Next Article in Journal
Innovative Approaches to Large-Area Perovskite Solar Cell Fabrication Using Slit Coating
Previous Article in Journal
Enhanced Thermoelectric Performance of SnTe via Introducing Resonant Levels
Previous Article in Special Issue
Strain Engineering of Unconventional Crystal-Phase Noble Metal Nanocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Stoichiometric Nickel Sulfide-Based Catalysts for Hydrogen Evolution Reaction in Alkaline Media

Department of Chemical Engineering, Department of Integrative Engineering for Hydrogen Safety, Kangwon National University, Chuncheon 24341, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(20), 4975; https://doi.org/10.3390/molecules29204975
Submission received: 23 September 2024 / Revised: 18 October 2024 / Accepted: 18 October 2024 / Published: 21 October 2024
(This article belongs to the Special Issue Metal-Based Nanomaterials in Catalysis and Electrochemistry)

Abstract

:
Efficient and cost-effective catalysts for hydrogen evolution reaction (HER) are essential for large-scale hydrogen production, which is a critical step toward reducing carbon emissions and advancing the global transition to sustainable energy. Nickel sulfide-based catalysts, which exist in various stoichiometries, show promise for HER in alkaline media. However, as single-phase materials, they do not demonstrate superior activity compared to Pt-based catalysts. This review highlights recent strategies to enhance the HER performance of nickel sulfides, including heteroatom doping, heterostructure construction, and vacancy engineering, tailored to their different stoichiometric ratios. The study also examines synthesis methods, characterizations, and their impact on HER performance. Furthermore, it discusses the challenges and limitations of current research and suggests future directions for improvement.

1. Introduction

As climate and energy crises intensify, various agendas have been proposed to achieve net-zero emissions by 2050 [1,2]. In response, a transition to renewable energy, coupled with a reduction in fossil fuel consumption, is being proposed as a fundamental solution. To facilitate this energy transition, numerous policies and research initiatives have focused on renewable energy sources, such as solar, wind, hydro, geothermal, and hydrogen energy. Among these, hydrogen energy is gaining attention as a clean and sustainable renewable energy source due to its high gravimetric energy density, versatility, and zero emissions when used. These characteristics not only make it a valuable energy source but also enable it to serve as an energy carrier for storing excess power generated from other renewable sources.
Despite these advantages, the environmental impact of hydrogen varies significantly depending on its production method. Hydrogen is classified by color based on production methods and is broadly categorized as gray, blue, and green hydrogen. Gray hydrogen is produced from fossil fuels, primarily by steam methane reforming (SMR) of liquefied natural gas. Blue hydrogen incorporates carbon capture and storage (CCS) technology into the gray hydrogen process to reduce carbon emissions; however, both methods still rely on fossil fuels and contribute to greenhouse gas emissions. In contrast, green hydrogen is produced using electricity to electrolyze water, generating hydrogen without harmful byproducts [3].
Water electrolysis involves anodic oxygen evolution reaction (OER) and cathodic hydrogen evolution reaction (HER). Although the theoretical potential for water electrolysis is 1.23 V, the actual potential required in an electrolyzer is significantly higher because of overpotentials at both the anode and cathode. Particularly in HER, precious Pt-based catalysts are typically used, but their high cost and limited availability hinder the widespread adoption of water electrolysis systems. Therefore, developing efficient and cost-effective electrocatalysts for HER is crucial.
Recently, various catalysts of transition metal-based compounds, such as nitrides [4,5,6], carbides [7,8,9], phosphides [10,11,12], and sulfides [13,14,15,16,17,18], have been extensively studied as non-precious HER catalysts. Among these materials, nickel sulfide stands out due to its cost-effectiveness, natural abundance, excellent electrical conductivity, and stability in alkaline environments, where it remains insoluble within electrolytes [19]. These properties make nickel sulfide a promising replacement for traditional Pt-based catalysts. Notably, nickel sulfide exists in various stoichiometries and has been reported as a HER catalyst in several forms, including hexagonal NiS, rhombohedral NiS, cubic NiS2, trigonal Ni3S2, cubic Ni3S4, and tetragonal Ni9S8 (Figure 1). However, when used alone, nickel sulfide exhibited significantly lower activity and stability for HER compared to Pt-based catalysts. To address this limitation, numerous studies have explored strategies to enhance HER performance, such as heteroatom doping, the construction of nickel sulfide-based heterostructures, and vacancy engineering. These approaches aim at optimizing the electronic structure and surface properties of nickel sulfide to improve HER performance.
In this review, we summarize and categorize research findings on enhancing the performance of nickel sulfide-based catalysts, focusing on their synthesis methods, characterizations, and HER performance according to their stoichiometries. The limitations and potential improvements of these catalysts are also discussed.

2. Mechanisms of HER in Alkaline Media

The overall water splitting reaction occurring in the electrolysis is as follows, involving a two-electron transfer reaction:
H2O → H2 + 1/2 O2
In an alkaline environment, the HER occurs via two mechanisms at the cathode (Figure 2). Initially, water molecules react with electrons, leading to the dissociation of water into adsorbed hydrogen atoms (denoted as Hads) and hydroxide ions (OH). This initial reaction is known as the Volmer step:
H2O + e → Hads + OH
Following the Volmer step, the Hads can follow two distinct pathways to form hydrogen gas (H2). The first pathway is the Tafel step, where two Hads atoms on the catalyst surface combine chemically to produce molecular hydrogen (H2):
Hads + Hads → H2
In contrast, the second pathway is the Heyrovsky step, which involves an electrochemical process. In this pathway, a Hads reacts with a water molecule and an additional electron, producing hydrogen gas (H2) and a hydroxide ion (OH):
Hads + H2O + e → H2 + OH
The determination of the Tafel and Heyrovsky pathways depends on the desorption method of Hads. If Hads is released through chemical desorption, it follows the Tafel step, whereas if the release involves electrochemical desorption, it proceeds through the Heyrovsky step. These two steps occur simultaneously during the HER process, and depending on which step is dominant, the mechanism is referred to as either the Volmer–Tafel mechanism or the Volmer–Heyrovsky mechanism.

3. Stoichiometric Nickel Sulfide-Based Catalysts for HER

3.1. NiS-Based Catalysts

3.1.1. Hexagonal NiS-Based Catalysts

Li et al. synthesized NiS/NiSe2 via a one-step hydrothermal method (Figure 3a) [20]. Scanning electron microscopy (SEM) images confirmed the formation of a polyhedral shape, while energy dispersive X-ray spectroscopy (EDS) elemental mapping showed that Ni and Se were uniformly distributed within the nanoparticles, with S concentrated in the inner regions. Adjusting the concentrations of S and Se during synthesis (Nix–Sey–Sz, where x, y, and z denote mmol of each element) revealed that an increase in S concentration (Sz) led to irregular morphologies and smaller particle sizes. Controlling the morphology through concentration adjustments significantly affected the HER activity. NiS/NiSe2 (denoted as Ni4–Se2–S2), with its uniform polyhedral shape, exhibited a low overpotential of 155 mV at 10 mA cm−210) in 1.0 M KOH (Figure 3b). This heterostructure demonstrated superior HER performance compared to NiS (Ni4–S4, η10 = 689 mV) and NiSe2 (Ni4–Se4, η10 = 502 mV).
Mo2N/NiS was prepared by annealing followed by hydrothermal treatment (Figure 3c) [21]. Initially, ammonium molybdate was decomposed via annealing under an NH3 atmosphere to afford a porous Mo2N substrate. Subsequent hydrothermal treatment with Ni and S sources yielded the Mo2N/NiS heterostructure. The SEM images confirmed that the hydrothermal treatment resulted in NiS particles anchored to the porous Mo2N structure (Figure 3d). Due to the formation of the heterostructure, Mo2N/NiS exhibited improved activity in 1.0 M KOH, with an overpotential η10 = 254 mV, compared with Mo2N (η10 = 355 mV) and NiS (η10 = 300 mV) (Figure 3e). To understand the enhanced HER activity resulting from heterostructure formation, density functional theory (DFT) calculations were conducted. The model shows that Mo2N/NiS is connected by Mo–S bonds within the heterostructure, with electron accumulation at the S sites, consistent with X-ray photoelectron spectroscopy (XPS) results. Additionally, when the HER occurs, H2O first adsorbs on the Mo-site of Mo2N, followed by proton adsorption on the S site of NiS. Subsequently, OH desorbs, leading to the formation of H* and the desorption of H2. The calculated Gibbs free energy of the HER process indicated that water dissociation occurred more readily on NiS than on Mo2N, whereas stronger H* adsorption on NiS minimized the HER activity. However, when the Mo2N/NiS heterostructure was formed, water dissociation was facilitated, and the H* adsorption energy was optimized, leading to enhanced HER activity.
Shi et al. synthesized a nitrogen-doped carbon (NC) coated NiS–CeO2 directly on Ni foam using a hydrothermal method followed by annealing [22]. SEM images revealed that nanorod structures formed during the initial hydrothermal process. After annealing with thiourea, a coated morphology was observed on the existing nanorods. This coating layer was confirmed by Fourier transform infrared spectroscopy (FT-IR), which identified a C–S adsorption peak, and XPS analysis of the C 1s and N 1s spectra confirmed the presence N-doped carbon on the nanorods. The authors suggested that NC layer enhances conductivity and stability in alkaline media. The NC/NiS–CeO2 catalyst demonstrated excellent HER activity with overpotential of η10 = 47 mV and η50 = 108 mV, outperforming Ni–CeO210 = 207 mV and η50 = 352 mV) before the formation of the NiS and NC layer. Furthermore, the introduction of CeO2 improved activity compared to NC/NiS alone (η10 = 122 mV and η50 = 219 mV).
MoS2/NiS catalyst was synthesized using a one-step hydrothermal method, resulting in a 3D needle-like morphology directly grown on Ni foam (NF), as observed in SEM images [23]. High-resolution SEM confirmed that the framework was composed of 2D nanosheets, and transmission electron microscopy (TEM) images showed MoS2/NF was coated on the needle-like hierarchical nanostructure. This unique morphology contributed to the superior HER performance of MoS2/NiS/NF, which exhibited a low overpotential of η10 = 87 mV in 1.0 M KOH, surpassing those of the single components NiS (η10 = 191 mV) and MoS210 = 145 mV). Furthermore, the MoS2/NiS/NF exhibited nearly twice the double-layer capacitance (Cdl) compared to individual components (NiS/NF: 2.38 mF cm−2, MoS2/NF: 14.2 mF cm−2), with MoS2/NiS/NF showing 27.6 mF cm−2, indicating a significantly enhanced electrochemical active surface area. The HER performance of recently reported hexagonal NiS-based HER catalysts are summarized in Table 1.

3.1.2. Rhombohedral NiS-Based Catalysts

Shi et al. synthesized MoS2/NiS using a two-step electrodeposition process followed by a hydrothermal treatment (Figure 4a) [28]. Initially, a two-step electrodeposition formed layered Ni(OH)2–Ni(OH)2/carbon cloth (CC) structures. MoS2 was then generated via hydrothermal treatment of the Ni(OH)2/CC, resulting in MoS2/NiS/CC. Compared with NiS/CC (η20 = 134 mV) and MoS2/CC (η20 = 224 mV), MoS2/NiS (NM2020, where 20 denotes the applied current density during each electrodeposition step) demonstrated significantly enhanced HER activity, achieving an η20 of 97 mV (Figure 4b). The applied current density variations influenced the electrodeposition rate and consequently the crystal core size. Applying a high current density (50 mA cm−2) during the first step and a low current density (30 mA cm−2) in the second step resulted in the formation of loose networks in the inner layer and dense structures in the outer layer. This configuration resulted in the synthesized catalyst (NM5030) exhibiting the highest HER activity, with η10 = 18 mV and η100 = 93 mV, attributed to the lattice strain induced by the structural differences between the two layers.
ReS2/NiS catalyst was prepared using a two-step hydrothermal process (Figure 4c) [29]. Initially, NiS nanowires were formed on the Ni foam via vulcanization. In the second step, these nanowires were combined with a Re precursor to form ReS2/NiS, where ReS2 nanosheets was deposited onto the NiS nanowires. This unique morphology resulted in ReS2/NiS exhibiting an overpotential of η10 = 78 mV in 1.0 M KOH, demonstrating superior HER activity to the single-phase of NiS nanowires (Figure 4d). Notably, at current densities above 325 mA cm−2, the ReS2/NiS catalyst outperformed the Pt/C catalyst. In-situ Raman spectroscopy was employed to investigate the enhanced HER activity by comparing fresh samples with those subjected to the HER reaction at 150 mA cm−2 for 10 h (denoted as “tested samples”). For NiS, no bands were observed in the tested samples; however, for ReS2 and ReS2/NiS, an S–Hads band appeared at 2708 cm−1 and 2682 cm−1, respectively (Figure 4e). The red shift in the S–Hads band of ReS2/NiS indicates that the S–Hads bond is weaker than in the single phase of ReS2. Consequently, the formation of a heterostructure creates a more favorable surface state for HER. Work function derived from DFT calculations revealed that the Fermi level of NiS is higher than that of ReS2, indicating that a Schottky junction is created when these two materials form a heterojunction. This drives electron transfer from NiS to ReS2, consistent with the XPS results. During the HER in the single-phase NiS and ReS2, OH favored adsorption onto the Ni and Re sites, respectively, while H transferred to the S sites. However, NiS exhibited weak water adsorption, and ReS2 formed a strong S-Hads bond, both of which hindered HER activity. In contrast, in the NiS/ReS2 heterostructure, electron transfer from NiS to ReS2 resulted in OH adsorption on the Ni site of NiS and H+ adsorption on the S site of ReS2, resulting in a more suitable energy state for HER.
Huang et al. directly synthesized NiS/MoS2 on calcined carbon paper (CP) using a one-step hydrothermal method [30]. The calcination process altered the hydrophobic/ aerophilic surface of CP to hydrophilic/aerophobic, as confirmed by contact angle measurements. This transformation enhanced the interaction between water and the electrode during the HER, facilitating the detachment of the generated gas from the electrode surface. Consequently, NiS/MoS2 on the calcined CP demonstrated notable HER performance with a η10 = 119 mV in 1.0 M KOH, outperforming the individual components NiS/CP (η10 = 225 mV) and MoS2/CP (η10 = 127 mV).
Jiang et al. synthesized a MoS2/NiS core-shell structure via a one-step hydrothermal method [31]. SEM images revealed that MoS2/NiS formed nanorods on nickel foam, in contrast to the nanoparticle morphology of MoS2 and the nanosheet structure form of NiS. TEM images revealed that MoS2/NiS had fluffy surfaces with both MoS2 and NiS coexisting. After removing the surface components by ultrasonication, only NiS remained, indicating that NiS formed the backbone with MoS2 coating the surface. This morphological effect resulted in excellent HER activity in 1.0 M KOH, achieving η10 = 84 mV, which significantly improved upon the lower activity of NiS alone (η10 = 176 mV) due to the heterostructure formation with MoS2. The HER performance of recently reported rhombohedral NiS-based HER catalysts is summarized in Table 2.

3.2. NiS2-Based Catalysts

Liu et al. synthesized Co–doped CeO2/NiS2 on Ni foam (Co–CeO2/NiS2/NF) using a hydrothermal method followed by annealing [34]. During the hydrothermal step, Ce, Ni, and Co precursors formed an intermediate, which was then vulcanized under an Ar atmosphere. SEM images revealed that Co–CeO2/NiS2/NF exhibited a nanoflower-like morphology with nanosheets directly formed on Ni foam, providing a larger specific surface area. The double-layer capacitance (Cdl) value increased significantly with the introduction of Co doping (Co–NiS2/NF, 62.61 mF cm−2) or the formation of a heterostructure with CeO2 (NiS2/CeO2/NF, 61.92 mF cm−2) compared to single-phase NiS2/NF (25.37 mF cm−2). The highest Cdl value of 83.52 mF cm−2 was observed in the Co–NiS2/CeO2/NF, demonstrating that Co doping and heterostructure formation with CeO2 in NiS2 significantly enhances the electrochemical active surface area. The Co–CeO2/NiS2/NF catalyst exhibited improved HER performance in 1.0 M KOH, achieving an η10 value of 84 mV, outperforming single NiS2/NF (η10 = 149 mV), the catalyst without Co doping, NiS2/CeO2/NF (η10 = 108 mV), and the catalyst without heterostructure, Co–NiS2/NF (η10 = 109 mV).
Li et al. synthesized NiS2/MoS2/CNTs using a hydrothermal process followed by etching process in NH4F solution (Figure 5a) [35]. SEM images showed that NiS2/MoS2/CNTs exhibited a flower-like morphology on the CNT surface, unlike MoS2/CNTs and NiS2/CNTs. After the etching process, the nanospheres were partially destroyed, indicating the partial removal of the NiS2/MoS2. Raman spectroscopy detected Mo3S13 edge sites at 819, 889, and 937 cm−1 were detected after the etching process, indicating the formation of abundant Mo–S edge sites. These Mo–S edge sites have insufficient coordination compared to the basal plane, facilitating easier hydrogen adsorption and desorption. The etching process significantly enhanced the HER activity, reducing the overpotential from 178 to 149 mV at 10 mA cm−2 in 1.0 M KOH. This heterostructure exhibited superior performance compared to single-phase MoS210 = 316 mV) and NiS210 = 260 mV), demonstrating enhanced catalytic activity (Figure 5b).
NiS2/Ni3C@C catalyst was prepared via a two-step hydrothermal process followed by a two-step calcination process (Figure 5c) [36]. Initially, α-Ni(OH)2 was synthesized via hydrothermal treatment. Glucose was then added as a carbon source, followed by an additional hydrothermal treatment to coat the catalyst with carbon. The catalyst underwent calcination at 700 °C, followed by a second calcination at 450 °C using thiourea as the sulfur source, forming NiS2/Ni3C@C. SEM images revealed that NiS2/Ni3C@C exhibited a peapod-like morphology with nanoparticles embedded within amorphous carbon, while TEM images indicated that Ni3C encapsulated the NiS2 nanoparticles. NiS2/Ni3C@C demonstrated an η10 value of 78 mV in 1.0 M KOH, which was significantly lower than that of NiS2@C (η10 = 226 mV) (Figure 5d). Furthermore, at current densities exceeding 200 mA cm−2, NiS2/Ni3C@C showed superior activity compared to Pt/C. DFT calculations revealed that the density of states (DOS) of Ni3C exhibits a metal-like structure, whereas defected carbon displays a semiconductor-like structure, suggesting the formation of a Schottky barrier at the Ni3C/C interface. This configuration facilitates faster electron and proton diffusion, adjusting the high Gibbs free energy of hydrogen adsorption in NiS2 to near zero through heterostructure formation.
Vanadium-doped NiS2 on carbon cloth (V–NiS2/CC) catalyst was synthesized via a hydrothermal method followed by annealing [37]. For the single NiS2/CC catalyst, SEM images showed that NiS2 aggregated on the carbon cloth. However, V doping, resulted in vertically grown nanosheets on the carbon cloth, indicating that V introduction promotes the formation of a morphology with a larger surface area. As a result, V–NiS2/CC exhibited a lower η10 value of 85 mV than NiS2/CC (η10 = 115 mV). DFT calculations indicated that V doping alters the electronic structure of NiS2 from semiconductor to metallic, enhancing its electrical conductivity. Additionally, V doping significantly reduced the water dissociation energy barrier, thereby lowering the overpotential. The HER performance of recently reported NiS2-based HER catalysts is summarized in Table 3.

3.3. Ni3S2-Based Catalysts

Tang et al. synthesized Co–MoS2/Ni3S2/NF via a one-step hydrothermal process [46]. SEM images showed that, both Co–MoS2/Ni3S2/NF and undoped MoS2/Ni3S2/NF formed flower-like aggregates, irrespective of Co introduction. However, XPS analysis identified doublets in the Mo4+ peak, indicating the co-existence of both 1T–MoS2 and 2H–MoS2 phases in Co–MoS2/Ni3S2/NF and MoS2/Ni3S2/NF. Notably, Co doping significantly increased the proportion of the 1T–MoS2 phase, suggesting that Co promotes the phase transition from 2H–MoS2 to 1T–MoS2. The increased 1T–MoS2, content enabled Co–MoS2/Ni3S2/NF to achieve remarkable HER performance, with η10 and η100 values of 43 and 201 mV, respectively, in 1.0 M KOH. This represents a significant improvement compared to MoS2/Ni3S2/NF (η10 = 107 mV and η100 = 355 mV) and surpasses the activity of Ni3S2/NF (η10 = 245 mV and η100 = 494 mV), highlighting the advantages of heterostructure formation.
Ni3S2–Ni3N–Co2N0.67/NF (Ni3S2@NiCoN/NF) catalyst was prepared via a two-step hydrothermal process followed by annealing (Figure 6a) [47]. Initially, Ni3S2/NF was synthesized via conducting hydrothermal treatment of Ni foam with thiourea. Cobalt chloride introduced in a subsequent hydrothermal step to form Ni3S2@Co(OH)2. Finally, the catalyst was annealed under an ammonia (NH3) flow, serving as a nitrogen source, to synthesize Ni3S2@NiCoN/NF. SEM images revealed that the introduction of Ni3N and Co2N0.67 into Ni3S2/NF, transformed the original cone structure of Ni3S2/NF into a core–shell triangular cone structure, which exposed more active sites. This structural transformation led to Ni3S2@NiCoN/NF demonstrating superior HER activity compared to Ni3S2/NF (η10 = 215 mV), with η10, η50 and η100 values of 63, 141 and 174 mV, respectively in 1.0 M KOH (Figure 6b).
Hu et al. synthesized Ni3S2 decorated with amorphous MoS2 (A–MoS2–Ni3S2–NF) using a one-step hydrothermal method (Figure 6c) [48]. SEM images revealed that A–MoS2–Ni3S2 formed an urchin-like morphology composed of assembled nanorods, and SEM-EDS measurements confirmed the presence of amorphous MoS2 based on the observed atomic ratios. TEM images further demonstrated that Ni3S2 formed the core, with amorphous MoS2 acting as the shell (Figure 6d). XPS spectra indicate electron transfer from inner Ni3S2 to outer amorphous MoS2 in the core–shell structure. Additionally, the introduction of amorphous MoS2 likely induced defects on the basal plane, maximizing the exposure of active sites. Consequently, A–MoS2–Ni3S2–NF exhibited outstanding HER performance, with η10 of 95 mV and η100 of 191 mV in 1.0 M KOH, significantly surpassing single Ni3S2–NF (η10 = 198 mV, η100 = 359 mV) (Figure 6e). The superior activity of A–MoS2–Ni3S2–NF can be attributed to the electron-rich environment of amorphous MoS2 on the catalyst’s surface, which promotes water dissociation and facilitates the Volmer and Heyrovsky process through rapid electron transfer reactions.
Li, V co-doped Ni3S2 (Li, V–Ni3S2) catalysts were prepared via a one-step hydrothermal process [49]. SEM images revealed that undoped Ni3S2 formed a granular shape on the Ni foam, whereas V doping yielded a well-arranged nanorod array. With Li and V co-doping, the nanorods developed a rougher surface compared to V–doped Ni3S2, suggesting enhanced provision of active sites, thereby facilitating electron and mass transfer. The Li, V–Ni3S2 exhibited excellent HER activity, with η10 of 90 mV and η100 of 183 mV in 1.0 M KOH, outperforming single Ni3S210 = 167 mV and η100 = 324 mV). Notably, a 2 × 2 cm2 single cell using Li, V–Ni3S2 as both the cathode and anode was employed for overall water splitting. It demonstrated exceptional activity at high current densities of 500 and 1000 mA cm−2, with cell potentials of 1.92 and 2.02 V, respectively. The single cell remained stable for 200 h under operation at 1000 mA cm−2.
Zhang et al. synthesized Ni3S2/MoS2 via a two-step hydrothermal process (Figure 7a) [50]. MoS2 was first formed on the carbon cloth (CC) followed by the growth of Ni3S2 on the MoS2. SEM images revealed the formation of Ni3S2/CC nanoparticles ranging in size from 6 to 22 nm on the MoS2 nanosheets (Figure 7b). After incorporating Ni3S2 into MoS2, X-ray diffraction (XRD) patterns showed a negative shift in the MoS2 (002) peak. Additionally, the HRTEM images of Ni3S2/MoS2/CC indicated that the layer spacing of MoS2 increased with more discontinuous lattice fringes compared to MoS2 alone. These findings suggest that Ni3S2 incorporation introduced defects and exposed more active sites on MoS2, enhancing its activity. Ni3S2/MoS2/CC exhibited superior activity compared to MoS2/CC and Ni3S2/CC, achieving an excellent performance of 189.4 mV at 100 mA cm−2 in 1.0 M KOH (Figure 7c) and outperforming Pt/C at current densities above 261.7 mA cm−2. Tafel slope reveals that the rate-determining step (RDS) for Ni3S2/CC (140.94 mVdec−1), MoS2 (97.34 mVdec−1), Ni3S2/MoS2/CC(81.01 mVdec−1) is the Volmer step. The formation of the heterostructure results in small Ni3S2 particles being present on MoS2, where Ni3S2, which effectively attracts oxygen-containing groups, adsorbs OH, while MoS2 strongly binds to protons. This synergistic effect promotes the dissociation of water, leading to a reduction in the Tafel slope.
Yang et al. synthesized VS4/Ni3S2/NF via a one-step solvothermal process using vanadium (III) acetylacetonate and thioacetamide as the vanadium and sulfur sources, respectively (Figure 7d) [51]. SEM and TEM images confirmed that VS4 nanoparticles decorate the Ni3S2 nanobelt array grown on the Ni foam. To investigate the impact of the vanadium source on morphology, sodium metavanadate was used instead of vanadium acetylacetonate, resulting in the formation of nanorod arrays, highlighting the role of vanadium acetylacetonate in forming the nanobelt structure. The VS4/Ni3S2/NF exhibited significantly enhanced HER activity compared to Ni3S2/NF (η10 = 195 mV and η100 = 365 mV), achieving η10 of 140 mV and η100 of 268 mV in 1.0 M KOH (Figure 7e). DFT calculations indicated that the S–S bonds in VS4/Ni3S2 are crucial for enhancing HER activity (Figure 7f). During the HER process, H2O initially bonds to the S–S bonds, forming S–Hads. The bidentate S-S bond spontaneously opens, accumulating electrons around the S atom that are then transferred to Hads, enhancing HER activity. Subsequently, Hads is released as H2, and the opened bridge spontaneously closes, returning to its initial state. This reversible process is repeated, supporting sustained HER performance.
Co2P–Ni3S2/NF catalyst was fabricated through a two-step heat treatment process (hydrothermal followed by annealing) [52]. Initially, Co–Ni3S2/NF was synthesized via a hydrothermal reaction and then annealed under an N2 atmosphere with a phosphorus source to form Co2P–Ni3S2/NF. SEM images revealed that Co2P–Ni3S2/NF formed a uniform nanowire structure on the Ni foam, with the phosphating process resulting in thinner and finer nanowires. Co2P–Ni3S2/NF exhibited low overpotential at high current densities and superior activity compared to P–Ni3S2/NF, achieving η100 = 110 mV, η500 = 164 mV, η1000 = 196 mV in 1.0 M KOH, while P–Ni3S2 showed η100 = 262 mV and η500 = 385 mV. DFT calculations suggested that single Ni3S2 and Co2P have strong binding energies for H*, while Co2P–Ni3S2 exhibited near-zero Gibbs free energy of H*, enhancing the HER process.
In studies comparing NiS, NiS2, and Ni3S2, it has been reported that Ni3S2 exhibits relatively high activity [53,54]. The shorter bond lengths between Ni-S and Ni-Ni in Ni3S2 lead to stronger metalmetal bonding interactions. Furthermore, DOS calculations have shown that Ni3S2 possesses stronger metallic properties compared to NiS and NiS2, resulting in enhanced electrical conductivity and improved HER activity. As a result, Ni3S2-based catalysts have been the most extensively studied. The HER performance of recently reported Ni3S2-based HER catalysts is summarized in Table 4.

3.4. Ni3S4-Based Catalysts

Ni3S4@Ni(OH)2 catalyst was synthesized via a two-step hydrothermal method followed by electrodeposition (Figure 8a) [86]. The two-step hydrothermal process resulted in the formation of Ni3S4 on the Ni foam, followed by electrodeposition in an electrolyte containing 0.7 M NiCl2·6H2O to form Ni3S4@Ni(OH)2. SEM images revealed a morphology of interlocked nanosheets stacked together. The introduction of Ni(OH)2 onto Ni3S4 transformed its hydrophobic surface into a hydrophilic one, reducing the contact angle to nearly 0°. Ni3S4@Ni(OH)2 catalyst exhibited excellent HER activity with η100 of 212.6 mV in 1.0 M KOH, significantly outperforming single-phase Ni3S4100 = 312.4 mV) and Ni(OH)2100 = 350.4 mV) (Figure 8b). Furthermore, Ni3S4@Ni(OH)2 was employed as both anode and cathode materials in an anion exchange membrane water electrolyzer (AEMWE) system. The Ni3S4@Ni(OH)2|| Ni3S4@Ni(OH)2 setup exhibited superior performance, achieving a cell potential of 1.84 V at 500 mA cm−2, surpassing the noble metal-based IrO2||Pt/C catalyst system at high current densities. During 100 h of operation at 500 mA cm−2, the Ni3S4||Ni3S4 configuration showed a 3.7% potential increase, whereas Ni3S4@Ni(OH)2|| Ni3S4@Ni(OH)2 exhibited only a 1.9% increase, indicating that the introduction of Ni(OH)2 significantly enhanced the stability of Ni3S4.
Shi et al. synthesized a molybdenum-doped Ni3S4 catalyst grown on carbonized wood (Mo–Ni3S4/CW) via a hydrothermal process (Figure 8c) [87]. SEM images revealed that the CW exhibited a porous structure with microchannel walls, which enhanced the interaction between the electrolyte and the catalyst material. The Mo–Ni3S4 formed rough nanosheets on the CW, contributing to its excellent catalytic activity with overpotential of η10 = 17 mV and η100 = 270 mV in 1.0 M KOH. Mo doping resulted in superior activity compared to Ni3S4/CW (η10 = 280 mV) and Pt/C (η10 = 96 mV and η100 = 284 mV) (Figure 8d). In-situ Raman spectroscopy showed an increase in the intensity of Ni–S and Mo–S bands upon an applied potential, indicating that Mo partially substituted Ni and intercalated into Ni3S4, expanding the lattice distance and enhancing electronic interactions (Figure 8e). DFT calculations indicated that the water dissociation energy and hydrogen adsorption Gibbs free energy at the Mo sites in Mo–Ni3S4 were close to zero, the lowest values among the compared sites, suggesting that Mo not only improved the intrinsic activity of the Ni sites in Ni3S4 but also played a critical role in the overall catalytic performance enhancement.
Ge et al. synthesized Ni3S4–MoS2 using a one-step hydrothermal method [88]. The SEM images showed that Ni3S4–MoS2 formed hierarchical nanospheres composed of nanosheets, with the heterostructure resulting in larger and thinner nanosheets than MoS2. This structural advantage provided Ni3S4-MoS2 with superior HER performance, achieving η10 = 116 mV in 1.0 M KOH, compared to single-phase MoS210 = 235 mV) and Ni3S410 = 318 mV). DFT calculations were performed to evaluate the chemisorption free energies of OH (ΔEOH) and H (ΔEH), revealing that Ni3S4 had a lower ΔEOH, whereas MoS2 had a lower ΔEH, indicating a preference for OH chemisorption on Ni3S4 and H chemisorption on MoS2 in the heterostructure. Furthermore, Ni3S4–MoS2 demonstrated a lower free energy barrier for water dissociation than either single-component Ni3S4 or MoS2, highlighting the improved HER performance of the hetrostructure. The HER performance of recently reported Ni3S4-based catalysts is summarized in Table 5.

3.5. Ni9S8-Based Catalysts

Gao et al. synthesized Ru-doped Ni9S8 with S vacancies (Vs–Ru–Ni9S8) using a hydrothermal process followed by low-temperature annealing at 100 °C (Figure 9a) [91]. SEM images revealed that Vs–Ru–Ni9S8 exhibited a layered rock-like morphology, while TEM images confirmed its 2D nanosheet structure. Electron paramagnetic resonance (EPR) spectral measurements revealed that Vs–Ru–Ni9S8, unlike Ru–Ni9S8, displayed a strong symmetric EPR signal at g = 2.003, indicating the successful formation of S vacancies. In 1.0 M KOH, Vs–Ru–Ni9S8 demonstrated superior HER activity with η10 = 94 mV compared to Ru-Ni9S810 =123 mV) (Figure 9b). The presence of S vacancies facilitated rapid ion transfer and exposed more active sites, contributing to the enhanced catalytic performance.
Chen et al. synthesized Ni9S8/MoS2@NiMoO4 through a hydrothermal process followed by annealing under an Ar atmosphere (Figure 9c) [92]. SEM images revealed that Ni9S8/MoS2@NiMoO4 formed nanorods with a porous structure, unlike the smooth surface of NiMoO4. TEM images confirmed that 1D NiMoO4 nanorods formed the core, whereas the Ni9S8/MoS2 nanosheets constituted the shell (Figure 9d). These structural characteristics, which exposed more active sites and provided rapid charge transfer pathways, led to Ni9S8/MoS2@NiMoO4 exhibiting excellent HER activity with η10 = 190 mV in 1.0 M KOH and significantly outperforming single-phase NiMoO410 = 434 mV) (Figure 9e). The HER performance of recently reported Ni9S8-based HER catalysts is summarized in Table 6.

3.6. Heterostructured Catalysts Between Nickel Sulfides

Zhang et al. prepared Mo-doped NiS/Ni3S2 on Ni foam containing S vacancies (Mo–NiS/Ni3S2–Sv/NF), as illustrated in Figure 10a [93]. The synthesis involved a hydrothermal method to produce a Ni-Mo precursor, followed by thermal treatment under an Ar atmosphere to obtain NiMoO4/NF. Subsequent hydrothermal treatment with a sulfur source yielded Mo–NiS/Ni3S2/NF (denoted as Mo–NiS/Ni3S2–0.08S). To introduce S vacancies (Sv), the material was etched in HCl, and the vacancy amount was controlled by varying the etching times (0, 30, 60, and 90 min), resulting in samples named Mo–NiS/Ni3S2-free Sv, Mo–NiS/Ni3S2-poor Sv, Mo–NiS/Ni3S2-rich Sv, and Mo–NiS/Ni3S2-excess Sv. SEM images confirmed the formation of nanorod-aggregated spheres in Mo–NiS/Ni3S2/NF. The heterostructured Mo–NiS/Ni3S2 catalyst demonstrated enhanced HER activity with η100 of 167 mV in 1.0 M KOH, outperforming Mo–NiS (η100 = 322 mV) and Mo–Ni3S2100 = 283 mV) (Figure 10b). The catalysts with S vacancies exhibited superior HER activity compared to S-free catalysts, with the Mo–NiS/Ni3S2-rich Sv sample (etched for 60 min) achieving a remarkable performance of 230 mV at 100 mA cm−2 without iR compensation. In situ Raman spectra indicated that the S–H band of Ni3S2 appeared at a lower potential when S vacancies were present, indicating that S vacancies facilitate the formation of the S–H intermediate more easily, thereby enhancing HER activity. Furthermore, the intensity of the S–H band over reaction time showed that in the presence of S vacancies, the intensity gradually increased, whereas in the absence of S vacancies, the intensity increased abruptly (Figure 10c). This behavior suggests that S vacancies prevent the accumulation of S-H on the catalyst surface. This allows the proton, generated from water dissociation at the Ni site of Mo-NiS, to interact properly with Ni3S2 and enhance catalytic performance.
The Mn–NiS/Mn–Ni3S4 (denoted as Mn-NiSx) catalyst was prepared by electrochemical deposition followed by annealing (Figure 10d) [94]. Initially, Mn–Ni(OH)2 was synthesized via electrodeposition, and subsequent vulcanization with sulfur powder under an Ar atmosphere produced Mn–NiS/Mn–Ni3S4. SEM images revealed that the disordered lamellar spheres of Mn–Ni(OH)2 transformed into a well-ordered arrangement of nano-micro spheres with a specific pattern, indicating that Mn doping and vulcanization contributed to the regular morphology, unlike the disordered structure observed without Mn doping. This ordered morphology, induced by Mn doping and vulcanization, created surface areas highly favorable for HER. Mn–NiS/Mn–Ni3S4 demonstrated superior activity with an overpotential of η10 = 94.2 mV and η100 = 267 mV in 1.0 M KOH, outperforming irregular Mn–Ni(OH)210 = 197 mV and η100 = 373 mV) and NiS/Ni3S410 = 172 mV and η100 = 407 mV) (Figure 10e).
Chen et al. synthesized Ni3S2@NiS through a hydrothermal process followed by a two-step annealing process (Figure 11a) [95]. First, Ni(OH)2 nanowires were synthesized via the hydrothermal method followed by annealing with sulfur powder to form Ni3S2. The nanowires were then annealed in air at 200, 250, and 300 °C to yield Ni3S2@NiS, denoted as Ni3S2@NiS–200/NF, Ni3S2@NiS–250/NF, and Ni3S2@NiS–300/N, respectively. SEM images showed a uniform wire-like morphology, with nanowires forming an interconnected 3D network. XPS spectra confirmed the presence of Ni2+ and Ni3+ in both Ni3S2/NF and Ni3S2@NiS, with the proportion of Ni2+ increasing with higher calcination temperatures. Ni3S2@NiS–250/NF exhibited the highest Ni2+ proportion, indicating the highest NiS content, which correlated with enhanced HER performance (Figure 11b). Ni3S2@NiS–250/NF exhibited the best HER activity with η10 = 129 mV in 1.0 M KOH, with all heterostructure catalysts outperforming single-phase Ni3S210 = 298 mV), as shown in Figure 11c. DFT calculations revealed that Ni3S2@NiS has a hydrogen adsorption free energy (ΔGH*) closer to zero compared to single-phase NiS and Ni3S2, indicating that the heterostructure effectively optimizes hydrogen binding energy, enhancing the catalytic performance. The HER performance of recently reported heterostructured catalysts between nickel sulfides is summarized in Table 7.

4. Summary and Perspective

To achieve truly sustainable net-zero carbon emissions, hydrogen production must transition to clean green hydrogen production, which generates no greenhouse gas emissions. For widespread implementation of water electrolysis systems, efficient catalysts are essential for both cathodic HER and anodic OER. Currently, the typical catalysts are precious Pt-based materials, underscoring the urgent need for highly active and stable non-precious alternatives. Nickel sulfide is a promising candidate for HER because of its low cost, abundance, stability in electrolytes, and excellent electrical conductivity. Various nickel sulfide catalysts, including NiS, NiS2, Ni3S2, Ni3S4, and Ni8S9, have been explored as HER catalysts; however, their single-phase forms often exhibit low activity and poor stability, limiting their ability to replace noble metal-based catalysts. To address this issue, strategies such as heteroatom doping, heterostructure construction, and vacancy engineering have been employed. This review summarizes the recent research progress of nickel sulfide-based catalysts for HER, categorized by stoichiometry, along with their synthetic methods, characterization techniques, and impacts on HER performance.
For the further development of nickel sulfide-based catalysts, several considerations should be considered. First, a more systematic approach is needed for nickel sulfide catalysts. Ni3S2-based catalysts have been predominantly studied, while NiS, NiS2, Ni3S4, and Ni8S9 have received less attention. To enhance HER activity, nickel sulfide catalysts are often combined with other components, such as metals, metal sulfides, metal oxides, and metal hydroxides, to form heterostructure catalysts. However, the intrinsic properties of stoichiometric nickel sulfides and the effects of these additional components on HER performance are poorly understood. Therefore, systematic investigations into the interactions between stoichiometric nickel sulfides and other components are crucial for improving the activity and stability of nickel sulfide-based catalysts.
Second, the identification of active sites and the mechanisms behind the enhanced HER activity of nickel sulfide-based catalysts requires further exploration via using in situ characterization techniques. Although many studies have focused primarily on HER performance, understanding active sites and activity origins has often relied heavily on computational simulations. Given that most strategies involve heteroatom doping, heterostructure construction, and vacancy engineering, there is an urgent need for experimental evidence regarding the nature of real active sites and their origins. Advanced in situ characterizations, such as in situ Raman spectroscopy or in situ X-ray absorption spectroscopy (XAS), can provide valuable insights into the active sites and activity mechanism. A combined approach using in situ characterization techniques and computational simulations can significantly advance the development of nickel sulfide-based catalysts for HER.
Lastly, for industrial applications, improvements in HER performance and synthesis methods are necessary. Few reports have indicated that nickel sulfide-based catalysts exhibit superior activity compared with typical noble metal-based catalysts, particularly when operating at high current densities (e.g., ~ Acm−2). Moreover, limited studies have investigated the performance of nickel sulfide-based catalysts in single-cell systems. To assess the feasibility of nickel sulfide-based catalysts for water electrolysis systems, it is essential to demonstrate their operation at high current densities at the single-cell level. Additionally, economical synthesis methods are critical for mass production. Current synthesis processes often involve multistep thermal treatments, which limit the scalability and cost-effectiveness. Therefore, more straightforward synthesis methods are needed, along with processes that eliminate the need for additional post-synthesis treatments of the catalyst.
Water electrolysis is a promising method for hydrogen production, and the development of highly active and stable non-precious catalysts is essential for widespread adoption. Although various nickel sulfide-based catalysts have been studied for HER, further advancements are needed to improve their performance for practical applications. By combining systematic investigations into the interactions between stoichiometric nickel sulfides and other components with in situ characterizations, researchers can gain insights into the active sites and HER mechanisms, ultimately leading to the design of more efficient nickel sulfide-based catalysts.

Author Contributions

Conceptualization, Y.C. and D.H.Y.; methodology, Y.C.; investigation, Y.C. and J.-H.L.; data curation, Y.C. and J.-H.L.; writing—original draft preparation, Y.C. and J.-H.L.; writing—review and editing, D.H.Y.; visualization, Y.C. and J.-H.L.; supervision, D.H.Y.; project administration, D.H.Y.; funding acquisition, D.H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry, and Energy (MOTIE) of the Republic of Korea (No. 20224000000080).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rogelj, J.; Geden, O.; Cowie, A.; Reisinger, A. Net-zero emissions targets are vague: Three ways to fix. Nature 2021, 591, 365–368. [Google Scholar] [CrossRef] [PubMed]
  2. Davis, S.J.; Lewis, N.S.; Shaner, M.; Aggarwal, S.; Arent, D.; Azevedo, I.L.; Benson, S.M.; Bradley, T.; Brouwer, J.; Chiang, Y.M.; et al. Net-zero emissions energy systems. Science 2018, 360, 1419. [Google Scholar] [CrossRef] [PubMed]
  3. Le, P.A.; Trung, V.D.; Nguyen, P.L.; Bac Phung, T.V.; Natsuki, J.; Natsuki, T. The current status of hydrogen energy: An overview. RSC Adv. 2023, 13, 28262–28287. [Google Scholar] [CrossRef]
  4. Liang, J.; Zhang, B.; Shen, H.; Yin, Y.; Liu, L.; Ma, Y.; Wang, X.; Xiao, C.; Kong, J.; Ding, S. Self-supported nickel nitride nanosheets as highly efficient electrocatalysts for hydrogen evolution. Appl. Surf. Sci. 2020, 503, 144143. [Google Scholar] [CrossRef]
  5. Dong, Y.; Zhang, X.; Wang, X.; Liu, F.; Ren, J.; Wang, H.; Wang, R. Kirkendall effect Strengthened-Superhydrophilic/superaerophobic Co-Ni3N/NF heterostructure as electrode catalyst for High-current hydrogen production. J. Colloid Interface Sci. 2023, 636, 657–667. [Google Scholar] [CrossRef] [PubMed]
  6. Park, S.H.; Jo, T.H.; Lee, M.H.; Kawashima, K.; Mullins, C.B.; Lim, H.-K.; Youn, D.H. Highly active and stable nickel–molybdenum nitride (Ni2Mo3N) electrocatalyst for hydrogen evolution. J. Mater. Chem. A 2021, 9, 4945–4951. [Google Scholar] [CrossRef]
  7. Hwang, J.Y.; Kim, J.Y.; Youn, D.H. In-situ construction of Mo2C–MoC heterostructure on Ni foam for efficient hydrogen evolution in alkaline media. Int. J. Hydrogen Energy 2024, 49, 58–66. [Google Scholar] [CrossRef]
  8. Diaz-Coello, S.; Winkler, D.; Griesser, C.; Moser, T.; Rodriguez, J.L.; Kunze-Liebhauser, J.; Garcia, G.; Pastor, E. Highly Active W2C-Based Composites for the HER in Alkaline Solution: The Role of Surface Oxide Species. ACS Appl. Mater. Interfaces 2024, 16, 21877–21884. [Google Scholar] [CrossRef]
  9. Youn, D.H.; Han, S.; Kim, J.Y.; Kim, J.Y.; Park, H.; Choi, S.H.; Lee, J.S. Highly Active and Stable Hydrogen Evolution Electrocatalysts Based on Molybdenum Compounds on Carbon Nanotube–Graphene Hybrid Support. ACS Nano 2014, 8, 5164–5173. [Google Scholar] [CrossRef]
  10. Liu, S.; Li, Z.; Chang, Y.; Gyu Kim, M.; Jang, H.; Cho, J.; Hou, L.; Liu, X. Substantial Impact of Built-in Electric Field and Electrode Potential on the Alkaline Hydrogen Evolution Reaction of Ru-CoP Urchin Arrays. Angew. Chem., Int. Ed. 2024, 63, e202400069. [Google Scholar] [CrossRef]
  11. Boppella, R.; Tan, J.; Yang, W.; Moon, J. Homologous CoP/NiCoP Heterostructure on N-Doped Carbon for Highly Efficient and pH-Universal Hydrogen Evolution Electrocatalysis. Adv. Funct. Mater. 2018, 29, 1807976. [Google Scholar] [CrossRef]
  12. Lee, M.H.; Youn, D.H.; Lee, J.S. Nanostructured molybdenum phosphide/N-doped carbon nanotube-graphene composites as efficient electrocatalysts for hydrogen evolution reaction. Appl. Catal. A 2020, 594, 117451. [Google Scholar] [CrossRef]
  13. Wang, Z.; Zhang, X.; Gao, Y.; Wang, H. Transition metal (Fe, Co, Ni) doping enhanced monolith catalysts of 1T−MoS2/Mo for the large-current hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 56, 75–83. [Google Scholar] [CrossRef]
  14. Qian, Y.; Sun, Y.; Zhang, F.; Luo, X.; Li, K.; Shen, L.; Shi, H.; Kang, D.J.; Pang, H. In situ construction of layered transition metal phosphides/sulfides heterostructures for efficient hydrogen evolution in acidic and alkaline media. Chem. Eng. J. 2024, 490, 151693. [Google Scholar] [CrossRef]
  15. Youn, D.H.; Jang, J.-W.; Kim, J.Y.; Jang, J.S.; Choi, S.H.; Lee, J.S. Fabrication of graphene-based electrode in less than a minute through hybrid microwave annealing. Sci. Rep. 2014, 4, 5492. [Google Scholar] [CrossRef]
  16. Ali Shah, S.; Sayyar, R.; Xu, L.; Sun, H.; Khan, I.; Guo, J.; Shen, X.; Hussain, S.; Yuan, A.; Ullah, H. In-situ synthesis of NiS2 nanoparticles/MoS2 nanosheets hierarchical sphere anchored on reduced graphene oxide for enhanced electrocatalytic hydrogen evolution reaction. J. Colloid Interface Sci. 2022, 624, 150–159. [Google Scholar] [CrossRef]
  17. Khan, H.; Shah, S.A.; Rehman, W.U.; Chen, F. CoS2 Nanoparticles-Decorated MoS2/rGO Nanosheets as An Efficient Electrocatalyst for Ultrafast Hydrogen Evolution. Adv. Mater. Interface 2021, 9, 2101294. [Google Scholar] [CrossRef]
  18. Shah, S.A.; Shen, X.; Xie, M.; Zhu, G.; Ji, Z.; Zhou, H.; Xu, K.; Yue, X.; Yuan, A.; Zhu, J.; et al. Nickel@Nitrogen-Doped Carbon@MoS2 Nanosheets: An Efficient Electrocatalyst for Hydrogen Evolution Reaction. Small 2019, 15, e1804545. [Google Scholar] [CrossRef]
  19. Wang, S.; Geng, Z.; Bi, S.; Wang, Y.; Gao, Z.; Jin, L.; Zhang, C. Recent advances and future prospects on Ni3S2-Based electrocatalysts for efficient alkaline water electrolysis. Green Energy Environ. 2024, 9, 659–683. [Google Scholar] [CrossRef]
  20. Li, Y.; Du, H.; Su, Y.; Zhao, J.; Qu, K.; Zhang, X.; Zhang, Y.; Dong, Y.; Guo, Z. Construction of Heterostructured NiS/NiSe2 and their application in electrocatalytic water splitting. Int. J. Hydrogen Energy 2024, 66, 286–293. [Google Scholar] [CrossRef]
  21. Zang, Y.; Huang, S.; Yang, B.; Chen, G.; Liu, X.; Zhang, N. Constructing collaborative interface between Mo2N and NiS as efficient bifunctional electrocatalysts for overall water splitting. Appl. Surf. Sci. 2023, 611, 155656. [Google Scholar] [CrossRef]
  22. Shi, M.; Li, N.; Fu, W.; Sun, G.; Wu, M.; Li, Q.; Shen, W.; Ma, J. Spindle Nanorods of CeO2 and NiS Heterointerface Coated by the NC Layer: A High-Performance Bifunctional Electrocatalyst for Water Splitting. Langmuir 2023, 39, 17929–17938. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, X.; Zhong, W.; Zhang, L.; Liu, G.; Du, Y. MoS2/NiS heterostructure grown on Nickel Foam as highly efficient bifunctional electrocatalyst for overall water splitting. Int. J. Hydrogen Energy 2020, 45, 17329–17338. [Google Scholar] [CrossRef]
  24. Pitchai, C.; Gopalakrishnan, S.M.; Chen, C.M. Ultra-efficient Nitrogen-Doped Carbon Dots-Supported Nickel Sulfide as a Platinum-Free Electrocatalyst for Overall Water Splitting in Basic Medium. Energy Fuels 2024, 38, 2235–2247. [Google Scholar] [CrossRef]
  25. Dang, J.; Yun, S.; Zhang, Y.; Yang, G.; Yang, J.; Qiao, D.; Yang, T. Designing nitrogen-enriched heterogeneous NiS@CoNi2S4 embedded in nitrogen-doped carbon with hierarchical 2D/3D nanocage structure for efficient alkaline hydrogen evolution and triiodide reduction. J. Colloid Interface Sci. 2023, 630, 91–105. [Google Scholar] [CrossRef]
  26. Liu, G.; Thummavichai, K.; Lv, X.; Chen, W.; Lin, T.; Tan, S.; Zeng, M.; Chen, Y.; Wang, N.; Zhu, Y. Defect-Rich Heterogeneous MoS2/rGO/NiS Nanocomposite for Efficient pH-Universal Hydrogen Evolution. Nanomaterials 2021, 11, 662. [Google Scholar] [CrossRef]
  27. Chen, Z.; Liu, X.; Xin, P.; Wang, H.; Wu, Y.; Gao, C.; He, Q.; Jiang, Y.; Hu, Z.; Huang, S. Interface engineering of NiS@MoS2 core-shell microspheres as an efficient catalyst for hydrogen evolution reaction in both acidic and alkaline medium. J. Alloys Compd. 2021, 853, 157352. [Google Scholar] [CrossRef]
  28. Shi, C.; Boda, M.A.; Zhao, K.; Zhou, Y.; Ma, H.; Yi, Z. Boosting hydrogen production from alkaline water splitting by regulating interlayer stress via lattice mismatch in NiS/MoS2. Int. J. Hydrogen Energy 2024, 62, 397–404. [Google Scholar] [CrossRef]
  29. Liu, Y.; Zhang, H.; Song, W.; Zhang, Y.; Hou, Z.; Zhou, G.; Zhang, Z.; Liu, J. In-situ growth of ReS2/NiS heterostructure on Ni foam as an ultra-stable electrocatalyst for alkaline hydrogen generation. Chem. Eng. J. 2023, 451, 1238905. [Google Scholar] [CrossRef]
  30. Huang, L.; Li, Z.; Sun, S.; Sun, G.; Li, Y.; Han, S.; Lian, J. NiS/MoS2 complex grown on carbon paper as a bifunctional electrocatalyst for full water splitting. J. Alloys Compd. 2022, 926, 166870. [Google Scholar] [CrossRef]
  31. Jiang, H.; Zhang, K.; Li, W.; Cui, Z.; He, S.-A.; Zhao, S.; Li, J.; He, G.; Shearing, P.R.; Brett, D.J.L. MoS2/NiS core-shell structures for improved electrocatalytic process of hydrogen evolution. J. Power Sources 2020, 472, 228497. [Google Scholar] [CrossRef]
  32. Prakash, C.; Sahoo, P.; Yadav, R.; Pandey, A.; Singh, V.K.; Dixit, A. Nanoengineered Zn-modified Nickel Sulfide (NiS) as a bifunctional electrocatalyst for overall water splitting. Int. J. Hydrogen Energy 2023, 48, 21969–21980. [Google Scholar] [CrossRef]
  33. Zhou, H.; Liu, J.; Xu, C. NiFeCr–S/NiS/NF with modified 2D nanosheet array as bifunctional electrocatalysts for overall water splitting. Int. J. Hydrogen Energy 2023, 48, 33088–33097. [Google Scholar] [CrossRef]
  34. Liu, F.; Zhang, X.; Zong, H.; Xu, H.; Xu, J.; Liu, J. Facile construction Co-doped and CeO2 heterostructure modified NiS2 grown on foamed nickel for high-performance bifunctional water electrolysis catalysts. Int. J. Hydrogen Energy 2024, 51, 314–322. [Google Scholar] [CrossRef]
  35. Li, G.L.; Miao, Y.Y.; Qiao, X.Y.; Wang, T.Y.; Deng, F. Engineering edge sites based on NiS2/MoS2/CNTs heterojunction catalyst for overall water splitting. Appl. Surf. Sci. 2023, 615, 156309. [Google Scholar] [CrossRef]
  36. Feng, Y.; Guan, Y.; Zhou, E.; Zhang, X.; Wang, Y. Nanoscale Double-Heterojunctional Electrocatalyst for Hydrogen Evolution. Adv. Sci. 2022, 9, e2201339. [Google Scholar] [CrossRef]
  37. Zhang, K.; Jia, J.; Tan, L.; Qi, S.; Li, B.; Chen, J.; Li, J.; Lou, Y.; Guo, Y. Morphological and electronic modification of NiS2 for efficient supercapacitors and hydrogen evolution reaction. J. Power Sources 2023, 577, 232239. [Google Scholar] [CrossRef]
  38. Yang, Z.; Chen, H.; Bei, S.; Bao, K.; Zhang, C.; Xiang, M.; Yu, C.; Dong, S.; Qin, H. Ultralow RuO2 Doped NiS2 Heterojunction as a Multifunctional Electrocatalyst for Hydrogen Evolution linking to Biomass Organics Oxidation. Small 2024, 20, e2310286. [Google Scholar] [CrossRef]
  39. Zhou, W.; Li, X.; Li, X.; Shao, J.; Yang, H.; Chai, X.; Hu, Q.; He, C. Crafting amorphous VO2–crystalline NiS2 heterostructures as bifunctional electrocatalysts for efficient water splitting: The different cocatalytic function of VO2. Chem. Eng. J. 2023, 470, 144146. [Google Scholar] [CrossRef]
  40. He, N.; Chen, X.; Fang, B.; Li, Y.; Lu, T.; Pan, L. Zr-MOF/NiS2 hybrids on nickel foam as advanced electrocatalysts for efficient hydrogen evolution. J. Colloid Interface Sci. 2023, 640, 820–828. [Google Scholar] [CrossRef]
  41. Tong, L.; Liu, Y.; Song, C.; Zhang, Y.; Latthe, S.S.; Liu, S.; Xing, R. (Fe, Ni)S2@MoS2/NiS2 hollow heterostructure nanocubes for high-performance alkaline water electrolysis. Int. J. Hydrogen Energy 2022, 47, 11143–11152. [Google Scholar] [CrossRef]
  42. Cao, N.; Chen, S.; Di, Y.; Li, C.; Qi, H.; Shao, Q.; Zhao, W.; Qin, Y.; Zang, X. High efficiency in overall water-splitting via Co-doping heterointerface-rich NiS2/MoS2 nanosheets electrocatalysts. Electrochim. Acta 2022, 425, 140674. [Google Scholar] [CrossRef]
  43. Tong, L.; Song, C.; Liu, Y.; Xing, R.; Sekar, K.; Liu, S. Open and porous NiS2 nanowrinkles grown on non-stoichiometric MoOx nanorods for high-performance alkaline water electrolysis and supercapacitor. Int. J. Hydrogen Energy 2022, 47, 14404–14413. [Google Scholar] [CrossRef]
  44. Xu, J.; Rong, J.; Zheng, Y.; Zhu, Y.; Mao, K.; Jing, Z.; Zhang, T.; Yang, D.; Qiu, F. Construction of sheet-on-sheet hierarchical MoS2/NiS2 heterostructures as efficient bifunctional electrocatalysts for overall water splitting. Electrochim. Acta 2021, 385, 138438. [Google Scholar] [CrossRef]
  45. Huang, S.; Jin, Z.; Ning, P.; Gao, C.; Wu, Y.; Liu, X.; Xin, P.; Chen, Z.; Jiang, Y.; Hu, Z.; et al. Synergistically modulating electronic structure of NiS2 hierarchical architectures by phosphorus doping and sulfur-vacancies defect engineering enables efficient electrocatalytic water splitting. Chem. Eng. J. 2021, 420, 127630. [Google Scholar] [CrossRef]
  46. Tang, X.; Zhang, Y.; Zhang, Y.; Liu, X.; Cao, J.; Zhang, C. Three-dimensional Co–MoS2/Ni3S2/Ni assembly with interfacial engineering and electronic modulation for efficient hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 67, 42–49. [Google Scholar] [CrossRef]
  47. Wang, Y.; Wang, Y.; Fu, W.; Gao, G.; Liu, Y.; Wang, L. Self-templated synthesis of Ni3S2@NiCoN with core-shell structure for effective hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 59, 1419–1426. [Google Scholar] [CrossRef]
  48. Hu, M.; Qian, Y.; Yu, S.; Yang, Q.; Wang, Z.; Huang, Y.; Li, L. Amorphous MoS2 Decorated Ni3S2 with a Core-shell Structure of Urchin-Like on Nickel-Foam Efficient Hydrogen Evolution in Acidic and Alkaline Media. Small 2024, 20, e2305948. [Google Scholar] [CrossRef]
  49. Ha, Q.N.; Susanto Gultom, N.; Yeh, C.H.; Kuo, D.H. One-pot synthesized Li, V co-doped Ni3S2 nanorod arrays as a bifunctional electrocatalyst for industrialization-facile hydrogen production via alkaline exchange membrane water electrolysis. Chem. Eng. J. 2023, 472, 144931. [Google Scholar] [CrossRef]
  50. Zhang, N.; Li, Y.; Zhang, R.; Huang, S.; Wang, F.; Tang, M.; Liu, J. Tiny Ni3S2 boosting MoS2 hydrogen evolution in alkali by enlarging coupling boundaries and stimulating basal plane. J. Colloid Interface Sci. 2023, 642, 479–487. [Google Scholar] [CrossRef]
  51. Yang, D.; Cao, L.; Huang, J.; Jiao, G.; Wang, D.; Liu, Q.; Li, G.; He, C.; Feng, L. Reversible active bridging sulfur sites grafted on Ni3S2 nanobelt arrays for efficient hydrogen evolution reaction. J. Colloid Interface Sci. 2023, 649, 194–202. [Google Scholar] [CrossRef] [PubMed]
  52. Li, H.; Gao, X.; Li, G. Construction of Co2P-Ni3S2 /NF Heterogeneous Structural Hollow Nanowires as Bifunctional Electrocatalysts for Efficient Overall Water Splitting. Small 2023, 19, e2304081. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. Zheng, X.; Han, X.; Zhang, Y.; Wang, J.; Zhong, C.; Deng, Y.; Hu, W. Controllable synthesis of nickel sulfide nanocatalysts and their phase-dependent performance for overall water splitting. Nanoscale 2019, 11, 5646–5654. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, W.; Dong, J.; An, B.; Su, H.; Teng, Z.; Li, N.; Gao, Y.; Ge, L. Synergistic dual built-in electric fields in 1T-MoS2/Ni3S2/LDH for efficient electrocatalytic overall water splitting reactions. J. Colloid Interface Sci. 2024, 673, 228–238. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, L.; Lv, W.; Yang, Y.; Yang, M.; Yang, W.; Tao, H.; Wang, Z.; Xiao, X.; Dong, X. One-step hydrothermal synthesis of nanowire-like W/Mo-Ni3S2/NF electrocatalysts for highly efficient hydrogen evolution reactions. J. Alloys Compd. 2024, 995, 174754. [Google Scholar] [CrossRef]
  57. Zhang, T.; Xu, D.; Liu, P.; Chen, L.; Guo, W.; Gu, T.; Yu, F.; Liu, Y.; Wang, G. Constructing defect-rich S-doped NiMoO4/Ni3S2 2D/2D heterostructures as highly efficient and stable bifunctional electrocatalysts for overall water splitting. J. Alloys Compd. 2024, 970, 172530. [Google Scholar] [CrossRef]
  58. Du, M.; Ji, Y.; Li, Y.; Liu, S.; Yan, J. Construction of an Internal Charge Field: CoS1.097/Ni3S2 Heterojunction Promotes Efficient Urea Oxidation Reaction. Adv. Funct. Mater. 2024, 34, 2402776. [Google Scholar] [CrossRef]
  59. Teng, X.; Wang, Z.; Wu, Y.; Zhang, Y.; Yuan, B.; Xu, Y.; Wang, R.; Shan, A. Enhanced alkaline hydrogen evolution reaction of MoO2/Ni3S2 nanorod arrays by interface engineering. Nano Energy 2024, 122, 109299. [Google Scholar] [CrossRef]
  60. Fang, B.; Jin, J.; Li, Y.; Dang, H.; Shao, M.; Zhao, L.; Yin, N.; Wang, W. Interfacial Electronic Modulation of Mo5N6/Ni3S2 Heterojunction Array Boosts Electrocatalytic Alkaline Overall Water Splitting. Small 2024, 20, e2310825. [Google Scholar] [CrossRef]
  61. Wang, Z.; Li, B.; Wang, L.; Chu, L.; Yang, M.; Wang, G. Amorphous-crystalline coupling CoS/NixPy/Fe-Ni3S2 heterostructure nanoplate array on NF for efficient overall water splitting. Electrochim. Acta 2023, 462, 142767. [Google Scholar] [CrossRef]
  62. Zhong, B.; Cheng, B.; Zhu, Y.; Ding, R.; Kuang, P.; Yu, J. Hierarchically porous nickel foam supported Fe-Ni3S2 electrode for high-current-density alkaline water splitting. J. Colloid Interface Sci. 2023, 629, 846–853. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, T.; Li, B.; Wang, P.; Xu, M.; Wang, D.; Wang, Y.; Zhang, W.; Qu, C.; Feng, M. Modulation of electronic structure of Ni3S2 via Fe and Mo co-doping to enhance the bifunctional electrocatalytic activities for HER and OER. J. Colloid Interface Sci. 2024, 672, 715–723. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, N.; Huang, S.; Chen, L.; Li, Y.; Tang, M.; Pei, Q.; Liu, J. Superhydrophilic/superaerophobic amorphous Ni3S2/NiMoS electrocatalyst for enhanced hydrogen evolution. J. Colloid Interface Sci. 2023, 652, 95–103. [Google Scholar] [CrossRef] [PubMed]
  65. Poudel, M.B.; Logeshwaran, N.; Kim, A.R.; Karthikeyan, S.C.; Vijayapradeep, S.; Yoo, D.J. Integrated core-shell assembly of Ni3S2 nanowires and CoMoP nanosheets as highly efficient bifunctional electrocatalysts for overall water splitting. J. Alloys Compd. 2023, 960, 170678. [Google Scholar] [CrossRef]
  66. Wang, X.; Yu, X.; Bai, J.; Yuan, G.; He, P.; Zhu, Y.; Wu, S.; Qin, F.; Ren, L. Interface engineering assisted Fe-Ni3S2/Ni2P heterostructure as a high-performance bifunctional electrocatalyst for OER and HER. Electrochim. Acta 2023, 458, 142524. [Google Scholar] [CrossRef]
  67. Liu, Y.; Ding, M.; Tian, Y.; Zhao, G.; Huang, J.; Xu, X. In-situ growth of 3D hierarchical gamma-FeOOH/Ni3S2 heterostructure as high performance electrocatalyst for overall water splitting. J. Colloid Interface Sci. 2023, 639, 24–32. [Google Scholar] [CrossRef]
  68. Li, W.; Sun, Z.; Ge, R.; Li, J.; Li, Y.; Cairney, J.M.; Zheng, R.; Li, Y.; Li, S.; Li, Q.; et al. Nanoarchitectonics of La-Doped Ni3S2/MoS2 Hetetostructural Electrocatalysts for Water Electrolysis. Small Struct. 2023, 4, 2300175. [Google Scholar] [CrossRef]
  69. Xu, X.; Zhao, X.; Zhao, R.; Gu, M.; Cheng, Y.; Zhang, J. Ni3S2@MoS2@Ni3Si2 Seaweed-like Hybrid Structures In-situ Grown on Ni Foam as Efficient Bifunctional Electrocatalysts. ChemCatChem 2023, 15, e202201412. [Google Scholar] [CrossRef]
  70. Shi, R.; Yang, J.; Zhou, G. High-index-faceted and electron density-optimized Ni3S2 in hierarchical NiWO4-Ni3S2@NiO/NF nanofibers for robust alkaline electrocatalytic hydrogen evolution. Chem. Eng. J. 2023, 457, 141188. [Google Scholar] [CrossRef]
  71. He, W.; Zhang, R.; Cao, D.; Li, Y.; Zhang, J.; Hao, Q.; Liu, H.; Zhao, J.; Xin, H.L. Super-Hydrophilic Microporous Ni(OH)x/Ni3S2 Heterostructure Electrocatalyst for Large-Current-Density Hydrogen Evolution. Small 2023, 19, e2205719. [Google Scholar] [CrossRef] [PubMed]
  72. Li, S.; Luo, W.; Gao, Q.; Shen, W.; Jiang, Y.; He, R.; Su, W.; Li, M. Strong electronic coupling induced by synergy of dopant and interface in Ru-Ni3S2/NixPy to boost efficient water splitting. Appl. Surf. Sci. 2023, 637, 157940. [Google Scholar] [CrossRef]
  73. Wang, Z.; Qian, G.; Yu, T.; Chen, J.; Shen, F.; Luo, L.; Zou, Y.; Yin, S. Carbon encapsulated FeWO4-Ni3S2 nanosheets as a highly active catalyst for overall water splitting at large current density. Chem. Eng. J. 2022, 434, 134669. [Google Scholar] [CrossRef]
  74. Liu, H.; Cheng, J.; He, W.; Li, Y.; Mao, J.; Zheng, X.; Chen, C.; Cui, C.; Hao, Q. Interfacial electronic modulation of Ni3S2 nanosheet arrays decorated with Au nanoparticles boosts overall water splitting. Appl. Catal. B 2022, 304, 120935. [Google Scholar] [CrossRef]
  75. Feng, X.; Jiao, Q.; Zhang, J.; Cui, H.; Li, H.; Zhao, Y.; Feng, C. Integrating Amorphous Molybdenum Sulfide Nanosheets with a Co9S8@Ni3S2 Array as an Efficient Electrocatalyst for Overall Water Splitting. Langmuir 2022, 38, 3469–3479. [Google Scholar] [CrossRef]
  76. Wu, Z.; Feng, Y.; Qin, Z.; Han, X.; Zheng, X.; Deng, Y.; Hu, W. Bimetallic Multi-Level Layered Co-NiOOH/Ni3S2@NF Nanosheet for Hydrogen Evolution Reaction in Alkaline Medium. Small 2022, 18, e2106904. [Google Scholar] [CrossRef]
  77. Li, R.; Kuang, P.; Wang, L.; Tang, H.; Yu, J. Engineering 2D NiO/Ni3S2 heterointerface electrocatalyst for highly efficient hydrogen production coupled with benzyl alcohol oxidation. Chem. Eng. J. 2022, 431, 134137. [Google Scholar] [CrossRef]
  78. Wang, H.; Wang, C.; Zhang, W.; Yao, S. A nano-spherical structure Ni3S2/Ni(OH)2 electrocatalyst prepared by one-step fast electrodeposition for efficient and durable water splitting. Int. J. Hydrogen Energy 2022, 47, 14916–14929. [Google Scholar] [CrossRef]
  79. Guo, D.; Wan, Z.; Fang, G.; Zhu, M.; Xi, B. A Tandem Interfaced Ni3S2-MoS2@TiO2 Composite Fabricated by Atomic Layer Deposition as Efficient HER Electrocatalyst. Small 2022, 18, e2201896. [Google Scholar] [CrossRef]
  80. Yang, S.; Guo, Y.; Zhao, Y.; Zhang, L.; Shen, H.; Wang, J.; Li, J.; Wu, C.; Wang, W.; Cao, Y.; et al. Construction of Synergistic Ni3S2-MoS2 Nanoheterojunctions on Ni Foam as Bifunctional Electrocatalyst for Hydrogen Evolution Integrated with Biomass Valorization. Small 2022, 18, e2201306. [Google Scholar] [CrossRef]
  81. Liu, H.; Ouyang, D.; Zhou, Q.; Feng, C. Successional heterostructure MoS2-Ni3S2 nanospheres based on 3D nano-porous Ni: An efficient electrocatalyst for overall water splitting. J. Alloys Compd. 2022, 920, 165243. [Google Scholar] [CrossRef]
  82. Wu, Q.; Dong, A.; Yang, C.; Ye, L.; Zhao, L.; Jiang, Q. Metal-organic framework derived Co3O4@Mo-Co3S4-Ni3S2 heterostructure supported on Ni foam for overall water splitting. Chem. Eng. J. 2021, 413, 127482. [Google Scholar] [CrossRef]
  83. Tong, X.; Li, Y.; Pang, N.; Qu, Y.; Yan, C.; Xiong, D.; Xu, S.; Wang, L.; Chu, P.K. Co-doped Ni3S2 porous nanocones as high-performance bifunctional electrocatalysts in water splitting. Chem. Eng. J. 2021, 425, 130455. [Google Scholar] [CrossRef]
  84. Jin, C.; Zhai, P.; Wei, Y.; Chen, Q.; Wang, X.; Yang, W.; Xiao, J.; He, Q.; Liu, Q.; Gong, Y. Ni(OH)2 Templated Synthesis of Ultrathin Ni3S2 Nanosheets as Bifunctional Electrocatalyst for Overall Water Splitting. Small 2021, 17, e2102097. [Google Scholar] [CrossRef]
  85. Xu, H.; Liao, Y.; Gao, Z.; Qing, Y.; Wu, Y.; Xia, L. A branch-like Mo-doped Ni3S2 nanoforest as a high-efficiency and durable catalyst for overall urea electrolysis. J. Mater. Chem. A 2021, 9, 3418–3426. [Google Scholar] [CrossRef]
  86. Chai, H.; Ma, X.; Dang, Y.; Zhang, Y.; Yue, F.; Pang, X.; Wang, G.; Yang, C. Triple roles of Ni(OH)2 promoting the electrocatalytic activity and stability of Ni3S4@Ni(OH)2 in anion exchange membrane water electrolyzers. J. Colloid Interface Sci. 2024, 654, 66–75. [Google Scholar] [CrossRef]
  87. Shi, Z.; Mao, C.; Zhong, L.; Peng, J.; Liu, M.; Li, H.; Huang, J. Mo-doped Ni3S4 nanosheets grown on carbonized wood as highly efficient and durable electrocatalysts for water splitting. Appl. Catal. B 2023, 339, 123123. [Google Scholar] [CrossRef]
  88. Ge, J.; Jin, J.; Cao, Y.; Jiang, M.; Zhang, F.; Guo, H.; Lei, X. Heterostructure Ni3S4-MoS2 with interfacial electron redistribution used for enhancing hydrogen evolution. RSC Adv. 2021, 11, 19630–19638. [Google Scholar] [CrossRef]
  89. Ren, Y.; Zhu, S.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. Hierarchical Ni3S4@MoS2 nanocomposites as efficient electrocatalysts for hydrogen evolution reaction. J. Mater. Sci. Tech. 2021, 95, 70–77. [Google Scholar] [CrossRef]
  90. Yao, Z.; Wang, J.; Wang, Y.; Xie, T.; Li, C.; Jiang, Z. Boosting electrocatalytic activity toward alkaline hydrogen evolution by strongly coupled ternary Ni3S4/Ni/Ni(OH)2 hybrid. Electrochim. Acta 2021, 382, 138342. [Google Scholar] [CrossRef]
  91. Gao, Q.; Luo, W.; Ma, X.; Ma, Z.; Li, S.; Gou, F.; Shen, W.; Jiang, Y.; He, R.; Li, M. Electronic modulation and vacancy engineering of Ni9S8 to synergistically boost efficient water splitting: Active vacancy-metal pairs. Appl. Catal. B 2022, 310, 121356. [Google Scholar] [CrossRef]
  92. Chen, L.; Deng, Z.; Chen, Z.; Wang, X. Building Ni9S8/MoS2 Nanosheets Decorated NiMoO4 Nanorods Heterostructure for Enhanced Water Splitting. Adv. Mater. Interfaces 2021, 8, 2101483. [Google Scholar] [CrossRef]
  93. Zhang, K.; Duan, Y.; Graham, N.; Yu, W. Unveiling the synergy of polymorph heterointerface and sulfur vacancy in NiS/Ni3S2 electrocatalyst to promote alkaline hydrogen evolution reaction. Appl. Catal. B 2023, 323, 122144. [Google Scholar] [CrossRef]
  94. Huang, H.; Hu, X.; Hou, Z.; Yang, D.; Xiang, D.; Hu, L. Interfacial construction and lattice distortion-triggered bifunctionality of Mn-NiS/Mn-Ni3S4 for H2 production. Fuel 2022, 328, 125337. [Google Scholar] [CrossRef]
  95. Chen, M.; Su, Q.; Kitiphatpiboon, N.; Zhang, J.; Feng, C.; Li, S.; Zhao, Q.; Abudula, A.; Ma, Y.; Guan, G. Heterojunction engineering of Ni3S2/NiS nanowire for electrochemical hydrogen evolution. Fuel 2023, 331, 125794. [Google Scholar] [CrossRef]
  96. Liu, C.; Zhang, X.; Cheng, Y.; Hao, X.; Yang, P. NiS/Ni3S4 decorated double-layered hollow carbon spheres for efficient electrochemical hydrogen evolution reaction and supercapacitor. Electrochim. Acta 2024, 477, 143751. [Google Scholar] [CrossRef]
  97. Wang, L.; Xue, X.; Luan, Q.; Guo, J.; Chu, L.; Wang, Z.; Li, B.; Yang, M.; Wang, G. Interface engineering of Mo-doped Ni9S8/Ni3S2 multiphase heterostructure nanoflowers by one step synthesis for efficient overall water splitting. J. Colloid Interface Sci. 2023, 634, 563–574. [Google Scholar] [CrossRef]
  98. Wang, W.; Wang, W.; Xu, Y.; Ren, X.; Liu, X.; Li, Z. Synthesis of Ni3S4/NiS2/FeS2 nanoparticles for hydrogen and oxygen evolution reaction. Appl. Surf. Sci. 2021, 560, 149985. [Google Scholar] [CrossRef]
  99. Wang, H.; Zhang, W.; Zhang, X.; Hu, S.; Zhang, Z.; Zhou, W.; Liu, H. Multi-interface collaboration of graphene cross-linked NiS-NiS2-Ni3S4 polymorph foam towards robust hydrogen evolution in alkaline electrolyte. Nano Res. 2021, 14, 4857–4864. [Google Scholar] [CrossRef]
  100. Zhou, B.; Li, J.; Zhang, X.; Guo, J. Engineering P-doped Ni3S2-NiS hybrid nanorod arrays for efficient overall water electrolysis. J. Alloys Compd. 2021, 862, 158391. [Google Scholar] [CrossRef]
  101. Yang, D.; Cao, L.; Feng, L.; Huang, J.; Feng, Y.; Liu, Q.; He, D.; Wang, J. Controlled Synthesis of V-Doped Heterogeneous Ni3S2/NiS Nanorod Arrays as Efficient Hydrogen Evolution Electrocatalysts. Langmuir 2021, 37, 357–365. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Primitive cell structure of nickel sulfides as a HER catalyst.
Figure 1. Primitive cell structure of nickel sulfides as a HER catalyst.
Molecules 29 04975 g001
Figure 2. Schematic illustration of the HER mechanisms in alkaline media.
Figure 2. Schematic illustration of the HER mechanisms in alkaline media.
Molecules 29 04975 g002
Figure 3. (a) Schematic illustration of the synthesis of NiS/NiSe2. (b) Linear sweep voltammetry (LSV) curves of Nix–Sey–Sz in 1.0 M KOH. (c) Schematic representation of the synthesis of Mo2N/NiS. (d) SEM image showing the morphology of Mo2N/NiS. (e) LSV curves comparing Mo2N/NiS with its comparison group in 1.0 M KOH. Reproduced with permission from Elsevier [20,21].
Figure 3. (a) Schematic illustration of the synthesis of NiS/NiSe2. (b) Linear sweep voltammetry (LSV) curves of Nix–Sey–Sz in 1.0 M KOH. (c) Schematic representation of the synthesis of Mo2N/NiS. (d) SEM image showing the morphology of Mo2N/NiS. (e) LSV curves comparing Mo2N/NiS with its comparison group in 1.0 M KOH. Reproduced with permission from Elsevier [20,21].
Molecules 29 04975 g003
Figure 4. (a) Schematic illustration of the synthesis of MoS2/NiS/CC. (b) LSV curves comparing MoS2/NiS/CC (denoted as NM2020) with its control group in 1.0 M KOH. (c) Schematic representation of the synthesis of ReS2/NiS. (d) LSV curves comparing ReS2/NiS with its control group in 1.0 M KOH. (e) Raman spectra of NiS, ReS2 and ReS2/NiS before (dotted) and after (solid) stability tests. Reproduced with permission from Elsevier [28,29].
Figure 4. (a) Schematic illustration of the synthesis of MoS2/NiS/CC. (b) LSV curves comparing MoS2/NiS/CC (denoted as NM2020) with its control group in 1.0 M KOH. (c) Schematic representation of the synthesis of ReS2/NiS. (d) LSV curves comparing ReS2/NiS with its control group in 1.0 M KOH. (e) Raman spectra of NiS, ReS2 and ReS2/NiS before (dotted) and after (solid) stability tests. Reproduced with permission from Elsevier [28,29].
Molecules 29 04975 g004
Figure 5. (a) Schematic illustration of the synthesis of NiS2/MoS2/CNTs. (b) LSV curves comparing NiS2/MoS2/CNTs with control group in 1.0 M KOH. (c) Schematic representation of the synthesis of NiS2/Ni3C@C. (d) LSV curves comparing NiS2/Ni3C@C with its control samples in 1.0 M KOH. Reproduced with permission from Elsevier [35] and Wiley [36].
Figure 5. (a) Schematic illustration of the synthesis of NiS2/MoS2/CNTs. (b) LSV curves comparing NiS2/MoS2/CNTs with control group in 1.0 M KOH. (c) Schematic representation of the synthesis of NiS2/Ni3C@C. (d) LSV curves comparing NiS2/Ni3C@C with its control samples in 1.0 M KOH. Reproduced with permission from Elsevier [35] and Wiley [36].
Molecules 29 04975 g005
Figure 6. (a) Schematic illustration of the synthesis of Ni3S2@NiCoN/NF. (b) LSV curves comparing of Ni3S2@NiCoN/NF with control groups in 1.0 M KOH. (c) Schematic representation of the synthesis of A–MoS2–Ni3S2–NF. (d) HRTEM image of A-MoS2-Ni3S2-NF. (e) LSV curves comparing of A–MoS2–Ni3S2–NF with control samples in 1.0 M KOH. Reproduced with permission from Elsevier [47] and Wiley [48].
Figure 6. (a) Schematic illustration of the synthesis of Ni3S2@NiCoN/NF. (b) LSV curves comparing of Ni3S2@NiCoN/NF with control groups in 1.0 M KOH. (c) Schematic representation of the synthesis of A–MoS2–Ni3S2–NF. (d) HRTEM image of A-MoS2-Ni3S2-NF. (e) LSV curves comparing of A–MoS2–Ni3S2–NF with control samples in 1.0 M KOH. Reproduced with permission from Elsevier [47] and Wiley [48].
Molecules 29 04975 g006
Figure 7. (a) Schematic illustration of the synthesis of Ni3S2/MoS2/CC. (b) SEM image showing the morphology of Ni3S2/MoS2/CC. (c) LSV curves comparing of Ni3S2/MoS2/CC with control samples in 1.0 M KOH. (d) Schematic representation of the synthesis of VS4/Ni3S2/NF NBs. (e) LSV curves comparing of VS4/Ni3S2/NF NBs with control groups in 1.0 M KOH. (f) Schematic diagram illustrating the reversible HER process of VS4/Ni3S2. Reproduced with permission from Elsevier [50,51].
Figure 7. (a) Schematic illustration of the synthesis of Ni3S2/MoS2/CC. (b) SEM image showing the morphology of Ni3S2/MoS2/CC. (c) LSV curves comparing of Ni3S2/MoS2/CC with control samples in 1.0 M KOH. (d) Schematic representation of the synthesis of VS4/Ni3S2/NF NBs. (e) LSV curves comparing of VS4/Ni3S2/NF NBs with control groups in 1.0 M KOH. (f) Schematic diagram illustrating the reversible HER process of VS4/Ni3S2. Reproduced with permission from Elsevier [50,51].
Molecules 29 04975 g007
Figure 8. (a) Schematic illustration of the synthesis of Ni3S4/Ni(OH)2. (b) LSV curves comparing of Ni3S4/Ni(OH)2 with control samples in 1.0 M KOH. (c) Schematic representation of the synthesis of Mo–Ni3S4/CW. (d) LSV curves comparing Mo–Ni3S4/CW with control groups in 1.0 M KOH. (e) In situ Raman spectra of Mo-Ni3S4/CW within the HER potential range. Reproduced with permission from Elsevier [86,87].
Figure 8. (a) Schematic illustration of the synthesis of Ni3S4/Ni(OH)2. (b) LSV curves comparing of Ni3S4/Ni(OH)2 with control samples in 1.0 M KOH. (c) Schematic representation of the synthesis of Mo–Ni3S4/CW. (d) LSV curves comparing Mo–Ni3S4/CW with control groups in 1.0 M KOH. (e) In situ Raman spectra of Mo-Ni3S4/CW within the HER potential range. Reproduced with permission from Elsevier [86,87].
Molecules 29 04975 g008
Figure 9. (a) Schematic illustration of the synthesis of Vs–Ru–Ni9S8. (b) LSV curves comparing of Vs–Ru–Ni9S8 and its comparison group in 1.0 M KOH. (c) Schematic representation of the synthesis of Ni9S8/MoS2@NiMoO4. (d) TEM image of Ni9S8/MoS2@NiMoO4. (e) LSV curves comparing of Ni9S8/MoS2@NiMoO4 and control samples in 1.0 M KOH. Reproduced with permission from Elsevier [91] and Wiley [92].
Figure 9. (a) Schematic illustration of the synthesis of Vs–Ru–Ni9S8. (b) LSV curves comparing of Vs–Ru–Ni9S8 and its comparison group in 1.0 M KOH. (c) Schematic representation of the synthesis of Ni9S8/MoS2@NiMoO4. (d) TEM image of Ni9S8/MoS2@NiMoO4. (e) LSV curves comparing of Ni9S8/MoS2@NiMoO4 and control samples in 1.0 M KOH. Reproduced with permission from Elsevier [91] and Wiley [92].
Molecules 29 04975 g009
Figure 10. (a) Schematic illustration of the synthesis of Mo–NiS/Ni3S2–Sv/NF. (b) LSV curves of Mo-NiS/Ni3S2/NF compared with control samples in 1.0 M KOH. (c) S–H stretching vibration intensity ratio from in situ Raman spectra comparing Mo–NiS/Ni3S2-free Sv/NF and Mo–NiS/Ni3S2-rich Sv/NF. (d) Schematic representation of the synthesis of Mn–NiS/Mn–Ni3S4. (e) LSV curves of Mn–NiS/Mn–Ni3S4 and its comparison group in 1.0 M KOH. Reproduced with permission from Elsevier [93,94].
Figure 10. (a) Schematic illustration of the synthesis of Mo–NiS/Ni3S2–Sv/NF. (b) LSV curves of Mo-NiS/Ni3S2/NF compared with control samples in 1.0 M KOH. (c) S–H stretching vibration intensity ratio from in situ Raman spectra comparing Mo–NiS/Ni3S2-free Sv/NF and Mo–NiS/Ni3S2-rich Sv/NF. (d) Schematic representation of the synthesis of Mn–NiS/Mn–Ni3S4. (e) LSV curves of Mn–NiS/Mn–Ni3S4 and its comparison group in 1.0 M KOH. Reproduced with permission from Elsevier [93,94].
Molecules 29 04975 g010
Figure 11. (a) Schematic representation of the synthesis of NiS@Ni3S2/NF. (b) XPS analysis of surface Ni2+ percentage of Ni3S2/NF and NiS@Ni3S2/NF samples annealed at 200, 250, and 300 °C. (c) LSV curves comparing of NiS@Ni3S2/NF and control samples in 1.0 M KOH. Reproduced with permission from Elsevier [95].
Figure 11. (a) Schematic representation of the synthesis of NiS@Ni3S2/NF. (b) XPS analysis of surface Ni2+ percentage of Ni3S2/NF and NiS@Ni3S2/NF samples annealed at 200, 250, and 300 °C. (c) LSV curves comparing of NiS@Ni3S2/NF and control samples in 1.0 M KOH. Reproduced with permission from Elsevier [95].
Molecules 29 04975 g011
Table 1. HER performances of hexagonal NiS-based catalysts.
Table 1. HER performances of hexagonal NiS-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
NiS/NiSe2GC 1)1.0 M KOHHydrothermal155 mV-[20]
Mo2N/NiSCFP 2)1.0 M KOHCalcination
and Hydrothermal
254 mV403 mV *[21]
NC 3)/NiS–CeO2NF 4)1.0 M KOHHydrothermal47 mV139 mV *[22]
MoS2/NiSNF1.0 M KOHHydrothermal87 mV189 mV *[23]
α–NiS@NDCS 5)NF1.0 M KOHHydrothermal173 mV-[24]
NiS@CoNi2S4/NC 6)NF1.0 M KOHCalcination126 mV210 mV *[25]
MoS2/rGO/NiSGC1.0 M KOHHydrothermal169 mV301 mV *[26]
NiS@MoS2NF1.0 M KOHCalcination
and Hydrothermal
146 mV-[27]
1) Glassy carbon. 2) Carbon fiber paper. 3) N-doped carbon coated. 4) Nickel foam. 5) N-doped carbon dots. 6) N-doped carbon. Data marked with * are estimated from the LSV.
Table 2. HER performance of the rhombohedral NiS-based catalysts.
Table 2. HER performance of the rhombohedral NiS-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
NiS/MoS2CC 1)1.0 M KOHElectrodeposition
and Hydrothermal
18 mV93 mV[28]
ReS2/NiSNF1.0 M KOHHydrothermal78 mV181 mV *[29]
NiS/MoS2CP 2)1.0 M KOHHydrothermal119 mV314 mV *[30]
MoS2/NiSNF1.0 M KOHHydrothermal84 mV168 mV * [31]
Zn–NiSNF1.0 M KOHHydrothermal208 mV-[32]
NiFeCr–S–NiSNF1.0 M KOHCalcination
and Hydrothermal
131 mV254 mV *[33]
1) Carbon cloth. 2) Carbon paper. Data marked with * are estimated from the LSV.
Table 3. HER performance of NiS2-based catalysts.
Table 3. HER performance of NiS2-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
Co–NiS2–CeO2NF1.0 M KOHCalcination
and Hydrothermal
88 mV 213 mV *[34]
NiS2/MoS2/CNTsGC1.0 M KOHHydrothermal149 mV 320 mV *[35]
NiS2–Ni3C@CGC1.0 M KOHCalcination
and Hydrothermal
78 mV 189 mV *[36]
V–NiS2CC1.0 M KOHCalcination
and Hydrothermal
85 mV 247 mV *[37]
RuO2/NiS2NF1.0 M KOHHydrothermal71 mV -[38]
VO2–NiS2CC1.0 M KOHHydrothermal96 mV 225 mV *[39]
Zr–MOF/NiS2NF1.0 M KOHCalcination72 mV -[40]
(Fe, Ni)S2@MoS2/NiS2NF1.0 M KOHCalcination
and Hydrothermal
91 mV 225 mV *[41]
Co–NiS2/MoS2CFP1.0 M KOHCalcination89 mV 166 mV[42]
MoOx@NiS2NF1.0 M KOHCalcination
and Hydrothermal
101 mV 323 mV *[43]
MoS2/NiS2CC1.0 M KOHHydrothermal80 mV 175 mV[44]
P–NiS2NF1.0 M KOHCalcination
and Solvothermal
73 mV 173 mV *[45]
Data marked with * are estimated from the LSV.
Table 4. HER performance of Ni3S2-based catalysts.
Table 4. HER performance of Ni3S2-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
Co–MoS2/Ni3S2NF1.0 M KOHHydrothermal43 mV201 mV[46]
Ni3S2@NiCoNNF1.0 M KOHCalcination
and Hydrothermal
63 mV174 mV[47]
A–MoS2/Ni3S2 1)NF1.0 M KOHHydrothermal95 mV191 mV[48]
Li, V–Ni3S2NF1.0 M KOHHydrothermal90 mV183 mV[49]
Ni3S2/MoS2CC1.0 M KOHHydrothermal105 mV *189 mV[50]
VS4/Ni3S2NF1.0 M KOHSolvothermal140 mV268 mV[51]
Co2P–Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
-110 mV[52]
1T–MoS2/Ni3S2/LDHNF1.0 M KOHElectrodeposition
and Hydrothermal
104 mV342 mV *[55]
W/Mo–Ni3S2NF1.0 M KOHHydrothermal136 mV271 mV *[56]
S–NiMoO4/Ni3S2NF1.0 M KOHElectrodeposition
and Hydrothermal
107 mV244 mV *[57]
CoS1.097/Ni3S2NF1.0 M KOHHydrothermal74 mV-[58]
MoO2/Ni3S2NF1.0 M KOHHydrothermal74 mV200 mV *[59]
Mo5N6/Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
59 mV313 mV *[60]
CoS/NixPy/Fe–Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
49 mV203 mV[61]
Fe–Ni3S2NF1.0 M KOHElectrodeposition-98 mV[62]
Fe–MoS2/Ni3S2NF1.0 M KOHHydrothermal74 mV235 mV[63]
Ni3S2/NiMoSNF1.0 M KOHElectrodeposition
and Hydrothermal
197 mV197 mV[64]
CoMoP–Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
97 mV192 mV *[65]
Fe–Ni3S2/Ni2PNF1.0 M KOHCalcination
and Hydrothermal
112 mV198 mV[66]
FeOOH/Ni3S2NF1.0 M KOHElectrodeposition92 mV232 mV *[67]
La–Ni3S2/MoS2NF1.0 M KOHHydrothermal-154 mV[68]
Ni3S2@MoS2@Ni3Si2NF1.0 M KOHHydrothermal84 mV143 mV *[69]
NiWO4–Ni3S2@NiONF1.0 M KOHElectrodeposition
and Hydrothermal
89 mV210 mV *[70]
Ni(OH)x/Ni3S2NF1.0 M KOHElectrochemical activation
and Hydrothermal
54 mV126 mV[71]
Ru–Ni3S2/NixPyNF1.0 M KOHCalcination
and Hydrothermal
51 mV126 mV *[72]
FeWO4–Ni3S2@C 2)NF1.0 M KOHCalcination
and Solvothermal
50 mV173 mV *[73]
Au–Ni3S2NF1.0 M KOHHydrothermal97 mV 188 mV *[74]
MoSx@Co9S8@Ni3S2NF1.0 M KOHHydrothermal77 mV180 mV *[75]
Co–NiOOH/Ni3S2NF1.0 M KOHElectrodeposition
and Hydrothermal
87 mV203 mV[76]
NiO/Ni3S2CC1.0 M KOHElectrodeposition91 mV210 mV *[77]
Ni3S2/Ni(OH)2NF1.0 M KOHElectrodeposition66 mV312 mV *[78]
(Ni3S2–MoS2)/TiO2NF1.0 M KOHALD 3)
and Hydrothermal
49 mV118 mV *[79]
Ni3S2–MoS2NF1.0 M KOHSolvothermal103 mV207 mV *[80]
MoS2–Ni3S2NF1.0 M KOHHydrothermal
and Calcination
109 mV214 mV *[81]
Co3O4@Mo–Co3S4–Ni3S2NF1.0 M KOHHydrothermal116 mV214 mV *[82]
Co–Ni3S2NF1.0 M KOHALD
and Hydrothermal
62 mV160 mV *[83]
Co–Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
102 mV (η20)158 mV[84]
Mo–Ni3S2NF1.0 M KOHHydrothermal90 mV176 mV *[85]
1) Amorphous. 2) Carbon encapsulated. 3) Atomic layer deposition. Data marked with * are estimated from the LSV.
Table 5. HER performance of Ni3S4-based catalysts.
Table 5. HER performance of Ni3S4-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
Ni3S4@Ni(OH)2NF1.0 M KOHElectrodeposition
and Hydrothermal
118 mV213 mV[86]
Mo–Ni3S4CW 1)1.0 M KOHHydrothermal17 mV270 mV[87]
Ni3S4–MoS2NF1.0 M KOHHydrothermal116 mV-[88]
Ni3S4@MoS2CC1.0 M KOHHydrothermal97 mV174 mV[89]
Ni3S4/Ni/Ni(OH)2TM 2)1.0 M KOHElectrodeposition54 mV185 mV * [90]
1) Carbonized wood. 2) Ti mesh. Data marked with * are estimated from the LSV.
Table 6. HER performance of Ni9S8-based catalysts.
Table 6. HER performance of Ni9S8-based catalysts.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
Vs–Ru–Ni9S8NF1.0 M KOHHydrothermal94 mV170 mV *[91]
Ni9S8/MoS2@NiMoO4NF1.0 M KOHCalcination
and Hydrothermal
190 mV-[92]
Data marked with * are estimated from the LSV.
Table 7. HER performance of heterostructured catalysts between nickel sulfides.
Table 7. HER performance of heterostructured catalysts between nickel sulfides.
CatalystSubstrateElectrolyteSynthesis MethodOverpotentialRef
η10η100
Mo–NiS/Ni3S2NF1.0 M KOHCalcination
and Hydrothermal
-167 mV[93]
Mn–NiS/Mn–Ni3S4NF1.0 M KOHCalcination
and Electrodeposition
94 mV 267 mV[94]
Ni3S2@NiSNF1.0 M KOHCalcination
and Hydrothermal
129 mV-[95]
NiS/Ni3S4GC1.0 M KOHCalcination
and Hydrothermal
263 mV495 mV *[96]
Mo–Ni9S8/Ni3S2NF1.0 M KOHHydrothermal116 mV240 mV[97]
Ni3S4/NiS2/FeS2NF1.0 M KOHHydrothermal196 mV-[98]
NiS–NiS2–Ni3S4NF1.0 M KOHCalcination
and Hydrothermal
68 mV-[99]
P–Ni3S2–NiSNF1.0 M KOHHydrothermal141 mV376 mV *[100]
V–Ni3S2–NiSNF1.0 M KOHHydrothermal85 mV218 mV[101]
Data marked with * are estimated from the LSV.
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

Choi, Y.; Lee, J.-H.; Youn, D.H. A Review of Stoichiometric Nickel Sulfide-Based Catalysts for Hydrogen Evolution Reaction in Alkaline Media. Molecules 2024, 29, 4975. https://doi.org/10.3390/molecules29204975

AMA Style

Choi Y, Lee J-H, Youn DH. A Review of Stoichiometric Nickel Sulfide-Based Catalysts for Hydrogen Evolution Reaction in Alkaline Media. Molecules. 2024; 29(20):4975. https://doi.org/10.3390/molecules29204975

Chicago/Turabian Style

Choi, Yeji, Jun-Hee Lee, and Duck Hyun Youn. 2024. "A Review of Stoichiometric Nickel Sulfide-Based Catalysts for Hydrogen Evolution Reaction in Alkaline Media" Molecules 29, no. 20: 4975. https://doi.org/10.3390/molecules29204975

APA Style

Choi, Y., Lee, J. -H., & Youn, D. H. (2024). A Review of Stoichiometric Nickel Sulfide-Based Catalysts for Hydrogen Evolution Reaction in Alkaline Media. Molecules, 29(20), 4975. https://doi.org/10.3390/molecules29204975

Article Metrics

Back to TopTop