transition-metal-based catalysts, owing to their low formation energy. An oxygen vacancy is a type of point defect that arises from the removal of oxygen atoms in the metal oxide lattice without causing a phase transition. The resultant reduction in oxygen concentration induces electron deficiency in neighboring metal species, leading to a redistribution of electron density towards the metal atoms and a subsequent reduction in electron density around oxygen atoms. This electronic modulation promotes the interaction between hydroxyl ions and OER reaction intermediates. As shown in Figure 5a–d, Yang et al. incorporated N doping and oxygen vacancy into the Co3O4 catalyst and demonstrated that N atoms redistributed electronic configuration of Co atoms to facilitate OER kinetics, while generating rich oxygen vacancies could activate lattice oxygen oxidation mechanism during the OER process [96]. The synergistic effect of N doping and oxygen vacancies optimized the adsorption behavior of oxygen-containing intermediates (Figure 5e,f). Additionally, electronic states can be regulated by integrating heterogeneous metal atoms with oxygen vacancies [97]. Li et al. constructed W-doped NiFeW-LDHs with oxygen vacancies on nickel foam and demonstrated that the weakening of metal-oxygen bonds and the shift of the O 2p band center towards the Fermi level induced the formation of oxygen vacancies, thereby enhancing the adsorption capacity of OER intermediates [46]. The positive shift of the d-band center and generation of oxygen vacancies enhanced the adsorption capacity of intermediates in the OER process. While anion vacancies (e.g., P, S, and Se) have received considerable attention [98,99,100], cationic vacancies have been relatively less explored due to their higher hopping barriers. Recent studies indicated that cationic vacancies can play a similar role to their anionic counterparts in improving OER activity [101,102,103]. For example, Zhao et al. designed and synthesized highly efficient Fe-doped La0.5Sr0.5−δCoO3 with Sr vacancies for OER and proposed that the synergistic effect of Fe active sites and Sr vacancies activated the lattice oxygen mechanism [104]. Theoretical calculations revealed that surface Fe sites acted as the catalytic centers to trigger lattice OER, while Sr vacancies could promote oxidation of surface lattice oxygen through uplifting O 2p levels to facilitate OER.

3.2.2. Heterogeneous Atoms and Lattice Distortion/Grain Boundaries

Lattice distortion and grain boundary engineering have emerged as effective strategies for boosting the kinetics of OER by creating additional active sites. Liao et al. explored the introduction of cerium (Ce) atoms into NiFe-LDH to induce lattice distortion, as depicted in Figure 6a,b [105]. Experimental and theoretical results demonstrated that the incorporation of Ce and lattice distortion regulated the electronic structure of Ni atoms in active sites and lowered the Gibbs free energy of the potential-determining step: *OH → *O (Figure 6c,d). Additionally, the creation of a high density of grain boundaries has been proposed as a promising strategy for augmenting the number of active sites for OER due to the loose distribution of atoms along these boundaries [106,107]. Qiao et al. synthesized (FexCo1-x)B OER electrocatalyst with controllable grain boundary density [108]. Physical characterizations and DFT calculations confirmed that the presence of Fe atoms and manipulation of grain boundaries could effectively modulate the electronic states and provide more efficient active sites, respectively, thus synergistically enhancing the OER process.
Inducing heteroatoms and creating lattice distortion also can cause lattice strain due to the change in atom–atom bond length or by the induced lattice mismatch. The electronic structure of the catalysts’ surface is sensitive to lattice strain, which makes strain a useful strategy for regulating electrocatalysis [109,110,111]. Ma et al. induced tunable lattice strain into NiFeMo alloys through dual doping of Mo and Fe, which in turn changed d-band center and electronic interaction on catalytic active sites, thus improving OER performance [112]. In addition, combining lattice strain with other modifications can synergistically modulate the OER property of catalysts. Liu et al. investigated the coupling effect of lattice strain and oxygen defects on electrocatalytic OER activity of La0.7Sr0.3CoO3−δ thin films [113]. Experimental results and computational calculations indicated that excessive oxygen defects induced by strain increased the eg state occupancy and expanded the energy gap between Co 3d and O 2p bands, leading to lower OER activity.

3.3. Electronic Structure and Morphology

The OER enhancement strategies mentioned above mainly involve regulating the electronic structure, conductivity, and adsorption-free energy of the active intermediate species. On the other hand, regulating morphology is another effective strategy to improve the OER efficiency of transition-metal catalysts by increasing specific surface area, exposing more active sites, and accelerating the release of bubbles. By integrating morphology engineering with electronic modulation to enhance intrinsic activity, catalytic performance can be further boosted.
Cao et al. synthesized sub-2 nm NiFeCr trimetallic hydroxide nanodots that exhibited a low overpotential of 271 mV at 10 mA cm−2 for OER [114]. This excellent catalytic performance could be attributed to the synergistic effect of strong electronic interaction among three metals and zero-dimensional morphology. Moreover, one, two, and multi-dimensional materials also have the advantages of large specific surface area and sufficient exposure to edge sites and are regarded as ideal electrocatalyst materials. For example, Wang et al. decorated W single atoms on the substrate of NiS0.5Se0.5 nanosheet/NiS0.5Se0.5 nanorod heterostructure through a solvothermal method (Figure 7a) [36]. The structure of nanosheets and nanorods exposed more active sites and promoted electrolyte diffusion (Figure 7b). Meanwhile, the incorporated W single atoms delocalized the spin state of Ni (Figure 7c), leading to an increased d-electron density that caused a significant reduction in the adsorption free energy of the rate-determining step (*O → *OOH) (Figure 7d–f). These two main factors contributed to significant OER performance enhancement. Constructing a hierarchical pore structure with interconnected macropores, mesopores, and small pores can greatly increase the specific surface area and expose abundant active sites, which have become the main target for morphology regulation of transition metal oxide catalysts. For instance, Wan et al. developed meso–macro hierarchical porous Ni3S4 architectures derived from Ni metal–organic frameworks as an OER electrocatalyst [115]. The outstanding OER performance can be ascribed to the synergistic effect of enhanced chemisorption of OH resulting from Ni3+ formation, more active sites, and faster mass transfer.

3.4. Synergistic Modulation on Elementary Reactions

The adsorbate evolution mechanism (AEM) and the lattice-oxygen-mediated mechanism (LOM) are two well-established mechanisms that play crucial roles in OER. In alkaline media, the AEM involves a series of four concurrent proton-electron transfer reactions, where metal atoms act as reaction centers. These reactions can be described by the following equations [116]:
OH + * → *OH + e
*OH → *O + e + H+
*O + OH → *OOH + e
*OOH → * + O2(g) + e + H+
The scaling relation among the reaction intermediates in the AEM pathway imposes a theoretical lower limit of 0.37 eV on the overpotential [117,118]. Three strategies have been proposed to break this scaling relation to obtain better activity: (1) stabilizing OER intermediate *OOH while maintaining the adsorption of *OH; (2) inducing a proton acceptor to regulate the reaction pathway; and (3) activating lattice oxygen for direct coupling of O−O radical, which is also known as LOM. This means that the LOM can bypass the formation of *OOH, and thus the limitation in scaling relation between *OH and *OOH can be avoided. For example, amorphous NiFeP nanostructures were fabricated for highly active and stable OER electrocatalysts [119]. The electronic structure of metal sites could be modulated by the ligand effect of P, consequently breaking the scaling relationships among these OER intermediates. Specifically, the adsorption energy gap between *OH and *OOH can be reduced from 3.08 to 2.62 eV by the incorporation of P atoms in NiFeOOH, which resulted in the shift of rate-determining step for OER from the formation of *O to *OH. Similarly, Liu et al. synthesized S-doped NiFe2O4 nanocone arrays which showed a current density of 100 mA cm−2 with an overpotential of 270 mV, which was superior to reported spinel-type oxides [120]. The calculation results demonstrated that the PDOS of Ni-d of Ni atoms adjacent to S atoms was localized near the Fermi level, suggesting that the coupling of Ni-d orbitals and 2p orbitals of oxygen-containing intermediates was promoted. In NiFe2O4, *O to *OOH is the rated-determining step (RDS). DFT calculations revealed that the energy barrier of RDS on the Ni site decreased to 0.25 eV, significantly lower than that on the Fe site, after the introduction of S. These findings suggest that S doping imparts appropriate electronic states and enhanced adsorption capabilities to Ni sites, breaking the scaling relation during the OER process.
Introducing a second component on host materials to form a heterostructure is considered a simple and effective route to design efficient OER electrocatalysts [121,122]. In contrast to heteroatom doping, which necessitates limiting the number of dopants to a low level (typically, <10% of the total elements) to avoid the emergence of new crystal phases that could impact the original structures and block active sites, heterostructures offer several advantages. These include synergistic effects, strain effects, and electronic interactions, all of which contribute to enhanced catalytic performance [123,124,125]. The strong interaction in the heterostructure has been proven to effectively modify the local electronic configuration around active sites and optimize the adsorption/desorption energy of intermediates on different components [124,126]. For instance, Zhao et al. designed an Ir/Ni(OH)2 heterostructure which required merely 224 mV to reach 10 mA cm−2 because of the synergistic effect between the active species of IrOx and NiOOH (Figure 8a–c) [127]. Electrochemical analyses and theoretical calculations provided evidence that the formation of Ir–O–Ni bridging bonds across the interface, along with the in situ formation of IrOx and NiOOH (Figure 8d–f), stabilized metastable Ir4+ species, which were highly active for O–O bond formation. Moreover, OER intermediates, *OH and *O, could be adsorbed on Ni(OH)2/NiOOH and IrOx, respectively (Figure 8g). As a result, the combined effect of these two components promoted the formation of *OOH, which broke the scaling relation and led to accelerated OER kinetics.
Moreover, loading metal nanoparticles on supporting materials can trigger strong metal-support interaction (SMSI) which not only disperses and stabilizes metal sites to avoid aggregation but also creates electron transfer between metal nanoparticles and support, thus significantly changing electronic states of each component and optimizing the adsorption of intermediates [128,129,130]. Gorlin et al. investigated the interaction between MnOx and Au nanoparticles and found that the Au/MnOx showed an order of magnitude high turnover frequency than that of pristine MnOx [131]. In situ Mn L-edge X-ray absorption spectra indicated the enhanced OER performance resulting from local and interfacial effects. More specifically, reduced MnOx and oxidized metal at the interface caused by electron transfer between Au and MnOx optimized the adsorption of *OH and thereby enhanced OER activity. Similarly, Zhang et al. deposited ultrafine Ag nanoparticles on Co(OH)2 nanosheets by spontaneous redox reaction and found that the as-synthesized Ag@Co(OH)2 exhibited a low overpotential of 250 mV to reach current density of 10 mA cm−2, which was better than that of pristine Co(OH)2 and commercial IrO2 [132]. DFT results indicated that the electronic configuration of metallic Ag was tuned by underlying Co(OH)2 support because of SMSI, which provided reduced energetic barriers for the oxygen-contained intermediates, thus promoting OER catalytic process.

3.5. Synergistic Modulation on External Fields

Field-assisted electrocatalysis has emerged as a promising technique for enhancing electrochemical reactions, particularly in the context of OER. This technique utilizes external factors such as magnetic fields, strain, and light to provide additional means of engineering and optimizing the OER process.
Theoretical explanations for magnetic field-assisted OER primarily involve three key effects: magnetothermal, spin-polarized, and electron energy state enhancement effects. The overall OER performance can be improved by increasing the surface temperature of catalysts, optimizing the adsorption thermodynamic features of reactants and intermediates, and accelerating electron transfer. For example, Garcés-Pineda et al. conducted a comprehensive investigation on the influence of an external static magnetic field on a series of transition metal oxides during the electrocatalytic OER process in an alkaline electrolyte [133]. As depicted in Figure 9a,b, they found that the external magnetic field had negligible influence on non-magnetic catalysts but significantly enhanced the performance of materials with strong magnetism. The OER current of NiZnFe4Ox was significantly boosted under the magnetic field. Specifically, the increase in current density reached nearly 100% at a potential of 1.67 V (Figure 9c). Upon turning off the magnetic field, a clear transition of the anodic current from a high-current state to a low-current state was observed, providing further confirmation of the magnetic field effect (Figure 9d). They proposed that the magnetic field contributed to the parallel alignment of oxygen radicals during the formation of O–O bond which dominated OER mechanistic pathway under alkaline conditions.
Light-assisted electrocatalytic OER involves two primary mechanisms: photocarrier and photothermy. When photosensitive materials are subjected to light irradiation, the carriers become excited, facilitating the overcoming of potential barriers in charge transfer and redox reactions [134,135]. Thus, coupling photo-excited carriers with electrochemical reactions can significantly accelerate catalytic rates. Bai et al. successfully hybridized CoFe-LDH with WO3/SnSe2 n–p heterojunction and demonstrated that the overpotential for OER could be decreased by 80 mV under simulated sunlight irradiation [136]. During the OER process, photo-generated holes on the valence band of SnSe2 would be transferred to CoFe-LDH and oxidize Co/Fe into higher valence states (Figure 10a). Consequently, OH could rapidly adsorb on metal sites and undergo deprotonation to form *O species. In other words, the photoelectric synergy system in the heterojunction led to a reduction in the energy barrier for OER and a remarkable acceleration of the OER kinetics.
In solar light-assisted electrocatalysis, the photothermal effect represents another critical aspect that can provide an additional driving force, namely thermal energy, to reduce activation energy, thus promoting the electrochemical reaction kinetics [137,138,139]. Photo-sensitive materials, including plasmonic metals, semiconductors, and carbon materials, can respond to solar light and generate in situ thermal energy to promote electron transfer. For example, Liang et al. synthesized a self-supported reduced graphene oxide (rGO) film with abundant carbon defects and broad light absorption [140]. As illustrated in Figure 10b, rGO not only accelerated electron and mass transfers but also served as a substrate to provide active sites and thermal sources. The surface temperature increased to 70.9 °C in 1 min under simulated sunlight irradiation. With the assistance of sunlight, this catalyst exhibited a low OER overpotential of 215 mV, which was 93 mV lower than that without irradiation.
Figure 10. Schematic illustrations of (a) the proposed mechanism of CoFe-LDH with WO3/SnSe2 system for OER in alkaline media. Reproduced from [136] with permission from John Wiley and Sons License; (b) rGO for broad-spectrum solar enhanced OER. Reproduced from [140] with permission from Elsevier.
Figure 10. Schematic illustrations of (a) the proposed mechanism of CoFe-LDH with WO3/SnSe2 system for OER in alkaline media. Reproduced from [136] with permission from John Wiley and Sons License; (b) rGO for broad-spectrum solar enhanced OER. Reproduced from [140] with permission from Elsevier.
Catalysts 13 01230 g010

4. Conclusions and Perspective

The pursuit of cheap renewable energy to replace traditional fossil fuels remains a primary goal. Electrochemical OER plays a key role in energy conversion and storage in conjunction with renewable energy, such as solar, and wind. In this review, we presented a summary of recent progress in the synergistic modulation of heterogeneous atoms, defects, morphology, elemental reactions, and external fields for advancing transition-metal-based OER electrocatalysts (Figure 11). Synergistic modulation among these aspects can simultaneously optimize the electronic structure of catalysts, adsorption/desorption energy of oxygen-containing intermediates, charge transfer, mass diffusion, and chemical/structural stability. This synergy system provides insights into the design of next-generation catalysts.
Despite significant progress in the development of electrocatalytic OER, there remain considerable challenges and opportunities. These include: (1) Developing scalable fabrication approaches for synergistically engineering the electronic configuration, defects, morphology, elemental reactions, and external fields to meet the requirements of practical applications. (2) Gaining a deeper understanding of the real active sites and catalytic mechanisms in synergy systems. The complex structural reconstruction and dynamic electron-loss processes involved in oxygen gas generation pose challenges for experimental characterization and detection. Advanced ex situ/in situ techniques are required to probe the real state of active sites and intermediates during the OER process. (3) Seawater oxidation. Seawater splitting is a promising approach to large-scale green hydrogen production due to the abundance of seawater supply. However, special attention is required in the design of synergy systems to address competition and corrosion problems from chlorine evolution reaction. (4) Ensuring chemical and structural stability under large-current-density operating conditions. The OER performance of electrocatalysts is typically evaluated under a low current density of 10 mA cm−2, which does not reflect practical industrial applications (>1 A cm−2). More efforts should be directed towards improving chemical and structural durability. (5) Coupling electrocatalysis with external fields. Integrating external driving forces such as magnetic field, light, and strain may provide new insights into modifying catalytic performance beyond the electrocatalyst itself. Addressing these challenges and opportunities will require a multidisciplinary approach, including materials science, chemistry, and engineering, to develop next-generation electrocatalysts with optimized performance for practical applications.

Author Contributions

Conceptualization: Z.L. and L.Y.S.L.; formal analysis: Y.W. and Z.L.; writing—original draft preparation: Z.L.; writing—review and editing: Y.W. and L.Y.S.L.; funding acquisition: L.Y.S.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial supports from the Hong Kong Polytechnic University (1-ZVST) and Shenzhen Key Basic Research Project, China (JCYJ20220818102210023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, Z.P.; Lu, X.F.; Zang, S.Q.; Lou, X.W. Non-noble-metal-based electrocatalysts toward the oxygen evolution reaction. Adv. Funct. Mater. 2020, 30, 1910274. [Google Scholar] [CrossRef]
  2. Chen, F.Y.; Wu, Z.Y.; Adler, Z.; Wang, H.T. Stability challenges of electrocatalytic oxygen evolution reaction: From mechanistic understanding to reactor design. Joule 2021, 5, 1704–1731. [Google Scholar] [CrossRef]
  3. Suen, N.T.; Hung, S.F.; Quan, Q.; Zhang, N.; Xu, Y.J.; Chen, H.M. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.C.; Han, C.D.; Gao, J.; Pan, L.; Wu, J.T.; Zhu, X.D.; Zou, J.J. NiCo-based electrocatalysts for the alkaline oxygen evolution reaction: A Review. ACS Catal. 2021, 11, 12485–12509. [Google Scholar] [CrossRef]
  5. Xie, X.H.; Du, L.; Yon, L.T.; Park, S.Y.; Qiu, Y.; Sokolowski, J.; Wang, W.; Shao, Y.Y. Oxygen evolution reaction in alkaline environment: Material challenges and solutions. Adv. Funct. Mater. 2022, 32, 2110036. [Google Scholar] [CrossRef]
  6. Yang, Y.; Dang, L.N.; Shearer, M.J.; Sheng, H.Y.; Li, W.J.; Chen, J.; Xiao, P.; Zhang, Y.H.; Hamers, R.J.; Jin, S. Highly active trimetallic NiFeCr layered double hydroxide electrocatalysts for oxygen evolution reaction. Adv. Energy Mater. 2018, 8, 1703189. [Google Scholar] [CrossRef]
  7. Subbaraman, R.; Tripkovic, D.; Chang, K.C.; Strmcnik, D.; Paulikas, A.P.; Hirunsit, P.; Chan, M.; Greeley, J.; Stamenkovic, V.; Markovic, N.M. Trends in activity for the water electrolyser reactions on 3d M(Ni,Co,Fe,Mn) hydr(oxy)oxide catalysts. Nat. Mater. 2012, 11, 550–557. [Google Scholar] [CrossRef]
  8. Wu, Z.P.; Zhang, H.B.; Zuo, S.W.; Wang, Y.; Zhang, S.L.; Zhang, J.; Zang, S.Q.; Lou, X.W. Manipulating the local coordination and electronic structures for efficient electrocatalytic oxygen evolution. Adv. Mater. 2021, 33, 2103004. [Google Scholar] [CrossRef]
  9. Hao, Y.M.; Li, Y.F.; Wu, J.X.; Meng, L.S.; Wang, J.L.; Jia, C.L.; Liu, T.; Yang, X.J.; Liu, Z.P.; Gong, M. Recognition of surface oxygen intermediates on NiFe oxyhydroxide oxygen-evolving catalysts by homogeneous oxidation reactivity. J. Am. Chem. Soc. 2021, 143, 1493–1502. [Google Scholar] [CrossRef]
  10. Betley, T.A.; Wu, Q.; Van Voorhis, T.; Nocera, D.G. Electronic design criteria for O-O bond formation via metal-oxo complexes. Inorg. Chem. 2008, 47, 1849–1861. [Google Scholar] [CrossRef]
  11. Hong, W.T.; Risch, M.; Stoerzinger, K.A.; Grimaud, A.; Suntivich, J.; Shao-Horn, Y. Toward the rational design of non-precious transition metal oxides for oxygen electrocatalysis. Energ. Environ. Sci. 2015, 8, 1404–1427. [Google Scholar] [CrossRef]
  12. Moysiadou, A.; Lee, S.; Hsu, C.S.; Chen, H.M.; Hu, X.L. Mechanism of oxygen evolution catalyzed by cobalt oxyhydroxide: Cobalt superoxide species as a key intermediate and dioxygen release as a rate-determining step. J. Am. Chem. Soc. 2020, 142, 11901–11914. [Google Scholar] [CrossRef] [PubMed]
  13. Jiang, J.; Zhang, Y.J.; Zhu, X.J.; Lu, S.; Long, L.L.; Chen, J.J. Nanostructured metallic FeNi2S4 with reconstruction to generate FeNi-based oxide as a highly-efficient oxygen evolution electrocatalyst. Nano Energy 2021, 81, 105619. [Google Scholar] [CrossRef]
  14. Liao, H.X.; Luo, T.; Tan, P.F.; Chen, K.J.; Lu, L.L.; Liu, Y.; Liu, M.; Pan, J. Unveiling role of sulfate ion in nickel-iron (oxy)hydroxide with enhanced oxygen-evolving performance. Adv. Funct. Mater. 2021, 31, 2102772. [Google Scholar] [CrossRef]
  15. Dai, W.J.; Bai, X.W.; Zhu, Y.A.; Zhang, Y.; Lu, T.; Pan, Y.; Wang, J.L. Surface reconstruction induced in situ phosphorus doping in nickel oxides for an enhanced oxygen evolution reaction. J. Mater. Chem. A 2021, 9, 6432–6441. [Google Scholar] [CrossRef]
  16. Wang, C.Z.; Wang, R.; Peng, Y.; Chen, J.J.; Chen, Z.; Yin, H.B.; Li, J.H. Nb-incorporated Fe(oxy)hydroxide derived from structural transformation for efficient oxygen evolution electrocatalysis. J. Mater. Chem. A 2020, 8, 24598–24607. [Google Scholar] [CrossRef]
  17. Guo, T.Q.; Li, L.D.; Wang, Z.C. Recent development and future perspectives of amorphous transition metal-based electrocatalysts for oxygen evolution reaction. Adv. Energy Mater. 2022, 12, 2200827. [Google Scholar] [CrossRef]
  18. Zhang, K.X.; Zou, R.Q. Advanced transition metal-based OER electrocatalysts: Current status, opportunities, and challenges. Small 2021, 17, 2100129. [Google Scholar] [CrossRef]
  19. Peng, X.; Yan, Y.J.; Jin, X.; Huang, C.; Jin, W.H.; Gao, B.; Chu, P.K. Recent advance and prospectives of electrocatalysts based on transition metal selenides for efficient water splitting. Nano Energy 2020, 78, 105234. [Google Scholar] [CrossRef]
  20. Yang, B.P.; Liu, K.; Li, H.J.W.; Liu, C.X.; Fu, J.W.; Li, H.M.; Huang, J.E.; Ou, P.F.; Alkayyali, T.; Cai, C.; et al. Accelerating CO2 electroreduction to multicarbon products via synergistic electric-thermal field on copper nanoneedles. J. Am. Chem. Soc. 2022, 144, 3039–3049. [Google Scholar] [CrossRef]
  21. Wang, Y.H.; Xu, A.; Wang, Z.Y.; Huang, L.S.; Li, J.; Li, F.W.; Wicks, J.; Luo, M.C.; Nam, D.H.; Tan, C.S.; et al. Enhanced nitrate-to-ammonia activity on copper-nickel alloys via tuning of intermediate adsorption. J. Am. Chem. Soc. 2020, 142, 5702–5708. [Google Scholar] [CrossRef] [PubMed]
  22. Koolen, C.D.; Luo, W.; Zuttel, A. From single crystal to single atom catalysts: Structural factors influencing the performance of metal catalysts for CO2 electroreduction. ACS Catal. 2023, 13, 948–973. [Google Scholar] [CrossRef]
  23. Zhang, M.D.; Huang, J.R.; Shi, W.; Liao, P.Q.; Chen, X.M. Synergistic effect in a metal-organic framework boosting the electrochemical CO2 overall splitting. J. Am. Chem. Soc. 2023, 145, 2439–2447. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, J.; Pham, T.H.; Ko, Y.D.; Li, M.; Yang, S.L.; Koolen, C.D.; Zhong, L.P.; Luo, W.; Zuttel, A. Tandem effect of Ag@C@Cu catalysts enhances ethanol selectivity for electrochemical CO2 reduction in flow reactors. Cell Rep. Phys. Sci. 2022, 3, 100949. [Google Scholar] [CrossRef]
  25. Lu, F.; Zhou, M.; Zhou, Y.X.; Zeng, X.H. First-row transition metal based catalysts for the oxygen evolution reaction under alkaline conditions: Basic principles and recent advances. Small 2017, 13, 1701931. [Google Scholar] [CrossRef]
  26. Feng, C.; Faheem, M.B.; Fu, J.; Xiao, Y.Q.; Li, C.L.; Li, Y.B. Fe-based electrocatalysts for oxygen evolution reaction: Progress and perspectives. ACS Catal. 2020, 10, 4019–4047. [Google Scholar] [CrossRef]
  27. Han, L.; Dong, S.J.; Wang, E.K. Transition-metal (Co, Ni, and Fe)-based electrocatalysts for the water oxidation reaction. Adv. Mater. 2016, 28, 9266–9291. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, Y.; Park, B.J.; Paidi, V.K.; Huang, R.; Lee, Y.C.; Noh, K.J.; Lee, K.S.; Han, J.W. Precisely constructing orbital coupling-modulated dual-atom Fe pair sites for synergistic CO2 electroreduction. ACS Energy Lett. 2022, 7, 640–649. [Google Scholar] [CrossRef]
  29. Jia, G.; Wang, Y.; Sun, M.; Zhang, H.; Li, L.; Shi, Y.; Zhang, L.; Cui, X.; Lo, T.W.B.; Huang, B.; et al. Size effects of highly dispersed bismuth nanoparticles on electrocatalytic reduction of carbon dioxide to formic acid. J. Am. Chem. Soc. 2023, 145, 14133–14142. [Google Scholar] [CrossRef]
  30. Chen, M.X.; Li, H.J.; Wu, C.L.; Liang, Y.B.; Qi, J.; Li, J.; Shangguan, E.; Zhang, W.; Cao, R. Interfacial engineering of heterostructured Co(OH)2/NiPx nanosheets for enhanced oxygen evolution reaction. Adv. Funct. Mater. 2022, 32, 2206407. [Google Scholar] [CrossRef]
  31. Liu, Y.H.; Chen, G.Y.; Ge, R.Y.; Pei, K.; Song, C.X.; Li, W.X.; Chen, Y.Y.; Zhang, Y.; Feng, L.Y.; Che, R.C. Construction of CoNiFe trimetallic carbonate hydroxide hierarchical hollow microflowers with oxygen vacancies for electrocatalytic water oxidation. Adv. Funct. Mater. 2022, 32, 2200726. [Google Scholar] [CrossRef]
  32. Song, Y.Y.; Sun, M.Z.; Zhang, S.C.; Zhang, X.Y.; Yi, P.; Liu, J.Z.; Huang, B.L.; Huang, M.H.; Zhang, L.X. Alleviating the work function of vein-like CoXP by Cr doping for enhanced seawater electrolysis. Adv. Funct. Mater. 2023, 33, 2214081. [Google Scholar] [CrossRef]
  33. Wu, J.S.; Yang, T.; Fu, R.; Zhou, M.; Xia, L.X.; Wang, Z.Y.; Zhao, Y. Constructing electrocatalysts with composition gradient distribution by solubility product theory: Amorphous/crystalline CoNiFe-LDH hollow nanocages. Adv. Funct. Mater. 2023, 2300808. [Google Scholar] [CrossRef]
  34. Zhang, H.T.; Guo, H.R.; Li, Y.Y.; Zhang, Q.H.; Zheng, L.R.; Gu, L.; Song, R. The guest doping effects of Fe on bimetallic NiCo layered double hydroxide for enhanced electrochemical oxygen evolution reaction: Theoretical screening and experimental verification. Adv. Funct. Mater. 2023, 2304403. [Google Scholar] [CrossRef]
  35. Song, H.Q.; Yu, J.K.; Tang, Z.Y.; Yang, B.; Lu, S.Y. Halogen-doped carbon dots on amorphous cobalt phosphide as robust electrocatalysts for overall water splitting. Adv. Energy Mater. 2022, 12, 2102573. [Google Scholar] [CrossRef]
  36. Wang, Y.; Li, X.P.; Zhang, M.M.; Zhang, J.F.; Chen, Z.L.; Zheng, X.R.; Tian, Z.L.; Zhao, N.Q.; Han, X.P.; Zaghib, K.R.; et al. Highly active and durable single-atom tungsten-doped NiS0.5Se0.5 nanosheet@NiS0.5Se0.5 nanorod heterostructures for water splitting. Adv. Mater. 2022, 34, 2107053. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, X.; Mao, Z.L.; Mao, X.; Hu, X.M.; Gao, F.Y.; Gao, M.R.; Wu, Q.L.; Lyu, X.; Du, A.J.; Xu, X.S.; et al. Dual integrating oxygen and sulphur on surface of CoTe nanorods triggers enhanced oxygen evolution reaction. Adv. Sci. 2023, 10, 2206204. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, B.; Wu, Z.H.; Shao, W.J.; Gao, Y.; Wang, W.W.; Ma, T.; Ma, L.; Li, S.; Cheng, C.; Zhao, C.S. Interfacial atom-substitution engineered transition-metal hydroxide nanofibers with high-valence Fe for efficient electrochemical water oxidation. Angew. Chem. Int. Edit. 2022, 61, e202115331. [Google Scholar]
  39. Yao, N.; Wang, G.W.; Jia, H.N.; Yin, J.L.; Cong, H.J.; Chen, S.L.; Luo, W. Intermolecular energy gap-induced formation of high-valent cobalt species in CoOOH surface layer on cobalt sulfides for efficient water oxidation. Angew. Chem. Int. Edit. 2022, 61, e202117178. [Google Scholar] [CrossRef]
  40. Jo, S.; Park, W.B.; Lee, K.B.; Choi, H.; Lee, K.S.; Ahn, D.; Lee, Y.W.; Sohn, K.S.; Hong, J.; Sohn, J.I. Bi/BiFe(oxy)hydroxide for sustainable lattice oxygen-boosted electrocatalysis at a practical high current density. Appl. Catal. B-Environ. 2022, 317, 121685. [Google Scholar] [CrossRef]
  41. Meena, A.; Thangavel, P.; Jeong, D.; Singh, A.N.; Jana, A.; Im, H.; Nguyen, D.A.; Kim, K.S. Crystalline-amorphous interface of mesoporous Ni2P@FePOxHy for oxygen evolution at high current density in alkaline-anion-exchange-membrane water-electrolyzer. Appl. Catal. B-Environ. 2022, 306, 121127. [Google Scholar] [CrossRef]
  42. Liao, H.X.; Zhang, X.D.; Niu, S.W.; Tan, P.F.; Chen, K.J.; Liu, Y.; Wang, G.M.; Liu, M.; Pan, J. Dynamic dissolution and re-adsorption of molybdate ion in iron incorporated nickel-molybdenum oxyhydroxide for promoting oxygen evolution reaction. Appl. Catal. B-Environ. 2022, 307, 121150. [Google Scholar] [CrossRef]
  43. Wang, J.Q.; Tran, D.T.; Chang, K.; Prabhakaran, S.; Zhao, J.H.; Kim, D.H.; Kim, N.H.; Lee, J.H. Hierarchical Ni@CNTs-bridged MoxC/Ni2P heterostructure micro-pillars for enhanced seawater splitting and Mg/seawater battery. Nano Energy 2023, 111, 108440. [Google Scholar] [CrossRef]
  44. Lin, L.; Wang, Y.F.; Ye, Q.; Zhao, Y.X.; Cheng, Y.L. Rapid fabrication of FexNi2-xP4O12 and graphene hybrids as electrocatalyst for highly efficient oxygen evolution reaction. Appl. Catal. B-Environ. 2023, 334, 122834. [Google Scholar] [CrossRef]
  45. Xia, L.C.; Bo, L.L.; Shi, W.P.; Zhang, Y.N.; Shen, Y.X.; Ji, X.C.; Guan, X.L.; Wang, Y.X.; Tong, J.H. Defect and interface engineering of templated synthesis of hollow porous Co3O4/CoMoO4 with highly enhanced electrocatalytic activity for oxygen evolution reaction br. Chem. Eng. J. 2023, 452, 139250. [Google Scholar] [CrossRef]
  46. Li, H.X.; Zhang, C.Y.; Xiang, W.J.; Amin, M.A.; Na, J.; Wang, S.P.; Yu, J.X.; Yamauchi, Y. Efficient electrocatalysis for oxygen evolution: W-doped NiFe nanosheets with oxygen vacancies constructed by facile electrodeposition and corrosion. Chem. Eng. J. 2023, 452, 139104. [Google Scholar] [CrossRef]
  47. Wen, S.T.; Huang, J.; Li, T.T.; Chen, W.; Chen, G.L.; Zhang, Q.; Zhang, X.H.; Qian, Q.Y.; Ostrikov, K. Multiphase nanosheet-nanowire cerium oxide and nickel-cobalt phosphide for highly-efficient electrocatalytic overall water splitting. Appl. Catal. B-Environ. 2022, 316, 121678. [Google Scholar] [CrossRef]
  48. Zhang, L.L.; Rong, J.; Yang, Y.Q.; Zhu, H.Z.; Yu, X.H.; Chen, C.L.; Cheng, H.M.; Liu, G. Activated FeS2@NiS2 core-Shell structure boosting cascade reaction for superior electrocatalytic oxygen evolution. Small 2023, 19, 2207472. [Google Scholar] [CrossRef] [PubMed]
  49. Zhai, Y.Y.; Ren, X.R.; Sun, Y.; Li, D.; Wang, B.L.; Liu, S. Synergistic effect of multiple vacancies to induce lattice oxygen redox in NiFe-layered double hydroxide OER catalysts. Appl. Catal. B-Environ. 2023, 323, 122091. [Google Scholar] [CrossRef]
  50. Ai, T.T.; Wang, H.H.; Bao, W.W.; Feng, L.L.; Zou, X.Y.; Wei, X.L.; Liu, D.; Deng, Z.F.; Bin, R. Fe-V synergistic doping effect of hierarchical Ni3S2 oblate-nanorod arrays for efficient electrocatalytic oxygen evolution reaction. Chem. Eng. J. 2022, 450, 138358. [Google Scholar] [CrossRef]
  51. Bockris, J.O.; Otagawa, T. The electrocatalysis of oxygen evolution on perovskites. J. Electrochem. Soc. 1984, 131, 290–302. [Google Scholar] [CrossRef]
  52. Suntivich, J.; May, K.J.; Gasteiger, H.A.; Goodenough, J.B.; Shao-Horn, Y. A perovskite oxide pptimized for oxygen evolution catalysis from molecular orbital principles. Science 2011, 334, 1383–1385. [Google Scholar] [CrossRef] [PubMed]
  53. Dias, J.A.; Andrade, M.A.S.; Santos, H.L.S.; Morelli, M.R.; Mascaro, L.H. Lanthanum-based perovskites for catalytic oxygen evolution reaction. ChemElectroChem 2020, 7, 3173–3192. [Google Scholar] [CrossRef]
  54. Matienzo, D.D.; Kutlusoy, T.; Divanis, S.; Di Bari, C.; Instuli, E. Benchmarking perovskite electrocatalysts’ OER activity as candidate materials for industrial alkaline water electrolysis. Catalysts 2020, 10, 1387. [Google Scholar] [CrossRef]
  55. Liu, W.; Kamiko, M.; Yamada, I.; Yagi, S. Electrochemical deposition of amorphous cobalt oxides for oxygen evolution catalysis. RSC Adv. 2022, 12, 8731–8736. [Google Scholar] [CrossRef] [PubMed]
  56. Xu, H.; Yuan, J.J.; He, G.Y.; Chen, H.Q. Current and future trends for spinel-type electrocatalysts in electrocatalytic oxygen evolution reaction. Coordin. Chem. Rev. 2023, 475, 214869. [Google Scholar] [CrossRef]
  57. Chen, S.; Qiao, S.Z. Hierarchically porous nitrogen-doped graphene-NiCo2O4 hybrid paper as an advanced electrocatalytic water-splitting material. ACS Nano 2013, 7, 10190–10196. [Google Scholar] [CrossRef]
  58. Harada, M.; Kotegawa, F.; Kuwa, M. Structural changes of spinel MCo2O4 (M = Mn, Fe, Co, Ni, and Zn) electrocatalysts during the oxygen evolution reaction investigated by in situ X-ray absorption spectroscopy. ACS Appl. Energ. Mater. 2022, 5, 278–294. [Google Scholar] [CrossRef]
  59. Yu, J.; Yu, F.; Yuen, M.F.; Wang, C.D. Two-dimensional layered double hydroxides as a platform for electrocatalytic oxygen evolution. J. Mater. Chem. A 2021, 9, 9389–9430. [Google Scholar] [CrossRef]
  60. Zhou, D.J.; Li, P.S.; Lin, X.; McKinley, A.; Kuang, Y.; Liu, W.; Lin, W.F.; Sun, X.M.; Duan, X. Layered double hydroxide-based electrocatalysts for the oxygen evolution reaction: Identification and tailoring of active sites, and superaerophobic nanoarray electrode assembly. Chem. Soc. Rev. 2021, 50, 8790–8817. [Google Scholar] [CrossRef]
  61. Zhang, X.; Zhao, Y.F.; Zhao, Y.X.; Shi, R.; Waterhouse, G.I.N.; Zhang, T.R. A simple synthetic strategy toward defect-rich porous monolayer NiFe-layered double hydroxide nanosheets for efficient electrocatalytic water oxidation. Adv. Energy Mater. 2019, 9, 1900881. [Google Scholar] [CrossRef]
  62. Liang, H.F.; Meng, F.; Caban-Acevedo, M.; Li, L.S.; Forticaux, A.; Xiu, L.C.; Wang, Z.C.; Jin, S. Hydrothermal continuous flow synthesis and exfoliation of NiCo layered double hydroxide nanosheets for enhanced oxygen evolution catalysis. Nano Lett. 2015, 15, 1421–1427. [Google Scholar] [CrossRef] [PubMed]
  63. Karmakar, A.; Karthick, K.; Kumaravel, S.; Sankar, S.S.; Kundu, S. Enabling and inducing oxygen vacancies in cobalt iron layer double hydroxide via selenization as precatalysts for electrocatalytic hydrogen and oxygen evolution reactions. Inorg. Chem. 2021, 60, 2023–2036. [Google Scholar] [CrossRef] [PubMed]
  64. Sun, S.F.; Lv, C.D.; Hong, W.Z.; Zhou, X.; Wu, F.G.; Chen, G. Dual tuning of composition and nanostructure of hierarchical hollow nanopolyhedra assembled by NiCo-layered double hydroxide nanosheets for efficient electrocatalytic oxygen evolution. ACS Appl. Energ. Mater. 2019, 2, 312–319. [Google Scholar] [CrossRef]
  65. Pei, Y.; Ge, Y.C.; Chu, H.; Smith, W.; Dong, P.; Ajayan, P.M.; Ye, M.X.; Shen, J.F. Controlled synthesis of 3D porous structured cobalt-iron based nanosheets by electrodeposition as asymmetric electrodes for ultra-efficient water splitting. Appl. Catal. B-Environ. 2019, 244, 583–593. [Google Scholar] [CrossRef]
  66. Xue, Z.Q.; Li, X.; Liu, Q.L.; Cai, M.K.; Liu, K.; Liu, M.; Ke, Z.F.; Liu, X.L.; Li, G.Q. Interfacial electronic structure modulation of NiTe nanoarrays with NiS nanodots facilitates electrocatalytic oxygen evolution. Adv. Mater. 2019, 31, 1900430. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, C.Z.; Zhu, M.; Cao, Z.; Zhu, P.; Cao, Y.; Xu, X.Y.; Xu, C.; Yin, Z.Y. Heterogeneous bimetallic sulfides based seawater electrolysis towards stable industrial-level large current density. Appl. Catal. B-Environ. 2021, 291, 120071. [Google Scholar] [CrossRef]
  68. Zou, X.X.; Wu, Y.Y.; Liu, Y.P.; Liu, D.P.; Li, W.; Gu, L.; Liu, H.; Wang, P.W.; Sun, L.; Zhang, Y. In situ generation of bifunctional, efficient Fe-based catalysts from mackinawite iron sulfide for water splitting. Chem 2018, 4, 1139–1152. [Google Scholar] [CrossRef]
  69. Wang, Q.; Zhang, Z.; Cai, C.; Wang, M.Y.; Zhao, Z.L.; Li, M.H.; Huang, X.; Han, S.B.; Zhou, H.; Feng, Z.X.; et al. Single iridium atom doped Ni2P catalyst for optimal oxygen evolution. J. Am. Chem. Soc. 2021, 143, 13605–13615. [Google Scholar] [CrossRef]
  70. Chai, L.L.; Hu, Z.Y.; Wang, X.; Xu, Y.W.; Zhang, L.J.; Li, T.T.; Hu, Y.; Qian, J.J.; Huang, S.M. Stringing bimetallic metal-organic framework-derived cobalt phosphide composite for high-efficiency overall water splitting. Adv. Sci. 2020, 7, 1903195. [Google Scholar] [CrossRef]
  71. Li, Y.; Wu, Y.; Hao, H.; Yuan, M.; Lv, Z.; Xu, L.; Wei, B. In situ unraveling surface reconstruction of Ni5P4@FeP nanosheet array for superior alkaline oxygen evolution reaction. Appl. Catal. B: Environ. 2022, 305, 121033. [Google Scholar] [CrossRef]
  72. Zhang, Y.; Zheng, G.B.; Li, A.L.; Zhu, X.K.; Jiang, J.J.; Zhang, Q.; Deng, L.W.; Gao, X.H.; Ouyang, F.P. Hexagonal single-crystal CoS nanosheets: Controllable synthesis and tunable oxygen evolution performance. Inorg. Chem. 2022, 61, 7568–7578. [Google Scholar] [CrossRef] [PubMed]
  73. Shi, J.H.; Qiu, F.; Yuan, W.B.; Guo, M.M.; Lu, Z.H. Nitrogen-doped carbon-decorated yolk-shell CoP@FeCoP micro-polyhedra derived from MOF for efficient overall water splitting. Chem. Eng. J. 2021, 403, 126312. [Google Scholar] [CrossRef]
  74. Sun, Y.H.; Zhao, Y.; Deng, X.Y.; Dai, D.M.; Gao, H.T. An efficient amorphous ternary transition metal boride (WFeNiB) electrocatalyst for oxygen evolution from water. Sustain. Energy Fuels 2022, 6, 1345–1352. [Google Scholar] [CrossRef]
  75. Dickens, C.F.; Kirk, C.; Norskov, J.K. Insights into the electrochemical oxygen evolution reaction with ab initio calculations and microkinetic modeling: Beyond the limiting potential volcano. J. Phys. Chem. C 2019, 123, 18960–18977. [Google Scholar] [CrossRef]
  76. Man, I.C.; Su, H.Y.; Calle-Vallejo, F.; Hansen, H.A.; Martinez, J.I.; Inoglu, N.G.; Kitchin, J.; Jaramillo, T.F.; Norskov, J.K.; Rossmeisl, J. Universality in oxygen evolution electrocatalysis on oxide surfaces. Chemcatchem 2011, 3, 1159–1165. [Google Scholar] [CrossRef]
  77. Niu, S.; Jiang, W.J.; Wei, Z.X.; Tang, T.; Ma, J.M.; Hu, J.S.; Wan, L.J. Se-doping activates FeOOH for cost-effective and efficient electrochemical water oxidation. J. Am. Chem. Soc. 2019, 141, 7005–7013. [Google Scholar] [CrossRef] [PubMed]
  78. Dou, S.; Tao, L.; Huo, J.; Wang, S.Y.; Dai, L.M. Etched and doped Co9S8/graphene hybrid for oxygen electrocatalysis. Energ. Environ. Sci. 2016, 9, 1320–1326. [Google Scholar] [CrossRef]
  79. Wang, Y.; Kong, B.; Zhao, D.Y.; Wang, H.T.; Selomulya, C. Strategies for developing transition metal phosphides as heterogeneous electrocatalysts for water splitting. Nano Today 2017, 15, 26–55. [Google Scholar] [CrossRef]
  80. Zhang, B.; Wang, L.; Cao, Z.; Kozlov, S.M.; de Arquer, F.P.G.; Dinh, C.T.; Li, J.; Wang, Z.Y.; Zheng, X.L.; Zhang, L.S.; et al. High-valence metals improve oxygen evolution reaction performance by modulating 3dmetal oxidation cycle energetics. Nat. Catal. 2020, 3, 985–992. [Google Scholar] [CrossRef]
  81. Zhao, T.W.; Shen, X.J.; Wang, Y.; Hocking, R.K.; Li, Y.B.; Rong, C.L.; Dastafkan, K.; Su, Z.; Zhao, C. In situ reconstruction of V-doped Ni2P pre-catalysts with tunable electronic structures for water oxidation. Adv. Funct. Mater. 2021, 31, 2100614. [Google Scholar] [CrossRef]
  82. Wang, X.Y.; Tuo, Y.X.; Zhou, Y.; Wang, D.; Wang, S.T.; Zhang, J. Ta-doping triggered electronic structural engineering and strain effect in NiFe LDH for enhanced water oxidation. Chem. Eng. J. 2021, 403, 126297. [Google Scholar] [CrossRef]
  83. Zhao, H.Y.; Xia, J.L.; Yin, D.D.; Luo, M.; Yan, C.H.; Du, Y.P. Rare earth incorporated electrode materials for advanced energy storage. Coordin. Chem. Rev. 2019, 390, 32–49. [Google Scholar] [CrossRef]
  84. Chen, P.; Han, W.; Zhao, M.; Su, J.W.; Li, Z.X.; Li, D.Y.; Pi, L.J.; Zhou, X.; Zhai, T.Y. Recent advances in 2D rare earth materials. Adv. Funct. Mater. 2021, 31, 2008790. [Google Scholar] [CrossRef]
  85. Sun, Y.; Li, R.; Chen, X.X.; Wu, J.; Xie, Y.; Wang, X.; Ma, K.K.; Wang, L.; Zhang, Z.; Liao, Q.L.; et al. A-site management prompts the dynamic reconstructed active phase of perovskite oxide OER catalysts. Adv. Energy Mater. 2021, 11, 2003755. [Google Scholar] [CrossRef]
  86. Lee, W.H.; Ko, Y.J.; Kim, J.Y.; Min, B.K.; Hwang, Y.J.; Oh, H.S. Single-atom catalysts for the oxygen evolution reaction: Recent developments and future perspectives. Chem. Commun. 2020, 56, 12687–12697. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, Q.; Huang, X.; Zhao, Z.L.; Wang, M.Y.; Xiang, B.; Li, J.; Feng, Z.X.; Xu, H.; Gu, M. Ultrahigh-loading of Ir single atoms on NiO matrix to dramatically enhance oxygen evolution reaction. J. Am. Chem. Soc. 2020, 142, 7425–7433. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, W.; Wu, B.B.; Wang, Y.Y.; Zhou, W.; Li, Y.Y.; Liu, T.Y.; Xie, C.; Xu, L.T.; Du, S.Q.; Song, M.L.; et al. Deciphering the alternating synergy between interlayer Pt single-atom and NiFe layered double hydroxide for overall water splitting. Energ. Environ. Sci. 2021, 14, 6428–6440. [Google Scholar] [CrossRef]
  89. Li, S.S.; Wang, L.; Su, H.; Hong, A.N.; Wang, Y.X.; Yang, H.J.; Ge, L.; Song, W.Y.; Liu, J.; Ma, T.Y.; et al. Electron redistributed S-doped nickel iron phosphides derived from one-step phosphatization of MOFs for significantly boosting electrochemical water splitting. Adv. Funct. Mater. 2022, 32, 2200733. [Google Scholar] [CrossRef]
  90. Hao, J.H.; Yang, W.S.; Hou, J.W.; Mao, B.D.; Huang, Z.P.; Shi, W.D. Nitrogen doped NiS2 nanoarrays with enhanced electrocatalytic activity for water oxidation. J. Mater. Chem. A 2017, 5, 17811–17816. [Google Scholar] [CrossRef]
  91. Wang, W.H.; Yang, Y.; Huan, D.M.; Wang, L.K.; Shi, N.; Xie, Y.; Xia, C.R.; Peng, R.R.; Lu, Y.L. An excellent OER electrocatalyst of cubic SrCoO3- prepared by a simple F-doping strategy. J. Mater. Chem. A 2019, 7, 12538–12546. [Google Scholar] [CrossRef]
  92. Zhu, Y.L.; Lin, Q.; Wang, Z.B.; Qi, D.C.; Yin, Y.C.; Liu, Y.; Zhang, X.W.; Shao, Z.P.; Wang, H.T. Chlorine-anion doping induced multi-factor optimization in perovskties for boosting intrinsic oxygen evolution. J. Energy Chem. 2021, 52, 115–120. [Google Scholar] [CrossRef]
  93. Zhao, Z.B.; Chang, H.N.; Wang, R.Y.; Du, P.; He, X.; Yang, J.K.; Zhang, X.L.; Huang, K.; Fan, D.Y.; Wang, Y.G.; et al. Activity origin and catalyst design principles for electrocatalytic oxygen evolution on layered transition metal oxide with halogen doping. Small Struct. 2021, 2, 2100069. [Google Scholar] [CrossRef]
  94. Xu, T.Y.; Jiao, D.X.; Zhang, L.; Zhang, H.Y.; Zheng, L.R.; Singh, D.J.; Zhao, J.X.; Zheng, W.T.; Cui, X.Q. Br-induced P-poor defective nickel phosphide for highly efficient overall water splitting. Appl. Catal. B-Environ. 2022, 316, 121686. [Google Scholar] [CrossRef]
  95. Ghanem, M.A.; Amer, M.S.; Al-Mayouf, A.M.; Arunachalam, P.; Weller, M.T. Halide-doping effect of strontium cobalt oxide electrocatalyst and the induced activity for oxygen evolution in an alkaline solution. Catalysts 2021, 11, 1408. [Google Scholar] [CrossRef]
  96. Yang, X.; Cheng, J.; Li, H.; Xu, Y.; Tu, W.F.; Zhou, J.H. Self-supported N-doped hierarchical Co3O4 electrocatalyst with abundant oxygen vacancies for acidic water oxidation. Chem. Eng. J. 2023, 465, 142745. [Google Scholar] [CrossRef]
  97. Li, C.F.; Zhao, J.W.; Xie, L.J.; Wu, J.Q.; Li, G.R. Fe doping and oxygen vacancy modulated Fe-Ni5P4/NiFeOH nanosheets as bifunctional electrocatalysts for efficient overall water splitting. Appl. Catal. B-Environ. 2021, 291, 119987. [Google Scholar] [CrossRef]
  98. Yuan, G.J.; Bai, J.L.; Zhang, L.; Chen, X.; Ren, L.L. The effect of P vacancies on the activity of cobalt phosphide nanorods as oxygen evolution electrocatalyst in alkali. Appl. Catal. B-Environ. 2021, 284, 119693. [Google Scholar] [CrossRef]
  99. Gao, Q.; Luo, W.; Ma, X.Y.; Ma, Z.M.; Li, S.J.; Gou, F.L.; Shen, W.; Jiang, Y.M.; He, R.X.; Li, M. Electronic modulation and vacancy engineering of Ni9S8 to synergistically boost efficient water splitting: Active vacancy-metal pairs. Appl. Catal. B-Environ. 2022, 310, 121356. [Google Scholar] [CrossRef]
  100. Tian, L.; Chen, Z.Y.; Wang, T.J.; Cao, M.; Lu, X.H.; Cheng, W.J.; He, C.C.; Wang, J.; Li, Z. Mo doping and Se vacancy engineering for boosting electrocatalytic water oxidation by regulating the electronic structure of self-supported Co9Se8@NiSe. Nanoscale 2022, 15, 259–265. [Google Scholar] [CrossRef]
  101. Zhu, J.Q.; Ren, Z.Y.; Du, S.C.; Xie, Y.; Wu, J.; Meng, H.Y.; Xue, Y.Z.; Fu, H.G. Co-vacancy-rich Co1-xS nanosheets anchored on rGO for high-efficiency oxygen evolution. Nano Res. 2017, 10, 1819–1831. [Google Scholar] [CrossRef]
  102. Liu, Y.W.; Cheng, H.; Lyu, M.J.; Fan, S.J.; Liu, Q.H.; Zhang, W.S.; Zhi, Y.D.; Wang, C.M.; Xiao, C.; Wei, S.Q.; et al. Low overpotential in vacancy-rich ultrathin CoSe2 nanosheets for water oxidation. J. Am. Chem. Soc. 2014, 136, 15670–15675. [Google Scholar] [CrossRef] [PubMed]
  103. Chen, D.W.; Qiao, M.; Lu, Y.R.; Hao, L.; Liu, D.D.; Dong, C.L.; Li, Y.F.; Wang, S.Y. Preferential cation vacancies in perovskite hydroxide for the oxygen evolution reaction. Angew. Chem. Int. Edit. 2018, 57, 8691–8696. [Google Scholar] [CrossRef] [PubMed]
  104. Zhao, J.W.; Zhang, H.; Li, C.F.; Zhou, X.; Wu, J.Q.; Zeng, F.; Zhang, J.W.; Li, G.R. Key roles of surface Fe sites and Sr vacancies in the perovskite for an efficient oxygen evolution reaction via lattice oxygen oxidation. Energ. Environ. Sci. 2022, 15, 3912–3922. [Google Scholar] [CrossRef]
  105. Liao, Y.Y.; He, R.C.; Pan, W.H.; Li, Y.; Wang, Y.Y.; Li, J.; Li, Y.X. Lattice distortion induced Ce-doped NiFe-LDH for efficient oxygen evolution. Chem. Eng. J. 2023, 464, 142669. [Google Scholar] [CrossRef]
  106. He, Y.M.; Tang, P.Y.; Hu, Z.L.; He, Q.Y.; Zhu, C.; Wang, L.Q.; Zeng, Q.S.; Golani, P.; Gao, G.H.; Fu, W.; et al. Engineering grain boundaries at the 2D limit for the hydrogen evolution reaction. Nat. Commun. 2020, 11, 57. [Google Scholar] [CrossRef] [PubMed]
  107. Gu, Y.H.; Wang, X.Y.; Bao, A.T.; Dong, L.; Zhang, X.Y.; Pan, H.J.; Cui, W.Q.; Qi, X.W. Enhancing electrical conductivity of single-atom doped Co3O4 nanosheet arrays at grain boundary by phosphor doping strategy for efficient water splitting. Nano Res. 2022, 15, 9511–9519. [Google Scholar] [CrossRef]
  108. Qiao, X.S.; Kang, H.J.; Li, Y.; Cui, K.; Jia, X.; Wu, X.H.; Qin, W. Grain boundary density and electronic dual modulation of intermetallic Co2B by Fe doping toward efficient catalyst for oxygen evolution reaction. Appl. Catal. B-Environ. 2022, 305, 121034. [Google Scholar] [CrossRef]
  109. Khorshidi, A.; Violet, J.; Hashemi, J.; Peterson, A.A. How strain can break the scaling relations of catalysis. Nat. Catal. 2018, 1, 263–268. [Google Scholar] [CrossRef]
  110. Wang, L.; Zeng, Z.H.; Gao, W.P.; Maxson, T.; Raciti, D.; Giroux, M.; Pan, X.Q.; Wang, C.; Greeley, J. Tunable intrinsic strain in two-dimensional transition metal electrocatalysts. Science 2019, 363, 870–874. [Google Scholar] [CrossRef]
  111. Wang, H.T.; Xu, S.C.; Tsai, C.; Li, Y.Z.; Liu, C.; Zhao, J.; Liu, Y.Y.; Yuan, H.Y.; Abild-Pedersen, F.; Prinz, F.B.; et al. Direct and continuous strain control of catalysts with tunable battery electrode materials. Science 2016, 354, 1031–1036. [Google Scholar] [CrossRef] [PubMed]
  112. Ma, K.W.; Chang, X.R.; Wang, Z.H.; Deng, R.C.; Wu, X.; Yang, H. Tunable d-band center of a NiFeMo alloy with enlarged lattice strain enhancing the intrinsic catalytic activity for overall water-splitting. Nanoscale 2023, 15, 5843–5854. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, X.; Zhang, L.; Zheng, Y.; Guo, Z.; Zhu, Y.M.; Chen, H.J.; Li, F.; Liu, P.P.; Yu, B.; Wang, X.W.; et al. Uncovering the effect of lattice strain and oxygen deficiency on electrocatalytic activity of perovskite cobaltite thin films. Adv. Sci. 2019, 6, 1801898. [Google Scholar] [CrossRef] [PubMed]
  114. Cao, Y.H.; Li, W.Q.; Guo, H.A.; Yue, M.Y.; Wang, Y.J. Surface synergistic effect of sub-2 nm NiFeCr hydroxide nanodots yielding high oxygen evolution mass activities. Chem. Eng. J. 2023, 461, 141917. [Google Scholar] [CrossRef]
  115. Wan, K.; Luo, J.S.; Zhou, C.; Zhang, T.; Arbiol, J.; Lu, X.H.; Mao, B.W.; Zhang, X.; Fransaer, J. Hierarchical porous Ni3S4 with enriched high-valence Ni sites as a robust electrocatalyst for efficient oxygen evolution reaction. Adv. Funct. Mater. 2019, 29, 1900315. [Google Scholar] [CrossRef]
  116. Rossmeisl, J.; Logadottir, A.; Norskov, J.K. Electrolysis of water on (oxidized) metal surfaces. Chem. Phys. 2005, 319, 178–184. [Google Scholar] [CrossRef]
  117. Montoya, J.H.; Seitz, L.C.; Chakthranont, P.; Vojvodic, A.; Jaramillo, T.F.; Norskov, J.K. Materials for solar fuels and chemicals. Nat. Mater. 2017, 16, 70–81. [Google Scholar] [CrossRef]
  118. Koper, M.T.M. Theory of multiple proton-electron transfer reactions and its implications for electrocatalysis. Chem. Sci. 2013, 4, 2710–2723. [Google Scholar] [CrossRef]
  119. Liu, J.Y.; Liu, X.; Shi, H.; Luo, J.H.; Wang, L.; Liang, J.S.; Li, S.Z.; Yang, L.M.; Wang, T.Y.; Huang, Y.H.; et al. Breaking the scaling relations of oxygen evolution reaction on amorphous NiFeP nanostructures with enhanced activity for overall seawater splitting. Appl. Catal. B-Environ. 2022, 302, 120862. [Google Scholar] [CrossRef]
  120. Liu, H.J.; Luan, R.N.; Li, L.Y.; Lv, R.Q.; Chai, Y.M.; Dong, B. Sulphur-dopant induced breaking of the scaling relation on low-valence Ni sites in nickel ferrite nanocones for water oxidation with industrial-level current density. Chem. Eng. J. 2023, 461, 141714. [Google Scholar] [CrossRef]
  121. Xu, Q.C.; Zhang, J.H.; Zhang, H.X.; Zhang, L.Y.; Chen, L.; Hu, Y.J.; Jiang, H.; Li, C.Z. Atomic heterointerface engineering overcomes the activity limitation of electrocatalysts and promises highly-efficient alkaline water splitting. Energ. Environ. Sci. 2021, 14, 5228–5259. [Google Scholar] [CrossRef]
  122. Li, Y.Q.; Yin, Z.H.; Cui, M.; Liu, X.; Xiong, J.B.; Chen, S.R.; Ma, T.L. Interface engineering of transitional metal sulfide-MoS2 heterostructure composites as effective electrocatalysts for water-splitting. J. Mater. Chem. A 2021, 9, 2070–2092. [Google Scholar] [CrossRef]
  123. Zhang, H.J.; Maijenburg, A.W.; Li, X.P.; Schweizer, S.L.; Wehrspohn, R.B. Bifunctional heterostructured transition metal phosphides for efficient electrochemical water splitting. Adv. Funct. Mater. 2020, 30, 2003261. [Google Scholar] [CrossRef]
  124. Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C.F.; Liu, Z.C.; Kaya, S.; Nordlund, D.; Ogasawara, H.; et al. Lattice-strain control of the activity in dealloyed core-shell fuel cell catalysts. Nat. Chem. 2010, 2, 454–460. [Google Scholar] [CrossRef] [PubMed]
  125. Du, X.C.; Huang, J.W.; Zhang, J.J.; Yan, Y.C.; Wu, C.Y.; Hu, Y.; Yan, C.Y.; Lei, T.Y.; Chen, W.; Fan, C.; et al. Modulating electronic structures of inorganic nanomaterials for efficient electrocatalytic water splitting. Angew. Chem. Int. Edit. 2019, 58, 4484–4502. [Google Scholar] [CrossRef]
  126. Zhang, Y.P.; Gao, F.; Wang, D.Q.; Li, Z.L.; Wang, X.M.; Wang, C.Q.; Zhang, K.W.; Du, Y.K. Amorphous/crystalline heterostructure transition-metal-based catalysts for high-performance water splitting. Coordin. Chem. Rev. 2023, 475, 214916. [Google Scholar] [CrossRef]
  127. Zhao, G.Q.; Li, P.; Cheng, N.Y.; Dou, S.X.; Sun, W.P. An Ir/Ni(OH)2 heterostructured electrocatalyst for the oxygen evolution reaction: Breaking the scaling relation, stabilizing iridium (V), and beyond. Adv. Mater. 2020, 32, 2000872. [Google Scholar] [CrossRef] [PubMed]
  128. Van Deelen, T.W.; Mejia, C.H.; de Jong, K.P. Control of metal-support interactions in heterogeneous catalysts to enhance activity and selectivity. Nat. Catal. 2019, 2, 955–970. [Google Scholar] [CrossRef]
  129. Dong, C.Y.; Li, Y.L.; Cheng, D.Y.; Zhang, M.T.; Liu, J.J.; Wang, Y.G.; Xiao, D.Q.; Ma, D. Supported metal clusters: Fabrication and application in heterogeneous catalysis. ACS Catal. 2020, 10, 11011–11045. [Google Scholar] [CrossRef]
  130. Serna, P.; Gates, B.C. Molecular metal catalysts on supports: Organometallic chemistry meets surface science. Acc. Chem. Res. 2014, 47, 2612–2620. [Google Scholar] [CrossRef]
  131. Gorlin, Y.; Chung, C.J.; Benck, J.D.; Nordlund, D.; Seitz, L.; Weng, T.C.; Sokaras, D.; Clemens, B.M.; Jaramillo, T.F. Understanding interactions between manganese oxide and gold that lead to enhanced activity for electrocatalytic water oxidation. J. Am. Chem. Soc. 2014, 136, 4920–4926. [Google Scholar] [CrossRef] [PubMed]
  132. Zhang, Z.; Li, X.P.; Zhong, C.; Zhao, N.Q.; Deng, Y.D.; Han, X.P.; Hu, W.B. Spontaneous synthesis of silver-nanoparticle-decorated transition-metal hydroxides for enhanced oxygen evolution reaction. Angew. Chem. Int. Edit. 2020, 59, 7245–7250. [Google Scholar] [CrossRef] [PubMed]
  133. Garces-Pineda, F.A.; Blasco-Ahicart, M.; Nieto-Castro, D.; Lopez, N.; Galan-Mascaros, J.R. Direct magnetic enhancement of electrocatalytic water oxidation in alkaline media. Nat. Energy 2019, 4, 519–525. [Google Scholar] [CrossRef]
  134. Meng, H.Y.; Xi, W.; Ren, Z.Y.; Du, S.C.; Wu, J.; Zhao, L.; Liu, B.W.; Fu, H.G. Solar-boosted electrocatalytic oxygen evolution via catalytic site remodelling of CoCr layered double hydroxide. Appl. Catal. B-Environ. 2021, 284, 119707. [Google Scholar] [CrossRef]
  135. Tian, L.; Chen, H.Y.; Lu, X.H.; Liu, D.S.; Cheng, W.J.; Liu, Y.Y.; Li, J.; Li, Z. Local photothermal and photoelectric effect synergistically boost hollow CeO2/CoS2 heterostructure electrocatalytic oxygen evolution reaction. J. Colloid Interf. Sci. 2022, 628, 663–672. [Google Scholar] [CrossRef] [PubMed]
  136. Bai, L.; Jia, S.J.; Gao, Y.D.; Li, C.; Chen, X.; Zhou, S.; Han, J.W.; Yang, F.C.; Zhang, X.; Lu, S.Y. A p-n WO3/SnSe2 heterojunction for efficient photo-assisted electrocatalysis of the oxygen evolution reaction. Energy Environ. Mater. 2022, e12456. [Google Scholar] [CrossRef]
  137. Zhang, Y.J.; Wang, Y.C.; Jiang, H.Q.; Huang, M.H. Multifunctional nickel sulfide nanosheet arrays for solar-intensified oxygen evolution reaction. Small 2020, 16, 2002550. [Google Scholar] [CrossRef]
  138. Zhai, H.C.; Xu, G.S.; Zhu, C.H.; Yuan, Y.P. Understanding the photothermal contribution to electrocatalysis: A case study of carbon supported NiFe layered double hydroxide. Int. J. Hydrogen Energy 2022, 47, 23971–23979. [Google Scholar] [CrossRef]
  139. Jin, B.J.; Li, Y.C.; Wang, J.N.; Meng, F.Y.; Cao, S.S.; He, B.; Jia, S.R.; Wang, Y.; Li, Z.; Liu, X.Q. Promoting oxygen evolution reaction of Co-based catalysts (Co3O4, CoS, CoP, and CoN) through photothermal effect. Small 2019, 15, 1903847. [Google Scholar] [CrossRef]
  140. Liang, Y.G.; Zhang, Y.J.; Wang, X.K.; Zhou, J.; Cao, Z.W.; Huang, M.H.; Jiang, H.Q. Multifunctional reduced graphene oxide film as electrocatalysts and photothermal layer for broad spectrum solar-enhanced oxygen evolution reaction. Mater. Today Energy 2022, 25, 100966. [Google Scholar] [CrossRef]
Figure 1. Publication number of electrocatalytic OER regarding transition-metal-based materials over the last 10 years. Search keywords: OER and transition metal. Website: Web of Science.
Figure 1. Publication number of electrocatalytic OER regarding transition-metal-based materials over the last 10 years. Search keywords: OER and transition metal. Website: Web of Science.
Catalysts 13 01230 g001
Figure 2. (a) Negative values of theoretical overpotential plotted against the standard free energy of ∆G*O − ∆G*OH step. Reproduced from [76] with permission from John Wiley and Sons License. (b) Two-dimensional map of the overpotentials generated by assuming ∆EOOH = EOH + 3.2 for different dopants in FeCoX and NiFeX catalysts: the overpotential can be reduced significantly with the high-valence charge metals. Reproduced from [80] with permission from Springer Nature. (c) Atomic structures of OER intermediates on V25%-Ni2P/NF-AC, where Ni, V, O, and H are represented by green, yellow, red, and grey spheres, respectively. (d) Free energy diagrams of the OER processed on V25%-Ni2P/NF-AC and Ni2P/NF-AC on Ni sites. Reproduced from [81] with permission from John Wiley and Sons License.
Figure 2. (a) Negative values of theoretical overpotential plotted against the standard free energy of ∆G*O − ∆G*OH step. Reproduced from [76] with permission from John Wiley and Sons License. (b) Two-dimensional map of the overpotentials generated by assuming ∆EOOH = EOH + 3.2 for different dopants in FeCoX and NiFeX catalysts: the overpotential can be reduced significantly with the high-valence charge metals. Reproduced from [80] with permission from Springer Nature. (c) Atomic structures of OER intermediates on V25%-Ni2P/NF-AC, where Ni, V, O, and H are represented by green, yellow, red, and grey spheres, respectively. (d) Free energy diagrams of the OER processed on V25%-Ni2P/NF-AC and Ni2P/NF-AC on Ni sites. Reproduced from [81] with permission from John Wiley and Sons License.
Catalysts 13 01230 g002
Figure 3. (a) Scanning transmission electron microscopy (STEM) micrographs of Ir-NiO catalyst, in which the bright spots are ascribed to Ir single atoms. (b) Corresponding atomic modes. (c) Free energy diagrams of OER at a potential of 1.23 V vs. RHE on perfect NiO (001) and single Ir atoms doped NiO (001). The potential-determining steps of OER are indicated for these two surfaces. Insets show the optimized OER intermediates on the Ir-NiO (001) surface. Reproduced from [86] with permission from American Chemical Society.
Figure 3. (a) Scanning transmission electron microscopy (STEM) micrographs of Ir-NiO catalyst, in which the bright spots are ascribed to Ir single atoms. (b) Corresponding atomic modes. (c) Free energy diagrams of OER at a potential of 1.23 V vs. RHE on perfect NiO (001) and single Ir atoms doped NiO (001). The potential-determining steps of OER are indicated for these two surfaces. Insets show the optimized OER intermediates on the Ir-NiO (001) surface. Reproduced from [86] with permission from American Chemical Society.
Catalysts 13 01230 g003
Figure 4. (a) Current densities N-doped NiS2 obtained at different overpotentials. (b) Projected density of states (PDOS) of different N-doped NiS2. (c) Volcano plots for N-doped NiS2 (squares), pure NiS2 (circles), and common metal electrocatalysts (triangles). (d) Free energy diagram of the Ni-0 and Ni-1 sites at the different applied potentials. Reproduced from [89] with permission from Royal Society of Chemistry.
Figure 4. (a) Current densities N-doped NiS2 obtained at different overpotentials. (b) Projected density of states (PDOS) of different N-doped NiS2. (c) Volcano plots for N-doped NiS2 (squares), pure NiS2 (circles), and common metal electrocatalysts (triangles). (d) Free energy diagram of the Ni-0 and Ni-1 sites at the different applied potentials. Reproduced from [89] with permission from Royal Society of Chemistry.
Catalysts 13 01230 g004
Figure 5. (a) Extended X-ray absorption fine structure (EXAFS) fitting curves of k3-weighted, and (b) EXAFS fitting results in k3-space for 3D Co3O4/NC-250 electrocatalyst. (c) Electron paramagnetic resonance (EPR), (d) polarization curves, and (e) structural diagram of N-doped Co3O4-Ov. (f) Free energy diagrams of adsorbate evolution mechanism and lattice-oxygen-mediated mechanism on Ov active site for 3D Co3O4/NC. Reproduced from [96] with permission from Elsevier.
Figure 5. (a) Extended X-ray absorption fine structure (EXAFS) fitting curves of k3-weighted, and (b) EXAFS fitting results in k3-space for 3D Co3O4/NC-250 electrocatalyst. (c) Electron paramagnetic resonance (EPR), (d) polarization curves, and (e) structural diagram of N-doped Co3O4-Ov. (f) Free energy diagrams of adsorbate evolution mechanism and lattice-oxygen-mediated mechanism on Ov active site for 3D Co3O4/NC. Reproduced from [96] with permission from Elsevier.
Catalysts 13 01230 g005
Figure 6. (a) Transmission electron microscopic (TEM) and (b) high-resolution TEM images of NiFeCe-LDH@CP. (c) Proposed four-electron OER mechanism and (d) the Gibbs free energy diagram for the four steps of OER on NiFeCe-LDH@CP. Reproduced from [105] with permission from Elsevier.
Figure 6. (a) Transmission electron microscopic (TEM) and (b) high-resolution TEM images of NiFeCe-LDH@CP. (c) Proposed four-electron OER mechanism and (d) the Gibbs free energy diagram for the four steps of OER on NiFeCe-LDH@CP. Reproduced from [105] with permission from Elsevier.
Catalysts 13 01230 g006
Figure 7. (a) Schematic for the synthesis of W-NiS0.5Se0.5. (b) Scanning electron microscopy image, (c) high-angle annular dark-field scanning TEM (HAADF-STEM) image, and (d) calculated PDOS of W-NiS0.5Se0.5 and NiS0.5Se0.5. (e) The Gibbs free energy diagrams of various oxygen species for W-NiS0.5Se0.5 and NiS0.5Se0.5 during the OER process at 0 V. (f) Proposed OER mechanism on the Ni sites of W-NiS0.5Se0.5. Reproduced from [114] with permission from John Wiley and Sons License.
Figure 7. (a) Schematic for the synthesis of W-NiS0.5Se0.5. (b) Scanning electron microscopy image, (c) high-angle annular dark-field scanning TEM (HAADF-STEM) image, and (d) calculated PDOS of W-NiS0.5Se0.5 and NiS0.5Se0.5. (e) The Gibbs free energy diagrams of various oxygen species for W-NiS0.5Se0.5 and NiS0.5Se0.5 during the OER process at 0 V. (f) Proposed OER mechanism on the Ni sites of W-NiS0.5Se0.5. Reproduced from [114] with permission from John Wiley and Sons License.
Catalysts 13 01230 g007
Figure 8. (a) Geometric area-normalized LSV curves of Ir nanoparticles, Ir/Ni(OH)2, Ni(OH)2, and Ir black. (b) Turnover frequency (TOF) values at 1.53 V and the required overpotential to achieve a TOF of 3.0 s−1 for Ir/Ni(OH)2, Ir nanoparticles, and Ir black. (c) Tafel slopes of Ir nanoparticles, Ir/Ni(OH)2, Ni(OH)2. The inverse fast Fourier transform (FFT) patterns of (d) IrO2, (e) NiOOH, and (f) both IrO2 and NiOOH. (g) Illustration of the proposed OER mechanism. Reproduced from [127] with permission from John Wiley and Sons License.
Figure 8. (a) Geometric area-normalized LSV curves of Ir nanoparticles, Ir/Ni(OH)2, Ni(OH)2, and Ir black. (b) Turnover frequency (TOF) values at 1.53 V and the required overpotential to achieve a TOF of 3.0 s−1 for Ir/Ni(OH)2, Ir nanoparticles, and Ir black. (c) Tafel slopes of Ir nanoparticles, Ir/Ni(OH)2, Ni(OH)2. The inverse fast Fourier transform (FFT) patterns of (d) IrO2, (e) NiOOH, and (f) both IrO2 and NiOOH. (g) Illustration of the proposed OER mechanism. Reproduced from [127] with permission from John Wiley and Sons License.
Catalysts 13 01230 g008
Figure 9. (a) Bar diagram with the maximum magnetocurrent observed for the various magnetic OER catalysts expressed as the relative percentage of the based current, and corresponding applied potential (blue dots). (b) Correlation between the maximum relative magnetocurrent (at 1.67 V vs. RHE) and bulk magnetization. (c) Polarization data for Ni-foam electrodes magnetically decorated with NiZnFe4Ox particles (OFF, filled circles), and under an applied magnetic field (ON, open circles). (d) A pulsed magneto-chronoamperometry experiment was performed at a constant potential of 1.67 V vs. RHE for the NiZn4Ox electrode. Reproduced from [133] with permission from Springer Nature.
Figure 9. (a) Bar diagram with the maximum magnetocurrent observed for the various magnetic OER catalysts expressed as the relative percentage of the based current, and corresponding applied potential (blue dots). (b) Correlation between the maximum relative magnetocurrent (at 1.67 V vs. RHE) and bulk magnetization. (c) Polarization data for Ni-foam electrodes magnetically decorated with NiZnFe4Ox particles (OFF, filled circles), and under an applied magnetic field (ON, open circles). (d) A pulsed magneto-chronoamperometry experiment was performed at a constant potential of 1.67 V vs. RHE for the NiZn4Ox electrode. Reproduced from [133] with permission from Springer Nature.
Catalysts 13 01230 g009
Figure 11. A schematic showing the synergistic modulation for advanced OER electrocatalysts.
Figure 11. A schematic showing the synergistic modulation for advanced OER electrocatalysts.
Catalysts 13 01230 g011
Table 1. List of electrocatalysts and corresponding OER performance in alkaline environment published from 2022 to 2023.
Table 1. List of electrocatalysts and corresponding OER performance in alkaline environment published from 2022 to 2023.
Catalystsη10 (mV)η100 (mV)Tafel Slope (mV dec−1)References
Co(OH)2/NiPx23630452[30]
CoNiFe carbonate hydroxide258/48.7[31]
Vein-like Cr-doping CoxP/32579.2[32]
CoNiFe-LDH nanocages257/31.4[33]
Fe-NiCo-LDH/26251.9[34]
F-CDs/CoP/NF/32896[35]
W-NiS0.5Se0.517123941[36]
CoO@S-CoTe24636256[37]
NiFeV nanofibers/carbon cloth18126947[38]
CoOOH/Co9S8240/86.4[39]
Bi/BiFe(oxy)hydroxide232/34[40]
Ni2P@FePOxHy22026043[41]
NiMo-Fe21726430[42]
Ni@CNTs-MoxC/Ni2P22829743[43]
FexNi2-xP4O12/RGO/27743.8[44]
Co3O4/CoMoO421734272[45]
NiFeW3-LDHs21125636.4[46]
CeO2-NiCoPx/NCF260/72[47]
FeS2@NiS2237/31.4[48]
Ni0.3Fe0.7-LDH@NF18425656.7[49]
Fe-V-doped Ni3S2/NF/25922.4[50]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Z.; Wang, Y.; Lee, L.Y.S. Recent Advances in Synergistic Modulation of Transition-Metal-Based Electrocatalysts for Water Oxidation: A Mini Review. Catalysts 2023, 13, 1230. https://doi.org/10.3390/catal13091230

AMA Style

Li Z, Wang Y, Lee LYS. Recent Advances in Synergistic Modulation of Transition-Metal-Based Electrocatalysts for Water Oxidation: A Mini Review. Catalysts. 2023; 13(9):1230. https://doi.org/10.3390/catal13091230

Chicago/Turabian Style

Li, Zhen, Ying Wang, and Lawrence Yoon Suk Lee. 2023. "Recent Advances in Synergistic Modulation of Transition-Metal-Based Electrocatalysts for Water Oxidation: A Mini Review" Catalysts 13, no. 9: 1230. https://doi.org/10.3390/catal13091230

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop