Next Article in Journal
Correction: Song et al. Terahertz Optical Properties and Carrier Behaviors of Graphene Oxide Quantum Dot and Reduced Graphene Oxide Quantum Dot via Terahertz Time-Domain Spectroscopy. Nanomaterials 2023, 13, 1948
Previous Article in Journal
Quantum Dot-Enabled Biosensing for Prostate Cancer Diagnostics
Previous Article in Special Issue
Evaluation of Aqueous and Ethanolic Extracts for the Green Synthesis of Zinc Oxide Nanoparticles from Tradescantia spathacea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Strategies to Improve the Electrocatalytic Performance of Tungsten Oxide Nanostructures for the Hydrogen Evolution Reaction

School of Physics and Technology, University of Jinan, 336 Nanxinzhuang West Road, Jinan 250022, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(15), 1163; https://doi.org/10.3390/nano15151163
Submission received: 6 June 2025 / Revised: 13 July 2025 / Accepted: 16 July 2025 / Published: 28 July 2025
(This article belongs to the Special Issue Advanced Nanocatalysis in Environmental Applications)

Abstract

Hydrogen, as a renewable and clean energy with a high energy density, is of great significance to the realization of carbon neutrality. In recent years, extensive research has been conducted on the electrocatalytic hydrogen evolution reaction (HER) by splitting water, with a focus on developing efficient electrocatalysts that can perform the HER at an overpotential with minimal power consumption. Tungsten oxide (WO3), a non-noble-metal-based material, has great potential in hydrogen evolution due to its excellent redox capability, low cost, and high stability. However, it cannot meet practical needs because of its poor electrical conductivity and the limited number of active sites; thus, it is necessary to further improve HER performance. In this review, recent advances related to WO3-based electrocatalysts for the HER are introduced. Most importantly, several tactics for optimizing the electrocatalytic HER activity of WO3 are summarized, such as controlling its morphology, phase transition, defect engineering (anion vacancies, cation doping, and interstitial atoms), constructing a heterostructure, and the microenvironment effect. This review can provide insight into the development of novel catalysts with high activity for the HER and other renewable energy applications.

1. Introduction

With the increasing severity of the energy crisis and environmental problems, the world has reached a consensus on the goal of “carbon neutrality” and taken action. Therefore, it is urgent to develop green, safe, and renewable energy as an alternative to traditional fossil fuels [1,2,3]. Hydrogen (H2), which has a high energy density (about 120 kJ/g) and is environmentally friendly [4], will play a remarkable role in the renewable energy system in the future and will be conducive to achieving the goals of carbon dioxide emission reductions and carbon neutrality [5,6]. At present, hydrogen can be produced by coal gasification [7], methane cracking [8], and reforming [9] methods using fossil fuels such as natural gas, coal, and oil as raw materials. Nearly 96% of the world’s hydrogen is produced by the conversion of conventional fossil fuels. These processes are accompanied by the production of large amounts of CO2, sulfur oxides, and other waste gases, and their energy consumption is large, which could aggravate the global energy crisis and environmental pollution. However, the production of “green hydrogen” using renewable energy sources integrated with water electrolysis can achieve zero carbon emissions and is essential for a carbon-neutral strategy [6].
The water electrolysis reaction generally consists of two half-reactions: the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) [10]. The catalyst plays a crucial role in the electrocatalytic reaction, and a highly efficient electrocatalyst can significantly lower the energy consumption required for the reaction. Currently, electrocatalysts depend primarily on precious-metal-based materials, including Pt-based HER electrocatalysts [11,12] and Ir/Ru-based OER electrocatalysts [13,14,15,16,17]. However, the high costs and scarcity of precious-metal-based electrocatalytic materials have hindered their widespread application [18,19]. Therefore, researchers have devoted themselves to the development of sustainable non-precious-metal electrocatalytic materials with low costs and high efficiency and stability. At present, great progress has been made in research on transition metal-based electrocatalysts, including oxides [20,21,22], hydroxide/hydroxyl oxides [23,24,25,26], sulfides [10,25,27,28,29,30], phosphide [31,32,33], selenide [34,35], nitrides [36,37], and carbides [38,39,40], which are much better alternatives to precious-metal catalysts due to their abundant reserves, low prices, environmental friendliness, and great potential for improving catalytic performance.
At present, transition metal oxides (TMOs), including Fe, Co, Ni, W, and Mn oxides, are attracting attention as potential electrocatalytic materials [41,42,43,44,45]. In contrast to the common precious-metal catalysts, transition metal oxides have the advantages of lower costs and easier accessibility, making them more feasible for commercial applications [46]. Tungsten has a great number of stable oxidation states, so it has a variety of electrochemical properties; among these states, tungsten oxide is the most stable. Tungsten oxide is a typical n-type semiconductor with a band gap of 2.6–2.8 eV [47] and electron mobility of about 12 cm2·V−1·s−1 [48]. The crystal structure of WO3 belongs to the ReO3 type, which is similar to the perovskite structure of ABO3 except with the absence of the A-site cation. It can be regarded as having a W atom at the center surrounded by six oxygen atoms, with the basic unit being a regular octahedron formed by WO6. According to the rotation direction and tilt angle of the WO6 octahedron, the crystalline phase of WO3 can be divided into monoclinic II (ε-WO3), triclinic (δ-WO3), monoclinic I (γ-WO3), orthorhombic (β-WO3), tetragonal (α-WO3), and cubic WO3 [49], which are widely used in the fields of energy storage [50], sensors [51,52], and catalysis [53,54,55], among others. WO3, as a transition metal oxide, is a promising electrocatalyst for an efficient HER because of its excellent redox capability, low cost, and high stability [56]. However, its electrical conductivity is poor, and active sites are limited, which result in the poor HER performance of WO3. The conductivity of non-metrological WO3−x, with oxygen vacancies and defect structures, such as WO2 [57,58], WO2.72 [59,60], WO2.83 [61], WO2.9 [20], and WO2.92 [62], is significantly improved. Of course, a lot of effort has been made in regulating and activating WO3−x to improve their HER performance.
Based on previously reported research results, this review first summarizes recent advances in stoichiometric WO3 for the electrocatalytic HER. Then, this review chiefly discusses recent advances in different strategies for improving the catalytic activity of WO3-based materials, such as controlling their morphology, defect engineering, interfacial engineering, tailoring the microenvironment, and so on. Finally, the main conclusions and perspectives related to HER electrocatalysts are discussed. It is hoped that this review can bring some inspiration and reference value to future research on WO3 or other related catalysts.

2. WO3 Nanostructures for HER

The stoichiometric WO3 nanostructures used as electrocatalysts for the HER include nanoparticles, nanorods, nanowires, nanoplates, and so on. WO3 nanostructures have been synthesized using hydrothermal [63,64], sonochemical [65], calcining [66], ionic exchange [67], inorganic–organic hybrid [68], and ultrasonic processes [69], among others. Due to the quantum size effect and their large specific surface areas, the physical and/or chemical properties of nanomaterials deviate from those in the bulk phase. For example, Ganesan et al. prepared WO3 nanoparticles by calcining with a chitosan biopolymer as a template. They exhibited fourfold higher catalytic activity than bulk WO3 for the HER in 1.0 M H2SO4 [66]. In addition, a nanostructure with a one-dimensional (1D) morphology (e.g., nanorods and nanowires) can provide an enlarged electrode–electrolyte interface to facilitate reaction kinetics while simultaneously enabling rapid electron transfer. Consequently, electrocatalysts with 1D architectures can exhibit superior electrocatalytic hydrogen evolution properties. For example, Rajeswari et al. [70] synthesized WO3 nanorods by pyrolysis of tetrabutylammonium decatungstate at 450 °C, and the WO3 nanorods showed much better HER catalytic performance compared with the bulk phase due to the unique electrochemical behavior of 1D nanostructures. Lee et al. [71] prepared hexagonal WO3 nanowires with a large surface area and a large pore volume using a microwave-assisted hydrothermal process. They exhibited excellent HER activity, which was ascribed to their high aspect ratio and crystallinity. Lee et al. [63] synthesized monoclinic WO3 (m-WO3) nanoplates and nanorods using a hydrothermal method, and the m-WO3 nanorods displayed slightly better electrocatalytic activity for the HER than the m-WO3 nanoplates. The experimental results further demonstrated that the 1D nanostructure is not only conducive to increasing the number of active sites for reaction but also enables rapid transfer of electrons along the crystal growth direction. Other studies researching the electrocatalytic hydrogen evolution of pure WO3 have been reported, as summarized in Table 1. It can be seen that the electrocatalytic HER performances of 1D nanostructures (e.g., nanorods and nanowires) and two-dimensional (2D) nanostructures (e.g., nanoplates and nanosheets) are comparable, and both significantly outperform that of nanoparticles. Additionally, their HER activity in acidic electrolytes surpasses that in alkaline electrolytes and is much better than that in neutral solutions.
Many significant research efforts have been devoted to studying WO3 electrocatalysts, achieving notable advancements in catalytic performance, while the intrinsic activity of WO3 nanostructures remains unsatisfactory. Consequently, further designing efficient WO3-based electrocatalysts is imperative for practical application of hydrogen energy.

3. Different Strategies for Improving Electrocatalytic Performance of WO3 for HER

The electrocatalytic HER activity of a material is fundamentally governed by critical parameters, including the charge transport capability, the accessibility of active sites, and electronic structure configurations [76]. Therefore, multifaceted strategies such as controlling the morphology [77], tailoring the phase [74], doping, constructing a heterostructure, and microenvironment modulation [64,78] have been proposed to systematically enhance the intrinsic activity and reaction kinetics of electrocatalysts.

3.1. Morphological Control

Electrocatalytic reactions primarily occur on the surfaces of catalysts, where specific morphologies (e.g., nanosheets, porous structures, and nanowires) are favorable for the exposure of highly active crystal planes or edge sites, thereby increasing the density of active sites. Additionally, nanostructures (such as three-dimensional (3D) porous and hierarchical architectures) can significantly increase the specific surface area to provide additional reaction interfaces, thereby simultaneously promoting electrolyte penetration and gas bubble detachment to reduce mass transport resistance. Unique configurations (e.g., core–shell structures and self-supporting electrodes) effectively suppress catalyst aggregation and dissolution during reactions, thereby prolonging the service life and improving the stability of electrocatalysts. Therefore, construction and preparation of electrocatalytic materials with specific microstructures is an effective strategy to optimize the behavior of catalysts.
The morphologies and crystallographic phases of nanostructures can be controlled by synthetic techniques, conditions, and parameters. Lee et al. [63] prepared monoclinic-phase WO3 using hydrothermal synthesis by adding and tuning the amount of ammonium nitrate (NH4NO3). Furthermore, the morphology of m-WO3 was transformed from nanoplates to nanorods by adding polyethylene glycol in solution, making its HER activity much higher than that of commercial bulk m-WO3. Moreover, the catalytic activity of the m-WO3 nanorods was superior to the nanoplates in an acidic medium, which could be attributed to the increased active sites and improved electron transfer along the crystal growth direction in the unique 1D morphology. Zhang et al. [73] synthesized monoclinic hierarchical flower-like structures consisting of nanoplates and hexagonal nanorods of WO3 by adjusting the concentration of hydrochloric acid (HCL) used in the hydrothermal method. The hexagonal phase of WO3 composed of nanorods required a low overpotential of 55 mV to achieve a current density of 10 mA·cm2 and exhibited excellent electrochemical stability in an acidic medium. It demonstrated much better catalytic activity than the hierarchical monoclinic phase. Ponpandian et al. [74] adjusted the hydrothermal synthesis scheme according to surfactants and physical parameters (temperature and time) and synthesized WO3 electrocatalysts with different morphologies, including 1D-WO3 nanowires (W-NWs), WO3 nanorods (W-NRs), 2D-WO3 nanobelts (W-NBs), WO3 nanoflakes (W-NFs), 3D-WO3 nanoparticles (W-NPs), star-like WO3 (W-S), and WO3 globules (W-Gs), as displayed in Figure 1a. The linear sweep voltammetry (LSV) curves of the as-prepared WO3 electrocatalysts with different morphologies and Pt/C were investigated in acidic and alkaline solutions (Figure 1b,c), respectively. Obviously, the overpotentials of the samples with the same current density in acidic media were lower than that of the alkaline solution. In particular, the W-NRs displayed a much better electrocatalytic HER performance than the other catalysts, and their overpotentials were 152 and 201 mV at 10 mA·cm−2 in 0.5 M H2SO4 and 1 M KOH solutions, respectively. Furthermore, the Tafel slopes of the W-NR electrocatalyst (Figure 1d,e) were 96 mV·dec−1 and 105 mV·dec−1 in 0.5 M H2SO4 and 1 M KOH solutions, respectively. This work has a promising scope for constructing highly efficient interfacial active sites required for large-scale hydrogen production by tuning the morphology and crystalline structure of metal oxides at the molecular level.

3.2. Phase Transition

The intrinsic activity of catalysts fundamentally relies on their crystal structure and surface electronic state [76,79,80], and the latter could be adjusted by phase transitions [81]. Phase transitions can be induced by external stimuli including electric fields [82], pressure [83], temperature [84], and strain [85,86], which could optimize the intrinsic activity and quantity of active sites. Duerloo et al. discovered that application of stress to transition metal dichalcogenides (TMDs) can induce a phase transition from the 2H phase to the 1T phase [86]. Zhang et al. successfully fabricated curved and conformally wrapped cobalt-doped tungsten selenide (Co-WSe2) nanosheets on carbon nanotubes. This curved growth mode enhanced the anisotropy of the material, thereby boosting its intrinsic catalytic activity. Concurrently, the strain generated by the conformal coating further promoted the formation of the metallic 1T phase, significantly improving the electrical conductivity. The strain effects induced by small-diameter tubular multi-walled carbon nanotubes (MWNTs) endowed the catalyst with enhanced catalytic activity and long-term stability [85]. Zhou et al. demonstrated that phase transition from the polymorphic δ phase to the α phase can be induced in MnO2 by adjusting the hydrothermal time. Furthermore, experimental and theoretical density functional theory (DFT) calculation results showed that an α-MnO2 polymorph exhibited much better electrocatalytic activity than δ-MnO2 due to a compatible energy band gap and superior band alignment tunability [80].
Considering the effect of the crystal structure on the electronic structure and adsorption energy of WO3, Ponpandian [74] prepared different crystal phases and morphologies of WO3 using the hydrothermal method. The solvent, surfactant, and synthesis parameters (notably the reaction temperature and time) synergistically determined the structural orientation of the nanostructures. Xie [87] et al. synthesized a series of Pt-WO3 catalysts using the hydrothermal method, and the phase transition from WO3 to HxWO3 caused by an electrochemical H+ insertion/removal process was verified by in situ Raman spectroscopy analysis in Pt-WO3. HxWO3 was found to facilitate electron transfer dynamics and hydrogen transport from HxWO3 to adjacent Pt sites during the HER. Yang et al. [88] have investigated the relationship between intrinsic activity and the crystal structure in the hexagonal and monoclinic phases of WO3 (h-WO3 and m-WO3) during the HER. The WO2 terminations for the h-WO3 (100) and m-WO3 (002) surfaces were selected for investigation, as shown in Figure 2a,b. The computed energy profiles of the reaction coordinates for hydrogen adsorption (Figure 2c) demonstrated that the strong O-H binding interaction suppresses H2 desorption while concurrently diminishing the catalytic activity of bridge O atoms. Electronic structure analysis was employed to elucidate the hydrogen adsorption abilities of the h-WO3 (100) and m-WO3 (002) surfaces, and the calculation results (Figure 2d,e) implied weaker W-H bonds and higher catalytic activity on m-WO3 (002). Furthermore, the d-orbital projected density of state (PDOS) of surface W atoms after H adsorption on the WO3 surface is plotted in Figure 2f. Comparative d-orbital PDOS analysis revealed that the h-WO3 (100) surface exhibits a prominent d-band across the Fermi level, in contrast to the m-WO3 (002) surface. This elevation of the d-band near the Fermi level in the electronic configuration significantly facilitated the formation of W-H bonds. The calculated results conclusively demonstrated that the lower H-adsorption energy and weaker W-H bonding on the m-WO3 (002) surface synergistically accelerated the desorption of surface-adsorbed H* intermediates. According to DFT calculations, the m-WO3 displayed a suitable energy barrier for the H-adsorption/desorption energy, which could be favorable for rapid desorption of active H* intermediates compared with h-WO3, resulting in excellent HER catalytic activity. The h-WO3 and m-WO3 were prepared by the hydrothermal method and calcined with different temperatures under a N2 atmosphere, and the XRD spectra of all samples are shown in Figure 2i,j. The HER was performed in a 0.5 M H2SO4 solution at room temperature, and the experimental results for m-WO3 and h-WO3 are shown in Figure 2k–m. The m-WO3 showed a low Tafel slope of 83 mV·dec−1, and an overpotential of 168 mV was required to realize a current density of 10 mA·cm−2. For the h-WO3, the overpotential was 257 mV at 10 mA·cm−2 and the Tafel slope was 157 mV·dec−1. Obviously, the m-WO3 exhibited much higher electrocatalytic activity than the h-WO3. Consequently, to attain highly active WO3 catalysts, maximizing monoclinic-phase WO3 is recommended in experiments.

3.3. Defect Engineering

Introducing intrinsic defects in transition metal-based electrocatalysts can generate a large number of active sites, increase conductivity, modulate electronic states, promote ion diffusion, and thus enhance catalytic performance. Therefore, studying the regulation of intrinsic defects in non-precious-metal electrocatalysts is of great importance to improve their energy conversion efficiency. The defects include edge defects, vacancy defects [89], and doping-derived defects [90]. Introducing lattice vacancy or atom doping can destroy the periodic arrangement of crystalline structures and can affect the local electronic environment to form unsaturated coordination states. Moreover, the hydrogen adsorption/dissociation energy can also be optimized by defect engineering [91,92].

3.3.1. Anion Vacancies

Anion vacancies are one of the most widespread intrinsic defects. For example, oxygen vacancies widely exist in transition metal oxides due to their low formation energy, which could change the physicochemical properties of oxides [45,93,94,95]. It has been reported that the oxygen vacancies in TMOs are crucial for efficient hydrogen evolution. Especially, introduction of oxygen vacancies usually leads to variation in the oxidation state of a metal, accompanied by the formation of new electron states in the band gap and an increase in the carrier concentration. Wu and their coworkers [93] prepared WO2–carbon mesoporous nanowires (MWCMNs) with a high concentration of oxygen vacancies (VO) by calcinating hybrid WO3 and ethylenediamine precursors, and they exhibited excellent HER activity of 58 at 10 mA·cm−2 and a Tafel slope of 46 mV·dec−1. Experimental and DFT calculation results confirmed that the presence of a large number of VO played a vital role in the introduction of an unusual electron state near the Fermi level and provided more active sites in the MWCMNs. Zhou et al. [89] constructed a hydrogen adsorption model for WO3 and calculated the electronic structure of WO3 with oxygen vacancy using density functional theory. Three different structural models were built on the basis of a WO3 (010) (√2 × √2) R45° slab, as shown in Figure 3a–c. According to the density of state (DOS) results, introducing oxygen vacancies could make new energy levels for W-5d at the conduction band minimum (CBM). Furthermore, the increasing concentration of oxygen vacancies endowed the WO3 slab with a remarkably aggrandized DOS of W 5d at the CBM. The energy gap introduced by VO could transform a metal oxide from a traditional semiconductor to a degenerate semiconductor with high conductivity. The variation in the free energy (ΔGH) decreased to close to zero, which promoted hydrogen adsorption (Figure 3e). In the experiment, WO3 nanosheets with rich VO could be obtained by liquid exfoliation of the tungsten oxide precursors, and they exhibited superior HER activity (Figure 3f,g) with a low overpotential of 38 mV at 10 mA·cm−2 and a Tafel slope of 38 mV·dec−1. It can thus be seen that the experimental results were correspondent with the theoretical calculations.

3.3.2. Cationic Doping

Cationic doping can also regulate the interface structure, surface chemical state, and band gap; optimize the adsorption energies of reaction intermediates; and thus improve the HER of WO3. A summary of studies on metal-doped WO3 catalysts for the HER is presented in Table 2. Liu et al. [96] synthesized self-supported Co-doped WO3 on Cu foam, and the overpotentials were 117, 105, and 149 mV at 10 mA·cm−2 at the pH values of 0, 7, and 14, respectively. The ΔGH of the W sites on a (200) crystal plane of pure WO3 was −0.78 eV, while the ΔGH of the Co sites on a Co-doped WO3 surface was 0.44 eV, which was much closer to 0 than that of the W sites in bare WO3. These results demonstrated that the improved HER activity could be ascribed to the optimization of the adsorption energy of H* species and electric conductivity caused by Co doping. Deng and their coworkers [56] prepared manganese (Mn)- or vanadium (V)-doped WO3 nanoparticles using the hydrothermal method, and the Mn or V impurity atoms replacing the W atom could tune the electronic nature of the WO3. Furthermore, the introduction of Mn and V could reduce the free energy of hydrogen adsorption to −0.11 and 0.08 eV, respectively, at Ob sites. Therefore, the Mn- and V-doped WO3 showed significantly superior HER activity compared to the pure WO3, with overpotentials of 97 mV and 38 mV at 10 mA·cm−2 and Tafel slopes of 68 and 41 mV·dec−1, respectively. Huang et al. [59] regulated the surface electronic structure and morphology of WO2.7−x by doping different metals (Cu, Co, Ni, and Zn), introducing oxygen vacancies, and donating as TM-WO2.7−x, and a schematic illustration of the preparation process of TM-WO2.7−x is shown in Figure 4a. Alkaline HER measurements (Figure 4b–d) confirmed that introducing metal dopants could improve the catalytic performance of WO2.7. The as-prepared Co-WO2.7−x displayed much better HER activity than Ni-WO2.7−x, Cu-WO2.7−x, and Zn-WO2.7−x, with an overpotential of only 59 mV at 10 mA·cm−2 and a Tafel slope of 86 mV·dec−1. According to the calculations for the alkaline HER process with different dopants, the Co-WO2.7−x displayed the lowest energy barrier (0.65 eV) for the rate-determining step, namely the water dissociation step. The experimental data were consistent with the results for the calculated charge density distribution, confirming that the surface properties and intrinsic catalytic activity of the catalysts could be adjusted by doping with metal atoms.

3.3.3. Interstitial Atoms

Interstitial atoms are also an important kind of defect, and a new interstitial atom defect model is proposed to optimize the intrinsic catalytic activity of WO3 via modulation of the surface hydrogen adsorption energy. When small hydrogen atoms enter the interstitial lattice position of WO3, the strong interaction between interstitial hydrogen and lattice oxygen can significantly reduce the electron density of the d orbital in the active W center, thus weakening the hydrogen adsorption energy of WO3 and finally improving the HER performance [102]. Yang and their coworkers annealed ammonium tungstate in the presence of graphite oxide under N2, and hydrogen atoms were successfully inserted into the interstitial sites of tungsten oxide (denoted as H0.23WO3). WO3 can form a variety of crystal structures according to the rotation direction and tilt angle of the WO6 octahedron. m-WO3 presented a perovskite-like structure (Figure 5a) when a hydrogen atom occupied an interstitial position in an m-WO3 twisted lattice that was surrounded by near-regular WO6 octahedrons, and the angle of the W-O-W bonds changed from 165° to 180°. The rotation direction of the WO6 octahedron changed after proton intercalation, which caused the crystalline phase to transform from monoclinic to tetragonal (Figure 5b,c). The crystal phases of as-synthesized catalysts were confirmed by XRD (Figure 5d). The H0.23WO3/rGO electrocatalyst displayed superior activity, with a low overpotential of 33 mV at 10 mA·cm−2 and a Tafel slope of 32.5 mV·dec−1 in acidic media (Figure 5e,f). In addition, the H0.23WO3/rGO catalyst showed amazing stability, and it could work continuously for 200,000 s with no attenuation at a high hydrogen output. The PDOS before (Figure 5h,k) and after (Figure 5l) adsorption of H*, calculated free energy diagrams of the HER (Figure 5i), and the charge density difference (Figure 5j,m) indicated that interstitial H atoms interacted strongly with lattice oxygen in H0.23WO3, reducing the electron density of the d-orbital W sites on the surface. Therefore, a decrement in the electron density at the active sites of H0.23WO3 led to a lower hydrogen adsorption energy and weaker W-H bonds compared with WO3, which promoted rapid desorption of the surface H* intermediates. Therefore, the H0.23WO3 catalyst exhibited enhanced kinetic properties and superior catalytic performance in the hydrogen evolution reaction. It could be seen that the surface electronic structure was adjusted and the hydrogen adsorption energy was weakened by inserting hydrogen atoms into the interstitial positions of WO3, thereby effectively enhancing its catalytic activity.
The above findings confirm that introducing various defects, such as anion vacancies, cation vacancies, and interstitial atoms, can effectively enhance the electrocatalytic hydrogen evolution properties of WO3 nanostructures. However, issues related to the stability of defective WO3 during operation in a hydrogen atmosphere remain. For instance, the V-doped WO3 prepared by Chandrasekaran et al. exhibited approximately 17.3% current degradation after operating at a constant potential of −100 mV for 14,000 s (less than 4 h) [56]. Therefore, improving the stability of electrocatalysts, particularly at high current densities, represents a critical challenge for future research to meet industrial production requirements.

3.4. Constructing Heterostructure

Constructing heterostructures enables regulation of properties such as electrical conductivity, chemical stability, and hydrophilia through integration with different materials. Furthermore, interface engineering significantly optimizes the hydrogen adsorption free energy and charge transfer kinetics by regulating the surface atomic/electronic structures, the distribution of active sites, and the interfacial reaction environments in electrocatalysts [25,26]. Especially, the Mott–Schottky interface, a metal–semiconductor interface, can effectively adjust the electronic structure at the interface to form a built-in electric field, enhancing the electron transfer capacity, creating nucleophilic/electrophilic regions, and optimizing the adsorption free energy of reaction intermediates [103,104]. Some Mott–Schottky interfaces have been constructed for optimizing HER performance, such as Ni/W5N4 [105], Ni/CeO2 [104], Co/CoP [106], and so on. Chen and their coworkers [107] constructed crystalline cobalt nanoparticles on amorphous tungsten oxide (Co/a-WOx). They exhibited excellent HER activity with an overpotential of 36.3 mV at 10 mA·cm−2 and a small Tafel slope of 53.9 mV·dec−1. Peng et al. [108] prepared a Ru-WO2.72 heterostructure (denoted as WR) by adding a minimal number of Ru species to a WO2.72 support. It exhibited a spherical structure characterized by numerous radial nanowires (Figure 6a). Ru species measuring about 2 nm were tightly attached to the surfaces of the WO2.72 nanowires (Figure 6b,c). The WR required an overpotential of about 40 mV to reach a current density of 10 mA·cm−2, and the Tafel slope was 50 mV·dec−1 (Figure 6d,e). A chronoamperometry test of the WR electrocatalyst was performed at 10 mA·cm−2 and remained stable (Figure 6f). DFT calculations were performed to investigate the electronic and transfer properties of the WR heterostructure. The analysis of charge density differences clearly illustrated a pronounced redistribution of charges following the incorporation of Ru (Figure 6j), which could significantly boost the internal electron concentration of the WR composite, thereby improving the adsorption of H* and the HER properties. The DOS distributions of WR and WO2.72 were compared, and the increased electron states at the Fermi level were attributed to the addition of Ru (Figure 6g). A comparison of the partial DOS between WO2.72 and WR suggested that the improved DOS near the Fermi level in WR was primarily caused by Ru d orbitals, revealing that the Ru facilitated d-electron donation near the Fermi level (Figure 6h,i). Furthermore, the calculations indicated that hydrogen adsorption had a ΔGH value for WR (−0.33 eV) closer to zero compared to WO2.72 (−0.78 eV), which gave rise to much better H* adsorption/desorption on WR (Figure 6k). Based on the results of the experimental analysis and DFT calculations, the Mott–Schottky effect and HER mechanism in WR were discussed (Figure 6l). The work function of WO2.72 is lower than that of Ru. Thus, free electrons can spontaneously transfer from WO2.72 to Ru until the Fermi levels align, creating an interfacial electric field. Therefore, the redistribution of charges in Ru and WO2.72 could result in an electron-rich area on the Ru surface, which would enhance its ability to bind with protons. Moreover, WO3 can create a hydrogen–tungsten–bronze compound (HyWO3) in acidic environments by incorporating hydrogen ions, facilitating the transfer of hydrogen from the active site to the WO3 support. The HER process was expedited by the Mott–Schottky structure, which facilitated effective regulation of the electron distribution, and the catalyst underwent a phase transition from WO2.72 to HyWO2.72. This finding confirmed that the Mott–Schottky structure of the catalyst can significantly enhance HER performance by optimizing electronic structures and interfacial charge transfer.
Currently, Pt-based materials are still the most effective electrocatalysts for the HER in water decomposition. However, the scarcity and high cost of Pt severely restrict its practical application. Therefore, it is of great significance to reduce the contents of Pt in catalysts while maintaining their superior catalytic performance. An effective strategy to achieve this goal is to construct a single-atom catalyst in which Pt (or precious-metal) atoms are isolated and dispersed on the carriers, maximizing the exposure of the active site of Pt, significantly promoting its utilization efficiency or mass activity, and thereby considerably minimizing the amount of Pt in the catalyst. Some electronic metal–support interaction catalysts have attracted attention, such as Pt ACs/CoNCs [109], Pt/RuCeOx-PA [110], Pt/N-doped HCS [111], Pt1/OLC [112], and Pt-TiN NTAs [113]. Among them, WO3 is considered an excellent carrier because of its superior stability in acidic and neutral media, as well as its reversible and rapid hydrogen insertion [64,114]. Hou et al. first synthesized [115] monolayer WO3·H2O (ML-WO3·H2O) using a space-confined strategy, then fixed single Pt atoms (Pt-SA) on monolayer WO3 (ML-WO3). The Pt-SA/ML-WO3 maintained the monolayer structure of ML-WO3·H2O, with a monolayer ratio of 93% and abundant defects (O and W vacancies). The experimental results showed that Pt-SA/ML-WO3 had excellent electrocatalytic performance, with a low overpotential of −22 mV at −10 mA·cm−2, a Tafel slope of 27 mV·dec−1, a super-high turnover frequency of about 87 H2 s−1 at −50 mV, and long-term durability. In particular, PT-SA/ML-WO3 displayed ultra-high mass activity of 87 A mgPt−1 at 50 mV, which was about 160 times higher than that of a commercial catalyst of 20 wt% Pt/C (about 0.54 A mgPt−1). According to experimental and density functional theory calculation results, the excellent electrocatalytic HER performance of PT-SA/ML-WO3 was attributed to the strong synergistic effect between the single Pt atom and the carrier. Wang et al. [116] prepared Co-doped WO3-loading Pt nanoparticles (Pt/N-CoWO3) in the presence of ammonia, and a schematic illustration of the preparation process is shown in Figure 7a. As shown by the LSV curves of the as-prepared samples (Figure 7b), the Pt/N-CoWO3 catalyst only requires overpotentials of 94 and 108 mV to reach high current density values of 1 and 2 A·cm−2, respectively, in 0.5 M H2SO4, which are lower than those of commercial Pt/C, Pt/WO3, Pt/CoWO3, and Pt/N-WO3 (Figure 7c). The Pt/N-CoWO3 catalyst exhibited a much lower Tafel slope of 28 mV·dec−1 compared with the other catalysts (Figure 7d). Moreover, the Pt/N-CoWO3 exhibited long-term durability for the HER in acidic conditions. The current density remained almost unchanged in the LSV curves after 4000 cycles (Figure 7e). Especially, it could work continuously for more than 2000 h with little attenuation at 1 A·cm−2 (Figure 7f), and the morphology and structure of Pt/N-CoWO3 had only minor changes after the reaction, showing excellent stability. Two possible H spillover reaction pathways were investigated with Pt as the starting point of the reaction, and free energy diagrams of the HER on different electrocatalysts were generated (Figure 7g,h). For Pt/N-CoWO3, interface phases doped with Co, N, and potential CoO sites were studied (Figure 7i). Theoretical calculations indicated that introducing N was more effective for stabilizing Pt, while introducing Co was more effective for the stabilization of W. Meanwhile, the electronic localization function (ELF) confirmed that the interactions between Pt and the substrates in Pt/N-WO3 and Pt/N-CoWO3 were much stronger than that of Pt/CoWO3, indicating that formation of Pt-N/O bonds was the main reason for the stabilization of Pt on the carrier. In addition, Co doping weakened the absorption of O for H at CoO sites, which promoted hydrogen spillover in the WO3 phase. Of secondary importance, N can promote the migration of H across the interface. In a word, the synergistic effect of Co and N improved the activity and stability of the HER, which provided guidance for the development of high-performance HER catalysts with multi-functional synergies that run for a long time under acidic conditions.
Of course, WO3 can also be combined with other metals or metal compounds (oxides, phosphides, carbides, etc.) to construct highly efficient catalytic systems, and WO3-based heterostructure catalysts for the HER are summarized in Table 3. Wang et al. constructed a self-supported porous Ni17W3/WO3−x/MoO3−x heterostructure with excellent HER performance and good stability [117]. Kong and their coworkers synthesized a WS2/WO3−x heterostructure enriched with oxygen vacancies, which showed superior HER activity with an overpotential of 120 mV at 10 mA·cm−2 and a small Tafel slope of 84.67 mV·dec−1 in 0.5 M H2SO4 [118]. The Ru2P/WO3@NPC (N and P co-doped carbon) heterostructure was achieved by a simple hydrothermal method using ruthenium and tungsten salts as raw materials [119]. The experimental results displayed that the Ru2P/WO3@NPC electrocatalyst exhibited outstanding HER activity, with an overpotential of 15 mV at a current density of 10 mA·cm−2 and a small Tafel slope of 18 mV·dec−1 in a 1.0 M KOH solution, demonstrating superior performance compared with commercial Pt/C and most reported electrocatalysts. Additionally, Ru2P/WO3@NPC showed robust long-term durability with no significant current decay after a continuous chronoamperometry test. These results, combined with DFT calculations, indicated that the electron density redistribution in Ru2P/WO3@NPC was facilitated by electron transfer from NPC to Ru2P/WO3 and from Ru2P to WO3. The NPC served as a carrier for Ru2P/WO3, which prevents nanoparticle aggregation, provides electrons to facilitate water dissociation, and could directly promote dissociation of water at the W site and desorption of hydrogen at the Ru site. In a word, efficient interface engineering can accelerate the development of active catalysts for electrocatalytic processes.

3.5. Microenvironment Effect

In most cases, traditional strategies to improve catalysts by introducing oxygenophilic active elements, heterostructural regulation, nanoscale limiting, etc., can only gradually or gently adjust the adsorption capacities of electronic states and intermediates, thereby improving catalytic performance. However, the instability of these strategies to eliminate the PH-related dynamics of the HER process prevents them from achieving major breakthroughs in non-acidic electrolytes. In acidic media, the HER performance of catalysts is generally superior due to the higher concentration of H3O+, which serves as the primary proton source for the reaction. However, in neutral or alkaline conditions, H2O is the main proton source; thus, the catalysts are required to overcome a higher energy barrier. Especially, hydroxide ions are nearly absent in neutral electrolytes, and the transfer of intermediates diffusing from the electrode to the electrolyte interface is more complex than in alkaline electrolytes [136]. Consequently, the HER kinetics in neutral/alkaline media are significantly slower than in acidic environments. Therefore, selecting a suitable system to maximize the local acid-like environment through multiple physicochemical effects will provide an alternative strategy for improving electrocatalytic performance in non-acidic electrolytes and designing more efficient electrocatalysts [137].
The kinetics of the HER is closely related to the properties of the electrode materials and the local reaction microenvironment near the catalytic site at the electrolyte–solid interface [138,139]. Pang’s research team [78] utilized this method to significantly augment the catalytic activity of defect-rich NiS2/MoS2 nanoflake composites (NMS NFs). The interfacial microenvironment of the NMS NFs could optimize the adsorption energies of reaction intermediates and thereby improve HER performance. Zhao et al. [137] created local acid-like reaction microenvironments on an amorphous NiMoB (am-NiMoB) catalyst in both alkaline and neutral electrolytes, significantly enhancing HER activity. Plenty of hydronium ion (H3O+) intermediates were generated in situ on the catalyst surface during the HER process, which was verified by operando Raman spectroscopy. The am-NiMoB generally achieved excellent performance related to acidic environments, with overpotentials of 38 mV (1.0 M KOH) and 48 mV (1.0 M PBS) at 10 mA·cm−2, and the corresponding Tafel slopes were 34.4 and 39.4 mV·dec−1, respectively. In addition, the am-NiMoB catalyst displayed remarkable durability, maintaining operation for over 350 h during seawater electrolysis at industrial current densities (500 and 1000 mA·cm−2). Wang et al. [64] reported that WO3 achieved reversible and rapid hydrogen doping through electrochemical electron–proton co-doping to form HxWO3 bronzes at reduction potentials. Hydrogen and oxygen atoms combined to form W-OH species with Brønsted acidity, and the charge rearrangement reduced W6+ to W5+, which enhanced the conductivity of the electrocatalyst (Figure 8a,b). By testing the local pH of the material, it was observed that protonated HxWO3 could act as a proton sponge and an electron reservoir to create a more ‘acidic-like’ microenvironment in the electrochemical double layer compared with a traditional carbon substrate (Figure 8c,d). However, a pure HxWO3 support exhibited poor catalytic activity, which was mainly due to the hindered H-H coupling process on its surface (Figure 8e,f). Thus, additional metal sites needed to be introduced. An Ir-HxWO3 electrocatalyst was prepared by electrodeposition, and 1H solid-state nuclear magnetic resonance and temperature-programmed desorption (TPD) (Figure 8g) measurements confirmed that introducing Ir could enhance the fluidity of protons in the HxWO3 carrier. Further analysis of the electronic structures of different hydrogen species showed that the interfacial hydrogen species had more electronic states near the Fermi level, indicating that Ir metal could activate the interfacial hydrogen species, enabling them to participate in the HER process. The operando electrochemical Raman spectra of the neutral HER behavior on the HxWO3 (Figure 8h) and Ir-HxWO3 surfaces (Figure 8i), as well as poisoning (Figure 8j) and kinetic isotope effect (KIE) (Figure 8k) experiments, showed the synergistic catalytic mechanism of Ir particles and HxWO3 support. The Ir site in the Ir-HxWO3 composite had excellent electrocatalytic activity for the Volmer process, which could effectively adsorb H2O and further promote its dissociation, thus generating Ir-Had species at the interface. Ir-Had spontaneously combined with interfacial activated WO-Had to form H2. Meanwhile, the rapid hydrogen transfer on the surface of HxWO3 could quickly replenish the local hydrogen species consumed at the interface, thus achieving a closed loop for the entire catalytic reaction. Therefore, the superior neutral HER performance of the Ir-HxWO3 catalyst arises from coherent synergistic catalysis of the Ir metal site and lattice hydrogen species.

4. Conclusions and Perspectives

In conclusion, this review mainly summarizes the reasonable design of high-performance electrocatalysts based on WO3. Although WO3 nanostructures used as electrocatalysts have achieved some research progress in electrocatalytic hydrogen production, there remains a significant gap before practical application, and their HER performance still needs to be further optimized. The surface morphology and crystal phase of WO3 can be controlled to adjust band gaps and increase the number of active sites. The electronic structure of WO3 can be effectively adjusted by introducing oxygen vacancies and/or atom doping. Stimulating the synergistic effect of multiple phases is used to reduce the reaction barrier of water decomposition and promote adsorption and desorption of protons. In this case, the catalyst has a high electron transport capacity and a large active surface area, which can provide enough active sites and appropriate adsorption strength. It thereby achieves 1 + 1 > 2 in performance. As a result of these efforts, WO3 electrocatalysts have been significantly developed. However, the discussion primarily focuses on the HER activity of WO3-based nanostructures in alkaline and acidic electrolytes, whereas reports on their HER performance in neutral electrolytes and seawater are relatively scarce. Additionally, most reported WO3-based electrocatalysts only exhibit satisfactory catalytic activity at low current densities (≤50 mA cm−2) for durations of tens of hours. There is a conspicuous absence of reports on long-term stability (thousands of hours), particularly regarding sustained performance at industrial-grade current densities. These limitations fall short of the requirements for an industrial alkaline electrolyzer.
So far, many achievements have been made in water electrolysis for sustainable hydrogen production, but there is still a long way to go before it is commercially available. Although many challenges remain, there will be many opportunities for researchers to develop this field in the future.
Firstly, the research on the reaction mechanism is not in-depth. Studying the electrocatalytic mechanism is crucial for both deeply understanding the reaction processes and providing theoretical guidance for the design of new materials. Currently, the electrocatalytic mechanisms of most catalysts, particularly composite catalysts, are still not deeply studied. To address this, it is essential to employ theoretical simulations and in situ or operando characterization techniques to support and guide research into catalytic mechanisms. This approach truly integrates experimental design with theoretical analysis, offering theoretical support for the development of efficient electrocatalytic materials.
Secondly, it is necessary to establish standardized measurements. Standardization of testing methods and a unified evaluation system are essential for accurately assessing the performance of electrocatalysts. Various factors, including catalyst loading, electrode preparation methods, and electrolyte composition, can significantly influence catalytic activity. To facilitate fair comparisons and optimization of electrocatalysts, researchers can supply comprehensive data, including parameters like the Tafel slope, exchange current density, catalyst loading amount, electrode surface area, Faraday efficiency, and stability. Establishing standardized protocols ensures that evaluations of electrocatalytic performance are consistent and reliable, aiding in the selection and optimization of superior existing catalysts and the development of new materials. In addition, if necessary, they could also be notarized by third-party testing organizations.
Thirdly, although several strategies have been reported to prepare efficient HER catalysts based on precious metals, the synthesis processes are complex and expensive, resulting in high costs for the final products. The yields and quality of catalysts are not sufficient to satisfy industrial and commercial demand. Significant progress has been made in the development of highly efficient non-precious transition metal-based HER electrocatalysts, but only a few electrocatalysts have performance comparable to Pt-based catalysts, including their catalytic activity, durability, and stability in a wide pH range. Some research groups use precious-metal doping or disperse Pt-based precious metals onto non-precious-metal carriers, which can reduce the loading of precious metals as much as possible while maintaining the high activity and high stability of precious metals, such as the preparation of atomic-level catalysts or the synthesis of catalysts with a hollow or core–shell structure. The operating costs of the precious-metal catalyst are reduced, and the utilization efficiency of the precious-metal catalyst is improved. Boosting the catalytic performance of existing transition metal-based catalysts and exploring new catalysts will be key research objectives in the years to come.

Author Contributions

Investigation, Y.Q.; formal analysis, Y.Z.; writing—review and editing, Funding acquisition, M.D.; supervision, G.Z.; project administration, W.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shandong Province (grant number ZR2022MB142).

Data Availability Statement

No data was used for the research described in this article.

Acknowledgments

The authors gratefully thank the Natural Science Foundation of Shandong Province.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
HERhydrogen evolution reaction
WO3tungsten oxide
LSVlinear sweep voltammetry
PDOSprojected density of state
DFTdensity functional theory
TMOstransition metal oxides
OERoxygen evolution reaction
1Done-dimensional
3Dthree-dimensional
VOoxygen vacancies
CBMconduction band minimum

References

  1. Liu, Y.; Xiao, C.; Lyu, M.; Lin, Y.; Cai, W.; Huang, P.; Tong, W.; Zou, Y.; Xie, Y. Ultrathin Co3S4 Nanosheets that Synergistically Engineer Spin States and Exposed Polyhedra that Promote Water Oxidation under Neutral Conditions. Angew. Chem. Int. Ed. 2015, 54, 11231–11235. [Google Scholar] [CrossRef] [PubMed]
  2. Huang, H.; Yan, M.; Yang, C.; He, H.; Jiang, Q.; Yang, L.; Lu, Z.; Sun, Z.; Xu, X.; Bando, Y.; et al. Graphene Nanoarchitectonics: Recent Advances in Graphene-Based Electrocatalysts for Hydrogen Evolution Reaction. Adv. Mater. 2019, 31, 1903415. [Google Scholar] [CrossRef]
  3. Wang, C.; Ji, X.; Liang, J.; Zhao, S.; Zhang, X.; Qu, G.; Shao, W.; Li, C.; Zhao, G.; Xu, X.; et al. Activating and Stabilizing a Reversible four Electron Redox Reaction of I−/I+ for Aqueous Zn-Iodine Battery. Angew. Chem. Int. Ed. 2024, 63, e202403187. [Google Scholar] [CrossRef] [PubMed]
  4. Mondal, A.; Vomiero, A. 2D Transition Metal Dichalcogenides-Based Electrocatalysts for Hydrogen Evolution Reaction. Adv. Funct. Mater. 2022, 32, 2208994. [Google Scholar] [CrossRef]
  5. Shi, Z.; Zhang, X.; Lin, X.; Liu, G.; Ling, C.; Xi, S.; Chen, B.; Ge, Y.; Tan, C.; Lai, Z.; et al. Phase-dependent growth of Pt on MoS2 for highly efficient H2 evolution. Nature 2023, 621, 300–305. [Google Scholar] [CrossRef]
  6. Hu, S.; Ge, S.; Liu, H.; Kang, X.; Yu, Q.; Liu, B. Low-Dimensional Electrocatalysts for Acidic Oxygen Evolution: Intrinsic Activity, High Current Density Operation, and Long-Term Stability. Adv. Funct. Mater. 2022, 32, 2201726. [Google Scholar] [CrossRef]
  7. Gnanapragasam, N.V.; Reddy, B.V.; Rosen, M.A. Hydrogen production from coal gasification for effective downstream CO2 capture. Int. J. Hydrog. Energy 2010, 35, 4933–4943. [Google Scholar] [CrossRef]
  8. Abánades, A.; Rubbia, C.; Salmieri, D. Technological challenges for industrial development of hydrogen production based on methane cracking. Energy 2012, 46, 359–363. [Google Scholar] [CrossRef]
  9. Ávila-Neto, C.N.; Dantas, S.C.; Silva, F.A.; Franco, T.V.; Romanielo, L.L.; Hori, C.E.; Assis, A.J. Hydrogen production from methane reforming: Thermodynamic assessment and autothermal reactor design. J. Nat. Gas Sci. Eng. 2009, 1, 205–215. [Google Scholar] [CrossRef]
  10. Yan, Y.; Xia, B.; Xu, Z.; Wang, X. Recent Development of Molybdenum Sulfides as Advanced Electrocatalysts for Hydrogen Evolution Reaction. ACS Catal. 2014, 4, 1693–1705. [Google Scholar] [CrossRef]
  11. Kim, J.; Kim, H.; Lee, W.J.; Ruqia, B.; Baik, H.; Oh, H.S.; Paek, S.M.; Lim, H.K.; Choi, C.H.; Choi, S.I. Theoretical and Experimental Understanding of Hydrogen Evolution Reaction Kinetics in Alkaline Electrolytes with Pt-Based Core–Shell Nanocrystals. J. Am. Chem. Soc. 2019, 141, 18256–18263. [Google Scholar] [CrossRef]
  12. Zhang, W.; Huang, B.; Wang, K.; Yang, W.; Lv, F.; Li, N.; Chao, Y.; Zhou, P.; Yang, Y.; Li, Y.; et al. WOx-Surface Decorated PtNi@Pt Dendritic Nanowires as Efficient pH-Universal Hydrogen Evolution Electrocatalysts. Adv. Energy Mater. 2021, 11, 2003192. [Google Scholar] [CrossRef]
  13. Wang, J.; Yang, H.; Li, F.; Li, L.; Wu, J.; Liu, S.; Cheng, T.; Xu, Y.; Shao, Q.; Huang, X. Single-site Pt-doped RuO2 hollow nanospheres with interstitial C for high-performance acidic overall water splitting. Sci. Adv. 2022, 8, eabl9271. [Google Scholar] [CrossRef]
  14. Meng, G.; Sun, W.; Mon, A.A.; Wu, X.; Xia, L.; Han, A.; Wang, Y.; Zhuang, Z.; Liu, J.; Wang, D.; et al. Strain Regulation to Optimize the Acidic Water Oxidation Performance of Atomic-Layer IrOx. Adv. Mater. 2019, 31, 1903616. [Google Scholar] [CrossRef]
  15. Wang, J.; Ji, Y.; Yin, R.; Li, Y.; Shao, Q.; Huang, X. Transition metal-doped ultrathin RuO2 networked nanowires for efficient overall water splitting across a broad pH range. J. Mater. Chem. A 2019, 7, 6411–6416. [Google Scholar] [CrossRef]
  16. Gao, J.; Xu, C.Q.; Hung, S.F.; Liu, W.; Cai, W.; Zeng, Z.; Jia, C.; Chen, H.M.; Xiao, H.; Li, J.; et al. Breaking Long-Range Order in Iridium Oxide by Alkali Ion for Efficient Water Oxidation. J. Am. Chem. Soc. 2019, 141, 3014–3023. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, J.; Zheng, Y.; Jiao, Y.; Wang, Z.; Lu, Z.; Vasileff, A.; Qiao, S.Z. NiO as a Bifunctional Promoter for RuO2 toward Superior Overall Water Splitting. Small 2018, 14, 1704073. [Google Scholar] [CrossRef] [PubMed]
  18. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef]
  19. Zhou, P.; Zhai, G.; Lv, X.; Liu, Y.; Wang, Z.; Wang, P.; Zheng, Z.; Cheng, H.; Dai, Y.; Huang, B. Boosting the electrocatalytic HER performance of Ni3N-V2O3 via the interface coupling effect. Appl. Catal. B Environ. 2021, 283, 119590. [Google Scholar] [CrossRef]
  20. Li, Y.H.; Liu, P.F.; Pan, L.F.; Wang, H.F.; Yang, Z.Z.; Zheng, L.R.; Hu, P.; Zhao, H.J.; Gu, L.; Yang, H.G. Local atomic structure modulations activate metal oxide as electrocatalyst for hydrogen evolution in acidic water. Nat. Commun. 2015, 6, 8064. [Google Scholar] [CrossRef]
  21. Li, T.; Niu, K.; Yang, M.; Shrestha, N.K.; Gao, Z.; Song, Y.-Y. Ultrathin CoS2 shells anchored on Co3O4 nanoneedles for efficient hydrogen evolution electrocatalysis. J. Power Sources 2017, 356, 89–96. [Google Scholar] [CrossRef]
  22. Li, X.; Wang, Y.; Wang, J.; Da, Y.; Zhang, J.; Li, L.; Zhong, C.; Deng, Y.; Han, X.; Hu, W. Sequential Electrodeposition of Bifunctional Catalytically Active Structures in MoO3/Ni–NiO Composite Electrocatalysts for Selective Hydrogen and Oxygen Evolution. Adv. Mater. 2020, 32, 2003414. [Google Scholar] [CrossRef]
  23. Liu, P.F.; Yang, S.; Zhang, B.; Yang, H.G. Defect-Rich Ultrathin Cobalt–Iron Layered Double Hydroxide for Electrochemical Overall Water Splitting. ACS Appl. Mater. Interfaces 2016, 8, 34474–34481. [Google Scholar] [CrossRef]
  24. Hu, J.; Li, S.; Li, Y.; Wang, J.; Du, Y.; Li, Z.; Han, X.; Sun, J.; Xu, P. A crystalline–amorphous Ni–Ni(OH)2 core–shell catalyst for the alkaline hydrogen evolution reaction. J. Mater. Chem. A 2020, 8, 23323–23329. [Google Scholar] [CrossRef]
  25. Liu, Y.; Ding, M.; Tian, Y.; Zhao, G.; Huang, J.; Xu, X. In-situ growth of 3D hierarchical γ-FeOOH/Ni3S2 heterostructure as high performance electrocatalyst for overall water splitting. J. Colloid Interface Sci. 2023, 639, 24–32. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, Y.; Ding, M.; Qin, Y.; Zhang, B.; Zhang, Y.; Huang, J. Crystalline/Amorphous Mo-Ni(OH)2/FexNiy(OH)3x+2y hierarchical nanotubes as efficient electrocatalyst for overall water splitting. J. Colloid Interface Sci. 2024, 657, 219–228. [Google Scholar] [CrossRef] [PubMed]
  27. Huang, J.; Hou, D.; Zhou, Y.; Zhou, W.; Li, G.; Tang, Z.; Li, L.; Chen, S. MoS2 nanosheet-coated CoS2 nanowire arrays on carbon cloth as three-dimensional electrodes for efficient electrocatalytic hydrogen evolution. J. Mater. Chem. A 2015, 3, 22886–22891. [Google Scholar] [CrossRef]
  28. Wang, D.Y.; Gong, M.; Chou, H.L.; Pan, C.J.; Chen, H.A.; Wu, Y.; Lin, M.C.; Guan, M.; Yang, J.; Chen, C.W.; et al. Highly Active and Stable Hybrid Catalyst of Cobalt-Doped FeS2 Nanosheets–Carbon Nanotubes for Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2015, 137, 1587–1592. [Google Scholar] [CrossRef]
  29. Jiang, N.; Tang, Q.; Sheng, M.; You, B.; Jiang, D.E.; Sun, Y. Nickel sulfides for electrocatalytic hydrogen evolution under alkaline conditions: A case study of crystalline NiS, NiS2, and Ni3S2 nanoparticles. Catal. Sci. Technol. 2016, 6, 1077–1084. [Google Scholar] [CrossRef]
  30. Ling, Y.; Yang, Z.; Zhang, Q.; Zhang, Y.; Cai, W.; Cheng, H. A self-template synthesis of defect-rich WS2 as a highly efficient electrocatalyst for the hydrogen evolution reaction. Chem. Commun. 2018, 54, 2631–2634. [Google Scholar] [CrossRef]
  31. Li, W.; Xiong, D.; Gao, X.; Song, W.G.; Xia, F.; Liu, L. Self-supported Co-Ni-P ternary nanowire electrodes for highly efficient and stable electrocatalytic hydrogen evolution in acidic solution. Catal. Today 2017, 287, 122–129. [Google Scholar] [CrossRef]
  32. Zhang, P.; Chen, H.; Wang, M.; Yang, Y.; Jiang, J.; Zhang, B.; Duan, L.; Daniel, Q.; Li, F.; Sun, L. Gas-templating of hierarchically structured Ni–Co–P for efficient electrocatalytic hydrogen evolution. J. Mater. Chem. A 2017, 5, 7564–7570. [Google Scholar] [CrossRef]
  33. Liu, H.; Ma, X.; Hu, H.; Pan, Y.; Zhao, W.; Liu, J.; Zhao, X.; Wang, J.; Yang, Z.; Zhao, Q.; et al. Robust NiCoP/CoP Heterostructures for Highly Efficient Hydrogen Evolution Electrocatalysis in Alkaline Solution. ACS Appl. Mater. Interfaces 2019, 11, 15528–15536. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, K.; Li, Y.; Deng, S.; Shen, S.; Zhang, Y.; Pan, G.; Xiong, Q.; Liu, Q.; Xia, X.; Wang, X.; et al. Molybdenum Selenide Electrocatalysts for Electrochemical Hydrogen Evolution Reaction. ChemElectroChem 2019, 6, 3530–3548. [Google Scholar] [CrossRef]
  35. Sun, J.; Li, J.; Li, Z.; Hu, X.; Bai, H.; Meng, X. Phase Transition in Cobalt Selenide with a Greatly Improved Electrocatalytic Activity in Hydrogen Evolution Reactions. ACS Sustain. Chem. Eng. 2022, 10, 4022–4030. [Google Scholar] [CrossRef]
  36. Gao, D.; Zhang, J.; Wang, T.; Xiao, W.; Tao, K.; Xue, D.; Ding, J. Metallic Ni3N nanosheets with exposed active surface sites for efficient hydrogen evolution. J. Mater. Chem. A 2016, 4, 17363–17369. [Google Scholar] [CrossRef]
  37. Liu, B.; He, B.; Peng, H.Q.; Zhao, Y.; Cheng, J.; Xia, J.; Shen, J.; Ng, T.W.; Meng, X.; Lee, C.S.; et al. Unconventional Nickel Nitride Enriched with Nitrogen Vacancies as a High-Efficiency Electrocatalyst for Hydrogen Evolution. Adv. Sci. 2018, 5, 1800406. [Google Scholar] [CrossRef]
  38. Kumar, R.; Gaur, A.; Maruyama, T.; Bera, C.; Bagchi, V. Strong Interactions between the Nanointerfaces of Silica-Supported Mo2C/MoP Heterojunction Promote Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 57898–57906. [Google Scholar] [CrossRef]
  39. Yang, Y.; Qian, Y.; Luo, Z.; Li, H.; Chen, L.; Cao, X.; Wei, S.; Zhou, B.; Zhang, Z.; Chen, S.; et al. Water induced ultrathin Mo2C nanosheets with high-density grain boundaries for enhanced hydrogen evolution. Nat. Commun. 2022, 13, 7225. [Google Scholar] [CrossRef]
  40. Chen, W.; Niu, M.; Zhang, Z.; Chen, L.; Li, X.; Zhang, J.; Sun, R.; Cao, H.; Wang, X. Phase-Transition of Mo2C Induced by Tungsten Doping as Heterointerface-Rich Electrocatalyst for Optimizing Hydrogen Evolution Activity. Small 2024, 20, 2311026. [Google Scholar] [CrossRef]
  41. Masa, J.; Xia, W.; Sinev, I.; Zhao, A.; Sun, Z.; Grützke, S.; Weide, P.; Muhler, M.; Schuhmann, W. MnxOy/NC and CoxOy/NC Nanoparticles Embedded in a Nitrogen-Doped Carbon Matrix for High-Performance Bifunctional Oxygen Electrodes. Angew. Chem. Int. Ed. 2014, 53, 8508–8512. [Google Scholar] [CrossRef] [PubMed]
  42. Meng, Y.; Song, W.; Huang, H.; Ren, Z.; Chen, S.; Suib, S.L. Structure–Property Relationship of Bifunctional MnO2 Nanostructures: Highly Efficient, Ultra-Stable Electrochemical Water Oxidation and Oxygen Reduction Reaction Catalysts Identified in Alkaline Media. J. Am. Chem. Soc. 2014, 136, 11452–11464. [Google Scholar] [CrossRef]
  43. Prabu, M.; Ketpang, K.; Shanmugam, S. Hierarchical nanostructured NiCo2O4 as an efficient bifunctional non-precious metal catalyst for rechargeable zinc–air batteries. Nanoscale 2014, 6, 3173–3181. [Google Scholar] [CrossRef]
  44. Gao, Y.; Zhao, H.; Chen, D.; Chen, C.; Ciucci, F. In situ synthesis of mesoporous manganese oxide/sulfur-doped graphitized carbon as a bifunctional catalyst for oxygen evolution/reduction reactions. Carbon 2015, 94, 1028–1036. [Google Scholar] [CrossRef]
  45. Chen, J.; Yu, D.; Liao, W.; Zheng, M.; Xiao, L.; Zhu, H.; Zhang, M.; Du, M.; Yao, J. WO3–x Nanoplates Grown on Carbon Nanofibers for an Efficient Electrocatalytic Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2016, 8, 18132–18139. [Google Scholar] [CrossRef]
  46. Osgood, H.; Devaguptapu, S.V.; Xu, H.; Cho, J.; Wu, G. Transition metal (Fe, Co, Ni, and Mn) oxides for oxygen reduction and evolution bifunctional catalysts in alkaline media. Nano Today 2016, 11, 601–625. [Google Scholar] [CrossRef]
  47. Zhou, Y.; Hu, X.C.; Liu, X.H.; Wen, H.R. Core–shell hierarchical WO2/WO3 microspheres as an electrocatalyst support for methanol electrooxidation. Chem. Commun. 2015, 51, 15297–15299. [Google Scholar] [CrossRef]
  48. Li, W.; Da, P.; Zhang, Y.; Wang, Y.; Lin, X.; Gong, X.; Zheng, G. WO3 Nanoflakes for Enhanced Photoelectrochemical Conversion. ACS Nano 2014, 8, 11770–11777. [Google Scholar] [CrossRef]
  49. Zheng, H.; Ou, J.Z.; Strano, M.S.; Kaner, R.B.; Mitchell, A.; Kalantar zadeh, K. Nanostructured Tungsten Oxide—Properties, Synthesis, and Applications. Adv. Funct. Mater. 2011, 21, 2175–2196. [Google Scholar] [CrossRef]
  50. Pathak, R.; Gurung, A.; Elbohy, H.; Chen, K.; Reza, K.M.; Bahrami, B.; Mabrouk, S.; Ghimire, R.; Hummel, M.; Gu, Z.; et al. Self-recovery in Li-metal hybrid lithium-ion batteries via WO3 reduction. Nanoscale 2018, 10, 15956–15966. [Google Scholar] [CrossRef] [PubMed]
  51. Epifani, M.; Comini, E.; Díaz, R.; Andreu, T.; Genç, A.; Arbiol, J.; Siciliano, P.; Faglia, G.; Morante, J.R. Solvothermal, Chloroalkoxide-based Synthesis of Monoclinic WO3 Quantum Dots and Gas-Sensing Enhancement by Surface Oxygen Vacancies. ACS Appl. Mater. Interfaces 2014, 6, 16808–16816. [Google Scholar] [CrossRef]
  52. Hu, X.; Li, X.; Yang, H.; Liu, S.; Qin, Z.; Xie, C.; Zeng, D. Active edge effect of pothole-rich WO3 nanosheets for enhancing dimethyl trisulfide gas sensing performance. Sens. Actuators B Chem. 2022, 352, 131103. [Google Scholar] [CrossRef]
  53. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  54. Yu, W.; Chen, J.; Shang, T.; Chen, L.; Gu, L.; Peng, T. Direct Z-scheme g-C3N4/WO3 photocatalyst with atomically defined junction for H2 production. Appl. Catal. B Environ. 2017, 219, 693–704. [Google Scholar] [CrossRef]
  55. Pehlivan, İ.B.; Atak, G.; Niklasson, G.A.; Stolt, L.; Edoff, M.; Edvinsson, T. Electrochromic solar water splitting using a cathodic WO3 electrocatalyst. Nano Energy 2021, 81, 105620. [Google Scholar] [CrossRef]
  56. Chandrasekaran, S.; Zhang, P.; Peng, F.; Bowen, C.; Huo, J.; Deng, L. Tailoring the geometric and electronic structure of tungsten oxide with manganese or vanadium doping toward highly efficient electrochemical and photoelectrochemical water splitting. J. Mater. Chem. A 2019, 7, 6161–6172. [Google Scholar] [CrossRef]
  57. Meng, G.; Yao, H.; Tian, H.; Kong, F.; Cui, X.; Cao, S.; Chen, Y.; Chang, Z.; Chen, C.; Shi, J. An electrochemically reconstructed WC/WO2–WO3 heterostructure as a highly efficient hydrogen oxidation electrocatalyst. J. Mater. Chem. A 2022, 10, 622–631. [Google Scholar] [CrossRef]
  58. Chen, Z.; Gong, W.; Wang, J.; Hou, S.; Yang, G.; Zhu, C.; Fan, X.; Li, Y.; Gao, R.; Cui, Y. Metallic W/WO2 solid-acid catalyst boosts hydrogen evolution reaction in alkaline electrolyte. Nat. Commun. 2023, 14, 5363. [Google Scholar] [CrossRef]
  59. Huang, H.; Xu, L.; Woo, D.Y.; Kim, S.; Kim, S.M.; Kim, Y.K.; Byeon, J.; Lee, J. Refining the surface properties of WO2.7 by vacancy engineering and transition metals doping for enhanced alkaline hydrogen evolution reaction. Chem. Eng. J. 2023, 451, 138939. [Google Scholar] [CrossRef]
  60. Chen, H.; Song, L.; Ouyang, S.; Wang, J.; Lv, J.; Ye, J. Co and Fe Codoped WO2.72 as Alkaline-Solution-Available Oxygen Evolution Reaction Catalyst to Construct Photovoltaic Water Splitting System with Solar-To-Hydrogen Efficiency of 16.9%. Adv. Sci. 2019, 6, 1900465. [Google Scholar] [CrossRef]
  61. Zhou, Y.; Li, R.; Dong, L.; Yin, S.; Chu, B.; Chen, Z.; Wang, J.; Li, B.; Fan, M. Heterointerface and Defect Dual Engineering in a Superhydrophilic Ni2P/WO2.83 Microsphere for Boosting Alkaline Hydrogen Evolution Reaction at High Current Density. ACS Appl. Mater. Interfaces 2022, 14, 18816–18824. [Google Scholar] [CrossRef] [PubMed]
  62. Peng, Q.; He, Q.; Hu, Y.; Isimjan, T.T.; Hou, R.; Yang, X. Interface engineering of porous Fe2P-WO2.92 catalyst with oxygen vacancies for highly active and stable large-current oxygen evolution and overall water splitting. J. Energy Chem. 2022, 65, 574–582. [Google Scholar] [CrossRef]
  63. Ham, D.J.; Phuruangrat, A.; Thongtem, S.; Lee, J.S. Hydrothermal synthesis of monoclinic WO3 nanoplates and nanorods used as an electrocatalyst for hydrogen evolution reactions from water. Chem. Eng. J. 2010, 165, 365–369. [Google Scholar] [CrossRef]
  64. Zheng, X.; Shi, X.; Ning, H.; Yang, R.; Lu, B.; Luo, Q.; Mao, S.; Xi, L.; Wang, Y. Tailoring a local acid-like microenvironment for efficient neutral hydrogen evolution. Nat. Commun. 2023, 14, 4209. [Google Scholar] [CrossRef]
  65. Koltypin, Y.; Nikitenko, S.I.; Gedanken, A. The sonochemical preparation of tungsten oxide nanoparticles. J. Mater. Chem. 2002, 12, 1107–1110. [Google Scholar] [CrossRef]
  66. Ganesan, R.; Gedanken, A. Synthesis of WO3 nanoparticles using a biopolymer as a template for electrocatalytic hydrogen evolution. Nanotechnology 2008, 19, 025702. [Google Scholar] [CrossRef]
  67. Yan, Z.; Wei, W.; Xie, J.; Meng, S.; Lü, X.; Zhu, J. An ion exchange route to produce WO3 nanobars as Pt electrocatalyst promoter for oxygen reduction reaction. J. Power Sources 2013, 222, 218–224. [Google Scholar] [CrossRef]
  68. Chen, D.; Gao, L.; Yasumori, A.; Kuroda, K.; Sugahara, Y. Size- and Shape-Controlled Conversion of Tungstate-Based Inorganic–Organic Hybrid Belts to WO3 Nanoplates with High Specific Surface Areas. Small 2008, 4, 1813–1822. [Google Scholar] [CrossRef]
  69. Mohamed, M.J.S.; Gondal, M.A.; Hassan, M.; Almessiere, M.A.; Tahir, A.A.; Roy, A. Effective Hydrogen Production from Alkaline and Natural Seawater using WO3–x@CdS1–x Nanocomposite-Based Electrocatalysts. ACS Omega 2023, 8, 33332–33341. [Google Scholar] [CrossRef]
  70. Rajeswari, J.; Kishore, P.; Viswanathan, B.; Varadarajan, T. Facile Hydrogen Evolution Reaction on WO3 Nanorods. Nanoscale Res. Lett. 2007, 2, 496. [Google Scholar] [CrossRef]
  71. Phuruangrat, A.; Ham, D.J.; Hong, S.J.; Thongtem, S.; Lee, J.S. Synthesis of hexagonal WO3 nanowires by microwave-assisted hydrothermal method and their electrocatalytic activities for hydrogen evolution reaction. J. Mater. Chem. 2010, 20, 1683–1690. [Google Scholar] [CrossRef]
  72. Zheng, H.; Mathe, M. Hydrogen evolution reaction on single crystal WO3/C nanoparticles supported on carbon in acid and alkaline solution. Int. J. Hydrog. Energy 2011, 36, 1960–1964. [Google Scholar] [CrossRef]
  73. Zhang, X.; Jin, G.; Wang, D.; Chen, Z.; Zhao, M.; Xi, G. Crystallographic phase and morphology dependent hydrothermal synthesis of tungsten oxide for robust hydrogen evolution reaction. J. Alloys Compd. 2021, 875, 160054. [Google Scholar] [CrossRef]
  74. Rajalakshmi, R.; Rebekah, A.; Viswanathan, C.; Ponpandian, N. Evolution of intrinsic 1-3D WO3 nanostructures: Tailoring their phase structure and morphology for robust hydrogen evolution reaction. Chem. Eng. J. 2022, 428, 132013. [Google Scholar] [CrossRef]
  75. Nayak, A.K.; Verma, M.; Sohn, Y.; Deshpande, P.A.; Pradhan, D. Highly Active Tungsten Oxide Nanoplate Electrocatalysts for the Hydrogen Evolution Reaction in Acidic and Near Neutral Electrolytes. ACS Omega 2017, 2, 7039–7047. [Google Scholar] [CrossRef] [PubMed]
  76. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. Advancing the Electrochemistry of the Hydrogen-Evolution Reaction through Combining Experiment and Theory. Angew. Chem. Int. Ed. 2015, 54, 52–65. [Google Scholar] [CrossRef]
  77. Mohamed, M.M.; Salama, T.M.; Hegazy, M.A.; Abou Shahba, R.M.; Mohamed, S.H. Synthesis of hexagonal WO3 nanocrystals with various morphologies and their enhanced electrocatalytic activities toward hydrogen evolution. Int. J. Hydrog. Energy 2019, 44, 4724–4736. [Google Scholar] [CrossRef]
  78. Qian, Y.; Yu, J.; Zhang, Y.; Zhang, F.; Kang, Y.; Su, C.; Shi, H.; Kang, D.J.; Pang, H. Interfacial Microenvironment Modulation Enhancing Catalytic Kinetics of Binary Metal Sulfides Heterostructures for Advanced Water Splitting Electrocatalysts. Small Methods 2022, 6, 2101186. [Google Scholar] [CrossRef]
  79. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef]
  80. Zhou, Y.; Zhou, Z.; Hu, L.; Tian, R.; Wang, Y.; Arandiyan, H.; Chen, F.; Li, M.; Wan, T.; Han, Z.; et al. A facile approach to tailor electrocatalytic properties of MnO2 through tuning phase transition, surface morphology and band structure. Chem. Eng. J. 2022, 438, 135561. [Google Scholar] [CrossRef]
  81. Greeley, J.; Jaramillo, T.F.; Bonde, J.; Chorkendorff, I.; Nørskov, J.K. Computational high-throughput screening of electrocatalytic materials for hydrogen evolution. Nat. Mater. 2006, 5, 909–913. [Google Scholar] [CrossRef]
  82. Saleem, M.S.; Cui, B.; Song, C.; Sun, Y.; Gu, Y.; Zhang, R.; Fayaz, M.U.; Zhou, X.; Werner, P.; Parkin, S.S.P.; et al. Electric Field Control of Phase Transition and Tunable Resistive Switching in SrFeO2.5. ACS Appl. Mater. Interfaces 2019, 11, 6581–6588. [Google Scholar] [CrossRef] [PubMed]
  83. Nayak, A.P.; Bhattacharyya, S.; Zhu, J.; Liu, J.; Wu, X.; Pandey, T.; Jin, C.; Singh, A.K.; Akinwande, D.; Lin, J.F. Pressure-induced semiconducting to metallic transition in multilayered molybdenum disulphide. Nat. Commun. 2014, 5, 3731. [Google Scholar] [CrossRef] [PubMed]
  84. Blunt, M.O.; Adisoejoso, J.; Tahara, K.; Katayama, K.; Van der Auweraer, M.; Tobe, Y.; De Feyter, S. Temperature-Induced Structural Phase Transitions in a Two-Dimensional Self-Assembled Network. J. Am. Chem. Soc. 2013, 135, 12068–12075. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, G.; Zheng, X.; Xu, Q.; Zhang, J.; Liu, W.; Chen, J. Carbon nanotube-induced phase and stability engineering: A strained cobalt-doped WSe2/MWNT heterostructure for enhanced hydrogen evolution reaction. J. Mater. Chem. A 2018, 6, 4793–4800. [Google Scholar] [CrossRef]
  86. Duerloo, K.-A.N.; Li, Y.; Reed, E.J. Structural phase transitions in two-dimensional Mo- and W-dichalcogenide monolayers. Nat. Commun. 2014, 5, 4214. [Google Scholar] [CrossRef]
  87. Xie, C.; Chen, W.; Du, S.; Yan, D.; Zhang, Y.; Chen, J.; Liu, B.; Wang, S. In-situ phase transition of WO3 boosting electron and hydrogen transfer for enhancing hydrogen evolution on Pt. Nano Energy 2020, 71, 104653. [Google Scholar] [CrossRef]
  88. Yang, J.; Chen, X.; Liu, X.; Cao, Y.; Huang, J.; Li, Y.; Liu, F. From Hexagonal to Monoclinic: Engineering Crystalline Phase to Boost the Intrinsic Catalytic Activity of Tungsten Oxides for the Hydrogen Evolution Reaction. ACS Sustain. Chem. Eng. 2021, 9, 5642–5650. [Google Scholar] [CrossRef]
  89. Zheng, T.; Sang, W.; He, Z.; Wei, Q.; Chen, B.; Li, H.; Cao, C.; Huang, R.; Yan, X.; Pan, B.; et al. Conductive Tungsten Oxide Nanosheets for Highly Efficient Hydrogen Evolution. Nano Lett. 2017, 17, 7968–7973. [Google Scholar] [CrossRef]
  90. Tang, T.; Wang, Z.; Guan, J. A review of defect engineering in two-dimensional materials for electrocatalytic hydrogen evolution reaction. Chin. J. Catal. 2022, 43, 636–678. [Google Scholar] [CrossRef]
  91. Huang, Z.H.; Li, H.; Li, W.H.; Henkelman, G.; Jia, B.; Ma, T. Electrical and Structural Dual Function of Oxygen Vacancies for Promoting Electrochemical Capacitance in Tungsten Oxide. Small 2020, 16, 2004709. [Google Scholar] [CrossRef]
  92. Zhang, C.; Shi, Y.; Yu, Y.; Du, Y.; Zhang, B. Engineering Sulfur Defects, Atomic Thickness, and Porous Structures into Cobalt Sulfide Nanosheets for Efficient Electrocatalytic Alkaline Hydrogen Evolution. ACS Catal. 2018, 8, 8077–8083. [Google Scholar] [CrossRef]
  93. Wu, R.; Zhang, J.; Shi, Y.; Liu, D.; Zhang, B. Metallic WO2–Carbon Mesoporous Nanowires as Highly Efficient Electrocatalysts for Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2015, 137, 6983–6986. [Google Scholar] [CrossRef]
  94. Diao, J.; Yuan, W.; Qiu, Y.; Cheng, L.; Guo, X. A hierarchical oxygen vacancy-rich WO3 with “nanowire-array-on-nanosheet-array” structure for highly efficient oxygen evolution reaction. J. Mater. Chem. A 2019, 7, 6730–6739. [Google Scholar] [CrossRef]
  95. Cheng, H.; Klapproth, M.; Sagaltchik, A.; Li, S.; Thomas, A. Ordered mesoporous WO2.83: Selective reduction synthesis, exceptional localized surface plasmon resonance and enhanced hydrogen evolution reaction activity. J. Mater. Chem. A 2018, 6, 2249–2256. [Google Scholar] [CrossRef]
  96. Liu, L.; Zhao, Y.; Wang, Y.; Zhang, Z.; Liu, J.; Wu, T.; Qin, W.; Liu, S.; Jia, B.; Wu, H.; et al. Single-Atom Co Doped in Ultrathin WO3 Arrays for the Enhanced Hydrogen Evolution Reaction in a Wide pH Range. ACS Appl. Mater. Interfaces 2021, 13, 53915–53924. [Google Scholar] [CrossRef] [PubMed]
  97. Zhong, X.; Sun, Y.; Chen, X.; Zhuang, G.; Li, X.; Wang, J.G. Mo Doping Induced More Active Sites in Urchin-Like W18O49 Nanostructure with Remarkably Enhanced Performance for Hydrogen Evolution Reaction. Adv. Funct. Mater. 2016, 26, 5778–5786. [Google Scholar] [CrossRef]
  98. Zhao, Y.; Tang, Q.; Yang, P.; He, B. Robust electrocatalysts from metal doped W18O49 nanofibers for hydrogen evolution. Chem. Commun. 2017, 53, 4323–4326. [Google Scholar] [CrossRef]
  99. Hai, G.; Huang, J.; Cao, L.; Kajiyoshi, K.; Wang, L.; Feng, L.; Chen, J. Activation of urchin-like Ni-doped W18O49/NF by electrochemical tuning for efficient water splitting. J. Energy Chem. 2021, 63, 642–650. [Google Scholar] [CrossRef]
  100. Zhao, Y.; Tang, Q.; He, B.; Yang, P. Mo incorporated W18O49 nanofibers as robust electrocatalysts for high-efficiency hydrogen evolution. Int. J. Hydrog. Energy 2017, 42, 14534–14546. [Google Scholar] [CrossRef]
  101. Liang, W.; Zhou, M.; Lin, X.; Xu, J.; Dong, P.; Le, Z.; Yang, M.; Chen, J.; Xie, F.; Wang, N.; et al. Nickel-doped tungsten oxide promotes stable and efficient hydrogen evolution in seawater. Appl. Catal. B Environ. 2023, 325, 122397. [Google Scholar] [CrossRef]
  102. Yang, J.; Cao, Y.; Zhang, S.; Shi, Q.; Chen, S.; Zhu, S.; Li, Y.; Huang, J. Interstitial Hydrogen Atom to Boost Intrinsic Catalytic Activity of Tungsten Oxide for Hydrogen Evolution Reaction. Small 2023, 19, 2207295. [Google Scholar] [CrossRef] [PubMed]
  103. Gu, C.; Zhou, G.; Yang, J.; Pang, H.; Zhang, M.; Zhao, Q.; Gu, X.; Tian, S.; Zhang, J.; Xu, L.; et al. NiS/MoS2 Mott-Schottky heterojunction-induced local charge redistribution for high-efficiency urea-assisted energy-saving hydrogen production. Chem. Eng. J. 2022, 443, 136321. [Google Scholar] [CrossRef]
  104. Li, T.; Yin, J.; Sun, D.; Zhang, M.; Pang, H.; Xu, L.; Zhang, Y.; Yang, J.; Tang, Y.; Xue, J. Manipulation of Mott−Schottky Ni/CeO2 Heterojunctions into N-Doped Carbon Nanofibers for High-Efficiency Electrochemical Water Splitting. Small 2022, 18, 2106592. [Google Scholar] [CrossRef] [PubMed]
  105. Zhou, Y.; Chu, B.; Sun, Z.; Dong, L.; Wang, F.; Li, B.; Fan, M.; Chen, Z. Surface reconstruction and charge distribution enabling Ni/W5N4 Mott-Schottky heterojunction bifunctional electrocatalyst for efficient urea-assisted water electrolysis at a large current density. Appl. Catal. B Environ. 2023, 323, 122168. [Google Scholar] [CrossRef]
  106. Xue, Z.H.; Su, H.; Yu, Q.Y.; Zhang, B.; Wang, H.H.; Li, X.H.; Chen, J.S. Janus Co/CoP Nanoparticles as Efficient Mott–Schottky Electrocatalysts for Overall Water Splitting in Wide pH Range. Adv. Energy Mater. 2017, 7, 1602355. [Google Scholar] [CrossRef]
  107. Chen, J.; Zheng, J.; He, W.; Liang, H.; Li, Y.; Cui, H.; Wang, C. Self-standing hollow porous Co/a-WOx nanowire with maximum Mott-Schottky effect for boosting alkaline hydrogen evolution reaction. Nano Res. 2023, 16, 4603–4611. [Google Scholar] [CrossRef]
  108. Peng, L.; Su, L.; Yu, X.; Wang, R.; Cui, X.; Tian, H.; Cao, S.; Xia, B.Y.; Shi, J. Electron redistribution of ruthenium-tungsten oxides Mott-Schottky heterojunction for enhanced hydrogen evolution. Appl. Catal. B Environ. 2022, 308, 121229. [Google Scholar] [CrossRef]
  109. Zhao, Y.; Kumar, P.V.; Tan, X.; Lu, X.; Zhu, X.; Jiang, J.; Pan, J.; Xi, S.; Yang, H.Y.; Ma, Z.; et al. Modulating Pt-O-Pt atomic clusters with isolated cobalt atoms for enhanced hydrogen evolution catalysis. Nat. Commun. 2022, 13, 2430. [Google Scholar] [CrossRef]
  110. Liu, T.; Gao, W.; Wang, Q.; Dou, M.; Zhang, Z.; Wang, F. Selective Loading of Atomic Platinum on a RuCeOx Support Enables Stable Hydrogen Evolution at High Current Densities. Angew. Chem. Int. Ed. 2020, 59, 20423–20427. [Google Scholar] [CrossRef]
  111. Shang, X.; Liu, Z.Z.; Lu, S.S.; Dong, B.; Chi, J.Q.; Qin, J.F.; Liu, X.; Chai, Y.M.; Liu, C.G. Pt–C Interfaces Based on Electronegativity-Functionalized Hollow Carbon Spheres for Highly Efficient Hydrogen Evolution. ACS Appl. Mater. Interfaces 2018, 10, 43561–43569. [Google Scholar] [CrossRef]
  112. Liu, D.; Li, X.; Chen, S.; Yan, H.; Wang, C.; Wu, C.; Haleem, Y.A.; Duan, S.; Lu, J.; Ge, B.; et al. Atomically dispersed platinum supported on curved carbon supports for efficient electrocatalytic hydrogen evolution. Nat. Energy 2019, 4, 512–518. [Google Scholar] [CrossRef]
  113. Zhao, J.; Zeng, Y.; Wang, J.; Xu, Q.; Chen, R.; Ni, H.; Cheng, G.J. Ultrahigh electrocatalytic activity with trace amounts of platinum loadings on free-standing mesoporous titanium nitride nanotube arrays for hydrogen evolution reactions. Nanoscale 2020, 12, 15393–15401. [Google Scholar] [CrossRef]
  114. Li, Y.F.; Lu, W.; Chen, K.; Duchesne, P.; Ali, F.M.; Xia, M.; Wood, T.E.; Ulmer, U.; Ozin, G.A. Cu Atoms on Nanowire Pd/HyWO3−x Bronzes Enhance the Solar Reverse Water Gas Shift Reaction. J. Am. Chem. Soc. 2019, 141, 14991–14996. [Google Scholar] [CrossRef]
  115. Wang, D.; Li, H.; Du, N.; Hou, W. Single Platinum Atoms Immobilized on Monolayer Tungsten Trioxide Nanosheets as an Efficient Electrocatalyst for Hydrogen Evolution Reaction. Adv. Funct. Mater. 2021, 31, 2009770. [Google Scholar] [CrossRef]
  116. Chen, H.; Yu, J.; Liu, L.; Gao, R.-T.; Gao, Z.; Yang, Y.; Chen, Z.; Zhan, S.; Liu, X.; Zhang, X.; et al. Modulating Pt-N/O Bonds on Co-doped WO3 for Acid Electrocatalytic Hydrogen Evolution with Over 2000 h Operation. Adv. Energy Mater. 2024, 14, 2303635. [Google Scholar] [CrossRef]
  117. Sun, Y.; Bao, Y.; Yin, D.; Bu, X.; Zhang, Y.; Yue, K.; Qi, X.; Cai, Z.; Li, Y.; Hu, X.; et al. Oxygen vacancy-induced efficient hydrogen spillover in Ni17W3/WO3−x/MoO3−x for a superior pH-universal hydrogen evolution reaction. J. Mater. Chem. A 2024, 12, 11563–11570. [Google Scholar] [CrossRef]
  118. Kong, L.; Pan, L.; Guo, H.; Qiu, Y.; Alshahrani, W.A.; Amin, M.A.; Lin, J. Constructing WS2/WO3−x heterostructured electrocatalyst enriched with oxygen vacancies for accelerated hydrogen evolution reaction. J. Colloid Interface Sci. 2024, 664, 178–185. [Google Scholar] [CrossRef]
  119. Jiang, X.; Jang, H.; Liu, S.; Li, Z.; Kim, M.G.; Li, C.; Qin, Q.; Liu, X.; Cho, J. The Heterostructure of Ru2P/WO3/NPC Synergistically Promotes H2O Dissociation for Improved Hydrogen Evolution. Angew. Chem. Int. Ed. 2021, 60, 4110–4116. [Google Scholar] [CrossRef]
  120. Wang, F.; Dong, B.; Wang, J.; Ke, N.; Tan, C.; Huang, A.; Wu, Y.; Hao, L.; Yin, L.; Xu, X.; et al. Self-supported porous heterostructure WC/WO3−x ceramic electrode for hydrogen evolution reaction in acidic and alkaline media. J. Adv. Ceram. 2022, 11, 1208–1221. [Google Scholar] [CrossRef]
  121. Lv, Y.; Liu, Y.; Chen, C.; Wang, T.; Zhang, M. Octopus tentacles-like WO3/C@CoO as high property and long life-time electrocatalyst for hydrogen evolution reaction. Electrochim. Acta 2018, 281, 1–8. [Google Scholar] [CrossRef]
  122. Tian, H.; Cui, X.; Zeng, L.; Su, L.; Song, Y.; Shi, J. Oxygen vacancy-assisted hydrogen evolution reaction of the Pt/WO3 electrocatalyst. J. Mater. Chem. A 2019, 7, 6285–6293. [Google Scholar] [CrossRef]
  123. Sun, T.; Zhong, W.; Zhang, M.; Tang, Y.; Wang, J.; Xu, C.; Hu, L.; Wang, M. A low-cost 2D WO3/Ni3S2 heterojunction for highly stable hydrogen evolution. Mater. Chem. Front. 2021, 5, 8248–8254. [Google Scholar] [CrossRef]
  124. Wang, B.; Wang, L.; Qian, Y.; Yang, Y.; Isimjan, T.T.; Yang, X. Construction of a self-supporting Ni2P–WO3 heterostructure for highly efficient hydrogen evolution under both caustic and acidic conditions. Sustain. Energy Fuels 2021, 5, 2884–2892. [Google Scholar] [CrossRef]
  125. Chen, J.; Chen, C.; Qin, M.; Li, B.; Lin, B.; Mao, Q.; Yang, H.; Liu, B.; Wang, Y. Reversible hydrogen spillover in Ru-WO3−x enhances hydrogen evolution activity in neutral pH water splitting. Nat. Commun. 2022, 13, 5382. [Google Scholar] [CrossRef]
  126. Liu, L.; Wang, Y.; Zhao, Y.; Wang, Y.; Zhang, Z.; Wu, T.; Qin, W.; Liu, S.; Jia, B.; Wu, H.; et al. Ultrahigh Pt-Mass-Activity Hydrogen Evolution Catalyst Electrodeposited from Bulk Pt. Adv. Funct. Mater. 2022, 32, 2112207. [Google Scholar] [CrossRef]
  127. Nguyen, N.A.; Chuluunbat, E.; Nguyen, T.A.; Choi, H.S. High electrocatalytic activity of Rh-WO3 electrocatalyst for hydrogen evolution reaction under the acidic, alkaline, and alkaline-seawater electrolytes. Int. J. Hydrog. Energy 2023, 48, 32686–32698. [Google Scholar] [CrossRef]
  128. Paudel, D.R.; Pan, U.N.; Ghising, R.B.; Kandel, M.R.; Prabhakaran, S.; Kim, D.H.; Kim, N.H.; Lee, J.H. Multi-interfacial dendritic engineering facilitating congruous intrinsic activity of oxide-carbide/MOF nanostructured multimodal electrocatalyst for hydrogen and oxygen electrocatalysis. Appl. Catal. B Environ. 2023, 331, 122711. [Google Scholar] [CrossRef]
  129. Zhang, K.; Guo, F.; Graham, N.; Yu, W. Engineering Morphology and Electron Redistribution of a Ni/WO3 Mott–Schottky Bifunctional Electrocatalyst for Efficient Alkaline Urea Splitting. ACS Appl. Mater. Interfaces 2023, 15, 50116–50125. [Google Scholar] [CrossRef]
  130. Fan, X.; Liu, C.; Gao, B.; Li, H.; Zhang, Y.; Zhang, H.; Gao, Q.; Cao, X.; Tang, Y. Electronic Structure Engineering of Pt Species over Pt/WO3 toward Highly Efficient Electrocatalytic Hydrogen Evolution. Small 2023, 19, 2301178. [Google Scholar] [CrossRef]
  131. Jing, S.; Lu, J.; Yu, G.; Yin, S.; Luo, L.; Zhang, Z.; Ma, Y.; Chen, W.; Shen, P.K. Carbon-Encapsulated WOx Hybrids as Efficient Catalysts for Hydrogen Evolution. Adv. Mater. 2018, 30, 1705979. [Google Scholar] [CrossRef]
  132. Pi, Y.; Qiu, Z.; Sun, Y.; Ishii, H.; Liao, Y.F.; Zhang, X.; Chen, H.Y.; Pang, H. Synergistic Mechanism of Sub-Nanometric Ru Clusters Anchored on Tungsten Oxide Nanowires for High-Efficient Bifunctional Hydrogen Electrocatalysis. Adv. Sci. 2023, 10, 2206096. [Google Scholar] [CrossRef]
  133. Peng, Y.W.; Shan, C.; Wang, H.J.; Hong, L.Y.; Yao, S.; Wu, R.J.; Zhang, Z.M.; Lu, T.B. Polyoxometalate-Derived Ultrasmall Pt2W/WO3 Heterostructure Outperforms Platinum for Large-Current-Density H2 Evolution. Adv. Energy Mater. 2019, 9, 1900597. [Google Scholar] [CrossRef]
  134. Kadam, S.R.; Bar Ziv, R.; Bar Sadan, M. A cobalt-doped WS2/WO3 nanocomposite electrocatalyst for the hydrogen evolution reaction in acidic and alkaline media. New J. Chem. 2022, 46, 20102–20107. [Google Scholar] [CrossRef]
  135. Dogra, N.; Kushvaha, S.S.; Dewasi, A.; Sharma, S. Unveiling the origin of high catalytic activity of WO3/MWCNT nanocomposites for the hydrogen evolution reaction. Catal. Sci. Technol. 2025, 15, 2327–2338. [Google Scholar] [CrossRef]
  136. Yang, W.; Li, M.; Zhang, B.; Liu, Y.; Zi, J.; Xiao, H.; Liu, X.; Lin, J.; Zhang, H.; Chen, J.; et al. Interfacial Microenvironment Modulation Boosts Efficient Hydrogen Evolution Reaction in Neutral and Alkaline. Adv. Funct. Mater. 2023, 33, 2304852. [Google Scholar] [CrossRef]
  137. Zhao, R.; Zhang, C.; Wei, L.; Wei, D.; Su, J.; Guo, L. Tailoring a local acidic microenvironment on amorphous NiMoB catalyst to boost alkaline and neutral hydrogen evolution reactions. Appl. Catal. B Environ. Energy 2025, 365, 124928. [Google Scholar] [CrossRef]
  138. Tan, H.; Tang, B.; Lu, Y.; Ji, Q.; Lv, L.; Duan, H.; Li, N.; Wang, Y.; Feng, S.; Li, Z.; et al. Engineering a local acid-like environment in alkaline medium for efficient hydrogen evolution reaction. Nat. Commun. 2022, 13, 2024. [Google Scholar] [CrossRef]
  139. Guo, J.; Zheng, Y.; Hu, Z.; Zheng, C.; Mao, J.; Du, K.; Jaroniec, M.; Qiao, S.-Z.; Ling, T. Direct seawater electrolysis by adjusting the local reaction environment of a catalyst. Nat. Energy 2023, 8, 264–272. [Google Scholar] [CrossRef]
Figure 1. (a) Synthesis protocol for WO3 electrocatalysts with different morphologies. LSV curves of WO3 electrocatalysts in (b) 0.5 M H2SO4 and (c) 1 M KOH. Tafel plots of WO3 electrocatalysts in (d) 0.5 M H2SO4 and (e) 1 M KOH [74].
Figure 1. (a) Synthesis protocol for WO3 electrocatalysts with different morphologies. LSV curves of WO3 electrocatalysts in (b) 0.5 M H2SO4 and (c) 1 M KOH. Tafel plots of WO3 electrocatalysts in (d) 0.5 M H2SO4 and (e) 1 M KOH [74].
Nanomaterials 15 01163 g001
Figure 2. The optimized geometric configurations of (a) h-WO3 (100) and (b) m-WO3 (002) surfaces with W atoms (gray spheres) and O atoms (red spheres). (c) Calculated HER free energy profiles under equilibrium for W and O active sites on h-WO3 (100) and m-WO3 (002) surfaces. Three-dimensional contour plots of charge density differences for H adsorption at W sites on (d) h-WO3 (100) and (e) m-WO3 (002) surfaces. The yellow and light blue areas imply electron accumulation and depletion, respectively. (f) PDOS analysis of the d orbital for W atoms on h-WO3 (100) and m-WO3 (002) after hydrogen adsorption. Crystalline structure diagrams of the (g) hexagonal and (h) monoclinic phases. XRD of (i) h-WO3 and (j) m-WO3. (k) LSV curves. (l) The overpotentials required at 10 mA·cm−2 for different catalysts. (m) Tafel slopes of as-grown samples. All electrochemical measurements were performed in a 0.5 M H2SO4 solution at room temperature [88].
Figure 2. The optimized geometric configurations of (a) h-WO3 (100) and (b) m-WO3 (002) surfaces with W atoms (gray spheres) and O atoms (red spheres). (c) Calculated HER free energy profiles under equilibrium for W and O active sites on h-WO3 (100) and m-WO3 (002) surfaces. Three-dimensional contour plots of charge density differences for H adsorption at W sites on (d) h-WO3 (100) and (e) m-WO3 (002) surfaces. The yellow and light blue areas imply electron accumulation and depletion, respectively. (f) PDOS analysis of the d orbital for W atoms on h-WO3 (100) and m-WO3 (002) after hydrogen adsorption. Crystalline structure diagrams of the (g) hexagonal and (h) monoclinic phases. XRD of (i) h-WO3 and (j) m-WO3. (k) LSV curves. (l) The overpotentials required at 10 mA·cm−2 for different catalysts. (m) Tafel slopes of as-grown samples. All electrochemical measurements were performed in a 0.5 M H2SO4 solution at room temperature [88].
Nanomaterials 15 01163 g002
Figure 3. The crystal structures of WO3 (010) with (a) an ideal surface (labeled as W24O72), (b) a surface with one bridging VO (labeled as W24O71), and (c) a surface with all terminal O atoms and one bridging O atom removed (labeled as W24O67). (d) The DOS of W 5d. (e) The ΔGH at the W site on WO3 (010) with low and high concentrations of VO. (f) LSV curves. (g) Tafel plots of different samples measured in H2-saturated 0.5 M H2SO4 [89].
Figure 3. The crystal structures of WO3 (010) with (a) an ideal surface (labeled as W24O72), (b) a surface with one bridging VO (labeled as W24O71), and (c) a surface with all terminal O atoms and one bridging O atom removed (labeled as W24O67). (d) The DOS of W 5d. (e) The ΔGH at the W site on WO3 (010) with low and high concentrations of VO. (f) LSV curves. (g) Tafel plots of different samples measured in H2-saturated 0.5 M H2SO4 [89].
Nanomaterials 15 01163 g003
Figure 4. (a) A schematic illustration of TM-WO2.7−x preparation. (b) LSV curves of as-prepared samples. (c) A comparison of the involved overpotentials of different catalysts at 10 mA·cm−2. (d) The Tafel slopes of all samples measured in a 1 M KOH solution. (e) The Gibbs free energy changes of the alkaline HER process on different samples [59].
Figure 4. (a) A schematic illustration of TM-WO2.7−x preparation. (b) LSV curves of as-prepared samples. (c) A comparison of the involved overpotentials of different catalysts at 10 mA·cm−2. (d) The Tafel slopes of all samples measured in a 1 M KOH solution. (e) The Gibbs free energy changes of the alkaline HER process on different samples [59].
Nanomaterials 15 01163 g004
Figure 5. Crystalline structure diagrams of (a) monoclinic WO3 and (b) tetragonal H0.23WO3. (c) The transformation of the crystal structure from monoclinic WO3 to tetragonal H0.23WO3 caused by a hydrogen atom occupying an interstitial site. (d) XRD spectra. (e) The LSV curves of the samples. (f) Tafel slopes. (g) The LSV curves of the initial H0.23WO3/rGO composite and the composite after 5000 cycles. The inset shows a test of H0.23WO3/rGO stability performed for 200,000 s at 100 mV. All electrocatalytic HER performance tests were performed in 0.5 M H2SO4. (h,k) The DOS results for W 5d orbitals and O 2p orbitals on WO3 (002) and H0.23WO3 (020) surfaces. (i) Calculated free energy diagrams for the surface W and O sites on WO3 (002) and H0.23WO3 (020). (l) PDOS results for the 5d orbitals of surface W atoms on WO3 (002) and H0.23WO3 (020) surfaces after H* adsorption. (j,m) The variation in charge density at the active sites following hydrogen adsorption on the surface of the catalyst [102].
Figure 5. Crystalline structure diagrams of (a) monoclinic WO3 and (b) tetragonal H0.23WO3. (c) The transformation of the crystal structure from monoclinic WO3 to tetragonal H0.23WO3 caused by a hydrogen atom occupying an interstitial site. (d) XRD spectra. (e) The LSV curves of the samples. (f) Tafel slopes. (g) The LSV curves of the initial H0.23WO3/rGO composite and the composite after 5000 cycles. The inset shows a test of H0.23WO3/rGO stability performed for 200,000 s at 100 mV. All electrocatalytic HER performance tests were performed in 0.5 M H2SO4. (h,k) The DOS results for W 5d orbitals and O 2p orbitals on WO3 (002) and H0.23WO3 (020) surfaces. (i) Calculated free energy diagrams for the surface W and O sites on WO3 (002) and H0.23WO3 (020). (l) PDOS results for the 5d orbitals of surface W atoms on WO3 (002) and H0.23WO3 (020) surfaces after H* adsorption. (j,m) The variation in charge density at the active sites following hydrogen adsorption on the surface of the catalyst [102].
Nanomaterials 15 01163 g005
Figure 6. (a) TEM, (b) HRTEM, (c) high-angle annular dark-field scanning transmission electron microscopy, and elemental mapping images of WR. (d) LSV curves. (e) The Tafel slopes of different samples. (f) The LSV curves of the initial WR catalyst and the catalyst after 2000 cycles. The inset shows the chronoamperometry curve of WR at an overpotential of 50 mV. (g) The calculated DOS results for WO2.72 and WR. (h) The PDOS results for WO2.72. (i) The PDOS results for WR. (j) The calculated charge density differences of all samples. (k) A free energy diagram of the (010) planes in WO2.72 and WR. (l) A schematic diagram of the HER mechanism in WR [108].
Figure 6. (a) TEM, (b) HRTEM, (c) high-angle annular dark-field scanning transmission electron microscopy, and elemental mapping images of WR. (d) LSV curves. (e) The Tafel slopes of different samples. (f) The LSV curves of the initial WR catalyst and the catalyst after 2000 cycles. The inset shows the chronoamperometry curve of WR at an overpotential of 50 mV. (g) The calculated DOS results for WO2.72 and WR. (h) The PDOS results for WO2.72. (i) The PDOS results for WR. (j) The calculated charge density differences of all samples. (k) A free energy diagram of the (010) planes in WO2.72 and WR. (l) A schematic diagram of the HER mechanism in WR [108].
Nanomaterials 15 01163 g006
Figure 7. (a) A schematic illustration of the preparation process of Pt/N-CoWO3. (b) The LSV curves of the as-synthesized samples. (c) A comparison of the overpotentials at 0.5, 1, and 2 A·cm−2. (d) Tafel plots. (e) The LSV curves of the initial Pt/N-CoWO3 catalyst and the catalyst after 4000 cycles. (f) The tested stability curves. (g,h) The DFT calculations of free energy for the HER for different samples. (i) The optimized H* adsorption structures at different sites on Pt/N-CoWO3 [116].
Figure 7. (a) A schematic illustration of the preparation process of Pt/N-CoWO3. (b) The LSV curves of the as-synthesized samples. (c) A comparison of the overpotentials at 0.5, 1, and 2 A·cm−2. (d) Tafel plots. (e) The LSV curves of the initial Pt/N-CoWO3 catalyst and the catalyst after 4000 cycles. (f) The tested stability curves. (g,h) The DFT calculations of free energy for the HER for different samples. (i) The optimized H* adsorption structures at different sites on Pt/N-CoWO3 [116].
Nanomaterials 15 01163 g007
Figure 8. (a). Schematic diagram of hydrogen intercalation process of WO3 to form HxWO3. (b) EIS Nyquist plots. (c) Schematic illustration of local ‘acidic-like’ microenvironment on HxWO3 electrocatalyst. (d) PH values on HxWO3 surfaces with different potentials. (e) LSV of HxWO3. (f) Tafel plots. (g) TPD-MS thermal desorption profiles for Ir-HxWO3 and HxWO3. Operando electrochemical Raman spectra of (h) HxWO3 and (i) Ir-HxWO3. (j) LSV curves of Ir-HxWO3 before and after poisoning using Li+ and SCN. (k) LSV curves of Ir-HxWO3 in 1.0 M PBS (H2O and D2O). (l) Free energy diagram of HER process on Ir10-HxWO3 [64].
Figure 8. (a). Schematic diagram of hydrogen intercalation process of WO3 to form HxWO3. (b) EIS Nyquist plots. (c) Schematic illustration of local ‘acidic-like’ microenvironment on HxWO3 electrocatalyst. (d) PH values on HxWO3 surfaces with different potentials. (e) LSV of HxWO3. (f) Tafel plots. (g) TPD-MS thermal desorption profiles for Ir-HxWO3 and HxWO3. Operando electrochemical Raman spectra of (h) HxWO3 and (i) Ir-HxWO3. (j) LSV curves of Ir-HxWO3 before and after poisoning using Li+ and SCN. (k) LSV curves of Ir-HxWO3 in 1.0 M PBS (H2O and D2O). (l) Free energy diagram of HER process on Ir10-HxWO3 [64].
Nanomaterials 15 01163 g008
Table 1. A summary of pure WO3 catalysts with different morphologies for the HER.
Table 1. A summary of pure WO3 catalysts with different morphologies for the HER.
CatalystsElectrolyteOverpotential (mV)Current Density (mA·cm−2)Tafel Slope (mV·dec−1)References
WO3 nanoparticles1 M H2SO480045 [66]
WO3 nanoplates1 M H2SO420017.58122[63]
m-WO3 nanorods1 M H2SO420023.86113[63]
WO3 nanoparticles0.5 M H2SO42000.7229[72]
WO3 nanoparticles1 M KOH8000.11114[72]
WO3 nanorods1 M H2SO480023188[70]
bulk WO31 M H2SO480015213[70]
WO3 nanoplates0.5 M H2SO41061078[73]
WO3 nanorods0.5 M H2SO4831048[73]
WO3 nanowires1 M H2SO410038.4116[71]
WO3 nanorods0.5 M H2SO41521096[74]
WO3 nanorods1 M KOH20110105[74]
WO3 nanoplates0.5 M H2SO4731039.5[75]
WO3 nanoplatesdistilled water331151.59[75]
Table 2. A summary of metal-doped WO3−x catalysts for the HER.
Table 2. A summary of metal-doped WO3−x catalysts for the HER.
CatalystsElectrolyteOverpotential (mV)Current Density (mA·cm−2)Tafel Slope (mV·dec−1)References
V-WO30.5 M H2SO4381041[56]
Mn-WO30.5 M H2SO4971068[56]
Mo-W18O490.5 M H2-saturated H2SO4451054[97]
Pd-W18O490.5 M H2SO41371035[98]
Ni-W18O491 M KOH240/35010/10092[99]
1 at% Mo-W18O490.5 M H2SO4262/46210/5049[100]
Co-WO2.7−x1 M KOH591086[59]
Ni-WO2.7−x1 M KOH9510129[59]
Zn--WO2.7−x1 M KOH5301072[59]
Ni-WOx1 M KOH40.51/137.0410/10040[101]
Ni-WOx1 M KOH seawater45.69/125.8110/10046[101]
Table 3. A summary of catalysts with WO3-x-based heterostructures for the HER.
Table 3. A summary of catalysts with WO3-x-based heterostructures for the HER.
CatalystsElectrolyteOverpotential (mV)Current Density (mA·cm−2)Tafel Slope (mV·dec−1)References
W/WO21.0 M KOH351034[58]
Ni2P/WO2.831.0 M KOH22.8/254.510/100053[61]
WO3−x@CdS1−x1.0 M KOH1911061.9[69]
Co/a-WOx1.0 M KOH36.3/55.110/2053.9[107]
Ru-WO2.720.5 M H2SO4401050[108]
Pt-SA/ML-WO3N2-saturated 0.5 M H2SO4221027[115]
Pt/N-CoWO3N2-saturated 0.5 M H2SO483/94/108500/1000/200028[116]
Ni17W3/WO3−x/MoO3−x1.0 M KOH161034.9[117]
Ni17W3/WO3−x/MoO3−x0.5 M H2SO4141032.6[117]
Ni17W3/WO3−x/MoO3−x1.0 M PBS421073.9[117]
WS2/WO3−x0.5 M H2SO41201084.67[118]
WS2/WO3−x1.0 M KOH1511097.29[118]
Ru2P/WO3@NPC1.0 M KOH151018[119]
WC/WO3−x0.5 M H2SO41071059.3[120]
WC/WO3−x1.0 M KOH1231072.4[120]
Ar/H2-treated WO3/C@CoO/NF1.0 M KOH5510115[121]
Pt/def-WO3@CFC0.5 M H2SO4421061[122]
WO3/Ni3S21.0 M KOH24910045.06[123]
Ni2P-WO3/CC1.0 M KOH1051064.2[124]
Ni2P-WO3/CC0.5 M H2SO41071055.9[124]
Ru-WO3−x1.0 M PBS191041[125]
PtCu/WO3@CF0.5 M H2SO4411045.90[126]
Rh-WO30.5 M H2SO4481031[127]
Rh-WO31.0 M KOH1161073[127]
Rh-WO30.5 M NaCl/1.0 M KOH981084[127]
FeCu–BTC/WO3–WC1.0 M KOH99/220/28610/50/10073.2[128]
Ni/WO31.0 M NaOH163100 [129]
Pt/WO3-6000.5 M H2SO48/2610/10035[130]
WOx@C/C0.5 M H2SO4366019.17[131]
Ru SNC/W18O49 NWs0.5 M H2SO4211035[132]
Pt2W/WO3/RGO0.5 M H2SO4394500 [133]
Co-WS2/WO30.5 M H2SO432110108[134]
Co-WS2/WO30.5 M KOH33710136[134]
WO3/MWCNT1.0 M KOH2001070[135]
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

Ding, M.; Qin, Y.; Ji, W.; Zhang, Y.; Zhao, G. A Review of Strategies to Improve the Electrocatalytic Performance of Tungsten Oxide Nanostructures for the Hydrogen Evolution Reaction. Nanomaterials 2025, 15, 1163. https://doi.org/10.3390/nano15151163

AMA Style

Ding M, Qin Y, Ji W, Zhang Y, Zhao G. A Review of Strategies to Improve the Electrocatalytic Performance of Tungsten Oxide Nanostructures for the Hydrogen Evolution Reaction. Nanomaterials. 2025; 15(15):1163. https://doi.org/10.3390/nano15151163

Chicago/Turabian Style

Ding, Meng, Yuan Qin, Weixiao Ji, Yafang Zhang, and Gang Zhao. 2025. "A Review of Strategies to Improve the Electrocatalytic Performance of Tungsten Oxide Nanostructures for the Hydrogen Evolution Reaction" Nanomaterials 15, no. 15: 1163. https://doi.org/10.3390/nano15151163

APA Style

Ding, M., Qin, Y., Ji, W., Zhang, Y., & Zhao, G. (2025). A Review of Strategies to Improve the Electrocatalytic Performance of Tungsten Oxide Nanostructures for the Hydrogen Evolution Reaction. Nanomaterials, 15(15), 1163. https://doi.org/10.3390/nano15151163

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