Next Article in Journal
Modeling of Dry Reforming of Methane Using Artificial Neural Networks
Previous Article in Journal
Comparative Hydrogen Production Routes via Steam Methane Reforming and Chemical Looping Reforming of Natural Gas as Feedstock
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tuning the Morphology of Transition Metal Disulfides: Advances in Electrocatalysts for Hydrogen Evolution Reaction

by
Shravani S. Jakkanawar
1,
Vijay D. Chavan
2,
Deok-Kee Kim
2,
Tejasvinee S. Bhat
1,* and
Hemraj M. Yadav
1,*
1
School of Nanoscience and Biotechnology, Shivaji University, Kolhapur 416 004, MS, India
2
Department of Electrical Engineering and Convergence Engineering for Intelligent Drone, Sejong University, Seoul 05006, Republic of Korea
*
Authors to whom correspondence should be addressed.
Hydrogen 2024, 5(4), 776-799; https://doi.org/10.3390/hydrogen5040041
Submission received: 8 October 2024 / Revised: 26 October 2024 / Accepted: 30 October 2024 / Published: 2 November 2024

Abstract

:
The hydrogen evolution reaction (HER) in the renewable energy system has gained a lot of attention from researchers as hydrogen is assumed to be a clean and renewable carrier. Transition metals and their compounds have been used as promising alternatives to precious noble metals for the HER, offering low cost, more availability, and high activity. In this work, we discussed the mechanisms of the HER and how morphology influenced the catalytic performance of transition metal disulfide (TMD), focusing on structures that range from zero-dimensional (0D) to three-dimensional (3D) TMD materials. Notably, two-dimensional (2D) TMDs, like nanosheets, exhibit the lowest overpotential and a very small Tafel slope, which can be ascribed to their inherent layered structure and large surface area. According to recent research reports, the efficacy and efficiency of the HER process are influenced by surface chemistry, electrochemical characteristics, and the existence of active sites.

1. Introduction

Over the past centuries, immense amounts of fossil fuels have been used for energy storage and conversion technologies. However, burning fossil fuels emits large quantities of CO2 into the atmosphere, which results in increasing climate change and environmental deterioration [1,2]. To overcome these problems, we need to generate clean and renewable energy from renewable energy sources, such as wind, water, solar, and geothermal power [3,4].
Hydrogen is considered a high-density, clean, and renewable energy source; it is produced from fossil fuels, biomass, and water splitting [5]. Among these, water splitting is a promising way to generate sustainable pure hydrogen, which can serve as an important alternative to conventional fossil fuels in specific applications. There are two significant reactions involved in water splitting, including the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) at the anode and cathode, respectively [6,7]. During the water splitting process, an energy output of 237.1 kJ mol−1 is required, with a theoretical potential of 1.23 V. But the real potential required to achieve this exact current density varies with theoretical potential from 1.23 to 2.5 V due to the sluggish kinetics of the HER and OER. Precious noble metals like platinum are used to overcome this obstacle and enhance the reaction kinetics. However, the noble metals lead to agglomeration, dissolution, and low durability during water splitting [8].
Researchers have gained a lot of interest in transition metals and their compounds, such as oxides, hydroxides, phosphides, carbides, nitrides, and sulfides, as catalysts for water splitting applications [9,10,11,12,13,14]. Due to their cost-effectiveness, generous availability, and good catalytic nature, these materials are beginning to replace noble metals. Their novelty and potential enhance energy storage and conversion technologies [15,16]. Of these transition metal compounds, TMDs, such as VS2, CoS2, NiS2, MoS2, WS2, TaS2, etc., have frequently been utilized as catalysts in recent studies because of their huge surface area, flexible electronic structure, and intrinsic catalytic capacity [17,18,19,20,21]. In addition to water splitting, these TMDs are used in various applications, such as field effect transistors, lubrication, electrocatalysis, and photocatalysis [22,23,24,25].
For the improvement of the catalytic efficiency of TMDs, researchers are extensively introducing new strategies, such as the doping of metal and the incorporation of carbonaceous materials. The doping of metals can alter the electronic structure, provide active sites, and increase the conductivity of TMDs [26,27]. Similarly, carbonaceous materials, including graphene, carbon nanotubes (CNTs), and activated carbon, are combined with TMDs to increase surface area and conductivity and to provide more active sites [28,29,30,31,32]. Even though these strategies modify the morphology of TMDs, the morphology plays a vital role in the enhancement of catalytic HER activity by exposing more edge sites. Each dimension, from 0D to 3D, offers additional advantages in hydrogen evolution. In short, 0D nanoparticles provide a high density of active sites; the one-dimensional (1D) nanowire facilitates charge transport; 2D nanosheets have more exposed edge sites for catalytic reactions; and 3D structures offer a high surface area, high active site density, and favorable scalability for application [33,34,35,36].
Here, we discussed the development of various morphologies of TMDs towards the electrochemical HER. First and foremost, we explained the principle of HER mechanisms and the details of the intermediate steps and pathways of the HER in different pH solutions. Also, we summarized the reported TMD composites with their morphology, synthesis method, and electrochemical performance. The specific dimensions of TMDs were then included, and their morphologies were evaluated for their effect on the catalytic efficiency of the HER. We also mention the advancements made in tuning these morphologies for higher performance, providing insight into research opportunities in HER applications.

2. HER Mechanism

The HER is one of the basic electrochemical reactions in water splitting, where the proton reduces to produce hydrogen gas at the cathode. Two different mechanisms are mainly involved in hydrogen evolution, namely the Volmer–Heyrovsky and Volmer–Tafel mechanisms [37]. These mechanisms are influenced by several factors, such as electrode materials, electrolyte pH, electrode surface, the concentration of the reactants, and overpotentials [8,38]. Based on the electrolyte pH conditions (acidic and alkaline), the HER involves different intermediate steps and pathways:
In acidic conditions,
H3O+ + e → H* + H2O         (via Volmer reaction)
H* + H3O+ + e → H2 + H2O       (via Heyrovsky reaction)
Or
H* + H* → H2               (via Tafel reaction)
In alkaline conditions,
H2O + e → H* + OH         (via Volmer reaction)
where H3O+ is a hydronium ion and H* is a hydrogen intermediate [39]. In alkaline conditions, the Heyrovsky reaction and the Tafel reaction follow the same path as in acidic conditions. H3O+ from the acidic electrolyte solution reduces in the Volmer reaction and the hydrogen atom absorbs on the surface of the electrode. In contrast, water molecules from an alkaline electrolyte solution absorb electrons and simultaneously adhere the hydrogen on the surface of the electrode and release OH molecules in the solution. In an alkaline solution, hydrogen evolution is slower because it requires water dissociation to occur before H* creation [40]. Hence, overpotential is required for each of those reactions to overcome the inherent barrier at the cathode [7].
In general, TMDs follow Volmer–Heyrovsky mechanisms because hydrogen desorption is easier in the Heyrovsky reaction compared to the Tafel reaction. Nevertheless, some differences are exhibited in the mechanisms due to the dependency of the transition metal and sulfide vacancies, their active sites, electronic structure, and hydrogen binding energy [41,42]. Researchers intend to add dopants or other materials, boost H2 adsorption rates, and decrease desorption rates on the TMD surface to give the best pathway for the HER.

3. Crystal Structure and Conductivity of TMDs

Electrochemical water splitting has gained a lot of attention with regard to producing clean hydrogen energy. However, concerns were raised about the HER regarding its electronic and catalytic properties, such as low conductivity, low electrochemical stability, and slower reaction kinetics due to the simultaneous participation of four electrons. To overcome these concerns, TMDs are extensively used for the HER because of their interesting mechanism and exceptional properties. However, TMDs have a significant crystal structure, which enhances the conductivity and catalytic activity in the HER.
The crystal structure plays a vital role in determining the conductivity and the catalytic behavior of TMDs for the HER. For example, MoS2 is a quasi-two-dimensional layered TMD, which consists of several layers of S−Mo−S coupled with weak van der Waals interactions [43]. In MoS2, molybdenum (Mo) atoms are sandwiched between two sulfur (S) atoms with strong covalent bonds between Mo and S atoms, as illustrated in Figure 1, where X represents the Mo atom and M represents the S atoms [42,44,45]. It has two polymorphs, namely 2H-MoS2 (hexagonal) and 1T-MoS2 (octahedral). The 2H phase of MoS2 is a semiconductor material with a band gap of about 1.2–1.8 eV, while the 1T phase of MoS2 is metallic in nature [46,47]. In the 1T phase of MoS2, the (002) plane enhances the conductivity of the material, which is characterized by efficient charge transfer behavior. Additionally, the 2D structure of the materials has a high density of catalytic active sites that minimize ion diffusion distance and increase surface area for electrocatalytic processes [48,49]. In addition, more active edge sites were revealed by adding defects to 2H MoS2 nanosheets, which significantly offered structural and electrical advantages for HER activity. Notably, the conversion of the 2H to the 1T phase influences the performance of the HER [1,36,50].
CoS2 is favorable for the advancement of the HER based on both computational and experimental studies [8,51]. CoS2 is a metallic compound with a pyrite structure, where Co is face-centered cubic and S atoms are S-S dimers located at the edge of the unit cell. This material is relatively stable due to the overlapping Co d-orbitals with the S p-orbitals. Moreover, the CoS2 nanoparticles improve the rate of the HER because they facilitate minimal interference in electron transport during the catalytic processes. Consequently, the conductivity is extended by adding defects, impurities, and exact crystal orientation of the TMDs because these factors increase the electronic and catalytic properties of the HER [52,53,54].
The top-down and bottom-up methods are the two basic approaches to synthesizing TMD nanomaterials. Various synthesis methods, such as colloidal synthesis, chemical vapor deposition, pyrolysis, thermal decomposition, solid-state synthesis, liquid phase synthesis, coprecipitation, electrospinning, and hydrothermal and solvothermal methods, are used to synthesize TMDs in the 0D–3D range [55,56,57,58,59,60,61,62,63,64,65]. Among them, the hydrothermal method is frequently utilized to produce all the dimensional TMDs with fine, elaborate, and often specific architectures due to the parameters of the reaction, the interaction between precursors, and the solvent effects. Several factors, such as precursor concentration, microwave power, the exfoliation agent, pH, temperature, pressure, and duration, affect the morphology of materials throughout each step of the synthesis process that converts the material from 0D to 3D. Table 1 presents the various TMD materials, morphologies, synthesis methods, overpotentials, Tafel slopes, and electrolytes.

3.1. Zero-Dimensional (0D) TMDs

Zero-dimensional materials, particularly nanoparticles and nanospheres, showed significant improvement in HER performance compared to bulk materials [60,83]. They have large active sites, a huge surface area/volume ratio, and quantum confinement, which may alter the electronic structure of the catalyst to make it more catalytically favorable. Due to the lower particle sizes, the materials help to improve electron transfer during hydrogen evolution. Furthermore, a large number of edge sites are available in nanoparticles to improve proton adsorption and desorption in the HER [84].
MoS2 and WS2 nanoparticles are widely used to improve the catalytic activity of the HER due to the existence of unique phases (1T and 2H phases). However, the 1T phase is superior to the 2H phase by virtue of its excellent electrical conductivity, availability of more active sites, favorable hydrogen binding energy, larger surface area, and favorable charge transfer characteristics. For instance, Liu et al. described the fact that 1T/1T’ phase-dominated WS2 has a higher HER activity than the 2H phase. They used the colloidal synthesis method to synthesize controllable 1T’ phase-dominated WS2 nanoparticles. The synthesized WS2 nanoparticles exhibited superior catalytic activity with an overpotential of 200 mV at the 10 mA cm−2 current density, a small Tafel slope of 50.4 mV dec−1, and excellent stability [64]. Wang et al. prepared MoS2 nanoparticles on the 3D CFP substrate to improve the HER activity. MoS2 converted its phase from 2H to 1T through lithiation and layer expansion by electrochemical intercalation. This conversion provided more active sites and enhanced the activity of the catalyst by reducing the overpotential to 200 mV (200 mA cm−2) and providing long-term stability [66].
MXene serves as a unique 2D layered structure of transition metal substrate, which has weak van der Waals interaction with excellent electric conductivity. Wang et al. created MoS2 nanodots anchored on MXene as an electrocatalyst for the HER. Polydopamine (PDA)-coated Ti3C2Tx MXene flakes and chelated Mo ions equally dispersed MoS2 nanodots on the MXene surface. During the annealing process, the PDA encapsulated on the MXene surface may assist prevent surface oxidation while maintaining good electrical conductivity (Figure 2). MoS2/MXene demonstrated outstanding catalytic HER performance due to the high exposure of active sites. The composite required a low overpotential of 94 mV at a current density of 10 mA cm−2 and a Tafel slope of 59 mV dec−1 in 1.0 M KOH [85]. Chen et al. prepared a high percentage of a 1T phase of MoS2 quantum dots on a Ti3C2Tx MXene substrate for the HER. The MoS2 QDs/Ti3C2Tx nanohybrid composite facilitated the charge transport process during hydrogen evolution because it provided more active sites and superior electronic conductivity. Thus, the composite showed an overpotential of 220 mV at 10 mA cm−2 and a Tafel slope of 72 mV dec−1 [86].
Researchers are widely using CNTs to create composites with TMDs, increasing the mechanical strength, modifying the electronic structure, and enhancing the conductivity of TMDs for hydrogen evolution. For instance, Lin et al. synthesized a novel CoS2 nanoparticle embedded in an N-doped CNT hollow polyhedron (CoS2-NCNHP) for both hydrogen evolution and Al-S batteries. The CNT acted as a substrate to improve the conductivity and stability of CoS2, while the hollow polyhedron structure offered a larger active surface. CoS2-NCNHP showed superior HER electrocatalysis in both acid and alkaline electrolytes due to the strong synergic effect between the active CoS2 nanoparticles and the N-doped CNT hollow polyhedron. It required an overpotential of 158 and 192 mV at 10 mA cm −2 and a Tafel slope of 84 and 95 mV dec−1 in 0.5 M H2SO4 and 1 M KOH, respectively [87]. Ji et al. prepared NiS2 nanoparticles anchored on the surface of nitrogen-doped CNT (Co-CNTs) encapsulated with cobalt nanoparticles for both supercapacitor and water splitting applications. Co-CNTs provided a larger surface area and better conductivity, which facilitated the charge transport kinetics of composite structures. The composite exhibited overpotentials of 235 and 309 mV in the HER and OER to drive a current density of 10 mA cm−2 [88].
Reduced graphene oxide (r-GO) itself shows good catalytic activity towards the HER. So, researchers make r-GO-TMD composites to increase the two-fold efficiency of the catalyst. The r-GO acts as a good substrate that can accommodate the dispersion of TMDs, reducing their agglomeration, and can offer the largest exposed surface area. It also minimizes the overpotential due to the optimization of the transfer of charges, which is required in the process. For example, Cui et al. fabricated Co-Ni-Fe disulfide nanoparticles on the surface of rGO as an electrocatalyst for the HER. It showed superior catalytic performance with a low overpotential of 138 mV at 10 mA cm−2 and a small Tafel slope of 49 mV dec−1, due to the synergic effect of r-GO and doping with Ni and Fe metals. These doping metals offered numerous active sites, and adding r-GO improved the anti-aggregation and electron transport properties (Figure 3). Additionally, the material had a large electrochemically active surface area due to the modulation of the crystal lattice and interaction of rGO [67]. Salarizadeh et al. prepared MoS2 nanoparticles with coatings on different carbonaceous materials, such as rGO, hollow carbon nanotubes (HCNs), single-walled carbon nanotubes (SWCNTs), and multi-walled carbon nanotubes (MWNTs). In addition, they studied and compared the electrochemical performance of these composites for the HER. Among these composites, MoS2/rGO had a porous nature and possessed many more active sites, good electronic conductivity, and larger catalytic activity than other carbon material-derived MoS2 composites due to the uniformly distributed MoS2 nanoparticles on rGO. It resulted in an outstanding overpotential of 70 mV for the required current density of 10 mA cm−2 and a smaller Tafel slope of 44 mV dec−1 [68]. Sun et al. fabricated MoS2 nanodots supported by rGO for both the HER and sodium-ion battery applications. The size-controlled MoS2 nanodots were synthesized with a narrow size distribution ranging from 2.2 to 5.3 nm using a heating-up method by controlling reaction time. These monodispersed MoS2 nanodots were supported on graphene via an ultrasonication technique and subsequent thermal treatment. The composite displayed a synergic effect due to the high surface area of MoS2 NDs and the electrical conductivity of rGO, resulting in a good catalytic performance with an overpotential of 222 mV at 10 mA−2 and a Tafel slope of 59.8 mV dec−1 [89].

3.2. One-Dimensional (1D) TMDs

One-dimensional morphology, like nanowires, nanotubes, or nanorods, has inherent structural advantages that enhance the performance of the HER. It has a high aspect ratio, which gives a large surface area and a more exposed active site for hydrogen evolution. The elongated straight structure of these materials facilitates easy electron transportation from end to end, thus enhancing catalytic activity. Also, it maintains high catalytic activity for a long time due to the flexible material endurance strain and anti-aggregation properties [90,91,92,93,94].
Xie et al. developed graphdiyne-CoS2 nanowires on carbon cloth as an electrocatalyst for a highly efficient HER. CoS2 nanowires supported on graphdiyne increase the number of active sites, resulting in a low overpotential of 97 mV to obtain a current density of 10 mA cm−2 and a small Tafel slope of 56 mV dec−1. This excellent electrochemical performance is mainly due to the synergic effect of the SP1 hybridized carbon and unsaturated planar sulfurs, which promote electron transfer kinetics (Figure 4) [52]. Gong et al. explained that niobium disulfide (NbS2) nanowires increased the stability and conductivity of the HER through the addition of polypyrrole. The thickness of polypyrrole was controlled by the process of radical polymerization of pyrrole under different reaction durations of 2 min (NbS2-PPy-0), 3 min (NbS2-PPy-1), and 5 min (NbS2-PPy-2). The NbS2-PPy-1 sample showed a higher surface area (65.8 μF cm−2) and good electrochemical performance, with an overpotential of 219 mV at 10 mA cm−2 and a Tafel slope of 56 mV dec−1 compared to NbS2-PPy-0 and NbS2-PPy-2 [65]. Nguyen et al. investigated different compositions of materials, such as WO3 nanorods (WO3 NR), heterostructures of WS2/WO3 nanobricks (WS2/WO3 NB), and WS2/WO3 nanorods (WS2/WO3 NR), using batch reactors for the HER. Among them, the WS2/WO3 NR showed a significant improvement in catalytic activity due to the uniform formation of a heterostructure between the WS2 and WO3 elements in the material. The WS2/WO3 NR composite delivered an overpotential of 224 mV at 10 mA cm−2 with a Tafel slope of 82.7 mV dec−1 and decreased the resistance of 397.7 Ω compared to those of WO3 NR (1816 Ω) and WS2/WO3 NB (3597 Ω) [95]. Abdullah et al. synthesized a NiS2@SnS2 nanocomposite using the hydrothermal method for the HER. The composite exhibited nanoparticle (SnS2) and nanorod (NiS2) structures that showed several benefits, such as greater active surface area, increased edge-terminated structures, and better conductivity. Also, it displayed good catalytic performance with a low overpotential of 155 at 10 mA cm−2 and a very small Tafel slope of 34 mV dec−1 [96].
Hu et al. developed CoS2 nanostructures with different morphologies and stoichiometries using the cobalt nitrate precursor under low sulfur (LS) and high sulfur (HS) concentrations. Moreover, they explained the role of the sulfur concentration in the synthesis of CoS2 and structural analysis. CoS2 (LS) showed straight and smooth nanowires that were transformed into crooked nanowires with a roughened surface, while CoS2 (HS) displayed rougher nanowires that were covered by a large number of dispersed nanoparticles. CoS2 (LS) facilitated the charge and mass transfer processes due to its smaller size with a larger surface area, whereas CoS2 (HS) exhibited better electrocatalytic activity because of its low overpotential of 163 mV at 10 mA cm−2 and small Tafel slope of 70 mV dec−1[54]. Xue et al. designed tubular carbon matrix-supported NiCoP-NiS2 nanowires as a heterogenous electrocatalyst for water splitting applications. The carbon microtube substrates, derived from Metaplexis japonica fluff biomass, enhanced the exposure of more active areas and promoted charge transfer during the reaction. The inner and outer surfaces of the hollow tubular biomass were completely covered by dense and inconsistently orientated nanowires, which provided structural stability and delivered good electrocatalytic performance for both the HER and OER. The composite required overpotentials of 134 and 308 mV at a current density of 10 mA cm−2 for the HER and OER, respectively [97]. Liu et al. developed porous CoS2 nanowires enriched with sulfur vacancies on carbon cloth (Vs-CoS2/CC) via an argon plasma-assisted process for the HER. The introduction of sulfur vacancies into CoS2 played a beneficial role in controlling the electron density surrounding the S sites, effectively activating the inert S sites for the HER and increasing their intrinsic activity. By adjusting the material’s charge density, the Gibbs free energy of hydrogen adsorption (ΔGH*) can also be maximized. Vs-CoS2/CC showed porous nanowires with aggregated nanoparticles, resulting in an enhanced surface area and exposing more active sites. This structural feature exhibited a lower overpotential (170 mV) at a current density of 10 mA cm−2 compared to pure CoS2/CC (236 mV) [98].
The doping of metal is a good strategy to enhance the overall catalytic performance of the material and alter the electronic structure. For example, He et al. synthesized a pyrite-type bimetallic Ni-doped CoS2 nanoneedle on stainless steel to improve the bifunctional electrocatalytic performance, i.e., the mechanical flexibility and structural integrity. Figure 5 illustrates the SEM images of the synthesized Ni-Co precursors. Doping Ni into CoS2 resulted in a decrease in the geometric size of the nanoneedle, which improved both the surface roughness and the specific surface area. This material delivered superior catalytic activity for both the OER (overpotential of 286 mV at 50 mA cm−2 and Tafel slope of 55 mV dec−1) and the HER (overpotential of 350 mV at 30 mA cm−2 and Tafel slope of 76 mV dec−1) [69]. Chen et al. revealed the catalytic activity for the HER using nitrogen anion-modified CoS2 (N−CoS2) nanowires as an electrocatalyst. The incorporation of nitrogen in the CoS2 system altered the morphology and improved the catalytic properties, including the active sites, electronic structure, and reaction dynamics of the material. The porous nature of the N−CoS2 nanowire displayed an overpotential of 152 mV to drive the current density of 50 mA cm−2 and a Tafel slope of 58 mV dec−1 [70]. Nan et al. obtained an efficient P-doped CoS2 hybrid (P-doped CoS2-h) nanoelectrocatalyst for the HER. After sulfurization, CoS2-h displayed a hybrid structure consisting of octahedra and nanotubes, which increased the number of active sites. The synergic effect between the hybrid structure and P-doping enhanced the catalytic performance in both acidic and alkaline media. The P-doped CoS2-h showed an overpotential of 129.4 mV in acidic media and 170 mV in alkaline media at a current density of 10 mA cm−2. The catalyst also exhibited a long stability of 24 h in both acidic and alkaline media [99].

3.3. Two-Dimensional (2D) TMDs

Researchers focus heavily on 2D materials for the HER because of their remarkable features. Two-dimensional materials, especially nanosheets, have an inherent layered structural feature, which can promote the electrocatalytic activity of the HER. Additionally, their large surface area, small bandgap, and rich edge site make the materials excellent HER catalysts [100,101,102,103].
Wang et al. activated the basal plane of multilayered MoS2 via Li-ion intercalation, which resulted in atomic defects and an increased HER. Desulfurization on the MoS2 basal plane would produce nanopores, increasing the number of active sites. Due to the phase transition from 2H to 1T, the MoS2 nanosheets exhibited a low overpotential of 200 mV to drive a current density of 10 mA cm−2 and a small Tafel slope of 65 mV dec−1 [71]. Nguyen et al. created W2C@WS2 nanoflowers using a hydrothermal method with a size of about 200–400 nm for the HER. W2C@WS2 alloys have improved electronic and catalytic properties due to the combination of tungsten carbide and tungsten disulfide [72]. Dinh et al. demonstrated the electrocatalytic performance of Cu-doped NiS2 for both the HER and OER. NiS2 improved the electrical conductivity, surface area, active sites, and electron transport by introducing Cu into NiS2 (Figure 6). The material showed excellent catalytic activity for both the HER (overpotential of 139 mV at 10 mA cm−2) and OER (overpotential of 232 mV at 10 mA cm−2) [73]. S. K. Sumesh developed ZnO-functionalized MoS2 heterostructures for the HER. Due to the incorporation of ZnO, MoS2 increased the surface area and active sites, which showed higher electrochemical performance with an overpotential of 239 mV at 10 mA cm−2, a Tafel slope of 62 mV dec−1, and excellent cycling stability [74]. Do et al. reported the comparative study of different morphologies of WS2 for the HER. WS2 nanoflowers were prepared by using ammonium metatungstate hydrate and thioacetamide precursors, while WS2 nanoflakes were fabricated from ammonium tetrathiotungstate through the thermolysis technique. The WS2 nanoflakes showed a larger surface area compared to the WS2 nanoflakes, which implies higher catalytic activity and faster electrochemical reaction kinetics. This resulted in an overpotential of 325 mV at a current density of 10 mA cm−2 and a small Tafel slope of 92.3 mV dec−1 [104].
Homayounfard et al. synthesized a flower-like MoS2 nanosheet on the heteroatom-doped activated carbon derived from a rose petal biomass for the HER. The F108 surfactant was used as a soft template to control MoS2 morphology with a higher active edge. Also, this surfactant decreased the crystallite and particle size of MoS2 flowers. This resulted in a decrease in the overpotential of 136 mV in 10 mA cm−2 with a Tafel slope of 72 mV dec−1 [105]. Li et al. introduced polyvinylpyrrolidone (PVP) and r-GO into MoS2 dispersion for liquid phase exfoliation to prepare a 3D MoS2/PVP/RGO microstructure on Cu film by inkjet printing. This fabrication exposed high-density edge sites, partial 1T-phase MoS2, reduced charge transfer impedance, and enhanced electron injection. Thus, this novel architecture improved HER performance with a lower overpotential of 51 mV at 10 mA cm−2 and a smaller Tafel slope of 32 mV dec−1 [106]. Wang et al. reported on the tungsten (W) nanoparticle incorporated on the 2H-WS2 film by pulsed laser deposition for the HER. Furthermore, they investigated the effect of laser deposition, which significantly reduced the size of WS2 and the formation of films composed of small nanoparticles and fragments. The catalyst exposed several edge sites with the fragmentation of WS2 because the incorporation of W nanoparticles on the W-WS2 film provided abundant electrons on the S-edge. Thus, W-WS2 film delivered an overpotential of 210 mV at 10 mA cm−2 in 0.5 M H2SO4 compared to the commercial WS2 powder (526 mV) or pure WS2 film (341 mV) [107]. Li et al. investigated the phase engineering of W-doped MoS2 nanosheets as an efficient electrocatalyst for the HER. The high proportion of 1T-phase MoS2 was formed under a low magnetic field due to the synergistic effect of a micro-strain induced by W doping and the introduction of magnetic free energy. Thus, the material displayed an overpotential of 195 mV at a current density of 10 mA cm−2, a Tafel slope of 66 mV dec−1 and long-term stability over 50 h [108].
Platinum is one of the most studied catalysts for the HER because of the high intrinsic activity and the effectiveness of the adsorption and desorption of hydrogen atoms. Decorating TMDs with Pt provides an excellent performance of the Pt and TMD materials while also lowering the cost of the catalyst. For example, Zhu et al. developed a Pt-decorated VS2 nanosheet electrocatalyst for the HER. The Pt catalysts were prepared using a simple optothermal reaction with different structures, including single atoms, clusters, and nanoparticles, which were well controlled and anchored on VS2 nanosheets. This material delivered a low overpotential of 77 mV to obtain a current density of 10 mA cm−2 and a small Tafel slope of 54.27 mV dec−1 [75]. Zainal et al. prepared a ternary molybdenum disulfide-cobalt oxide-platinum (Mo-Co-Pt) composite as an efficient electrocatalyst for the HER. As shown in the TEM images of Figure 7, Pt was decorated on the surface of Co3O4 nanocubes, and a large amount of these Co3O4-Pt nanocubes were grown on the surface of MoS2, resulting in an increased surface area and numerous active sites. This composite exhibited outstanding electrochemical performance, with the lowest overpotential of 20 mV at 10 mA cm−2, the smallest Tafel slope of 34 mV dec−1, and long-term durability over 13 h [76]. Wang et al. reported on Pt nanoparticles anchored onto the VS2 nanosheets via the colloidal method under acidic conditions for the HER. The VS2 nanosheets functioned as stable substrates, providing a larger surface area and high electrical conductivity. This boosted the interfacial electron transfer and electrochemical contact with Pt nanoparticles. Thus, the Pt/VS2 composite delivered the lowest overpotential of 26 mV at a current density of 10 mA cm−2 compared to pure VS2 and long-term stability up to 110 h [109]. ERH.

3.4. Three-Dimensional (3D) TMDs

From recent studies, 3D nanomaterials open new possibilities for designing and attaining the high performance of the catalytic activity of the HER. These materials enhance the catalytic properties by using different strategies, like doping, defect engineering, and the creation of heterojunctions. These strategies optimize the number of available active sites, facilitate electron transfer, and improve mass transport to make 3D nanostructured materials an effective catalyst for HER application [110,111,112,113].
Wu et al. designed a hierarchical MoS2/MoN heterostructure as an efficient electrocatalyst for the HER. As shown in Figure 8, the spherical structure of MoS2 nanosheets with a large number of surface folds exposed more catalytic sites and enabled interfacial mass transfer. The MoS2/MoN heterostructure exhibited excellent electrocatalytic activity due to the synergic effect of its hierarchical architecture and the interfaces between MoS2 and MoN. The material showed overpotentials of 117 and 132 mV to drive a current density of 10 mA cm−2 in 0.5 M H2SO4 and 1 M KOH, respectively [77]. Liu et al. constructed novel WS2 triangular nanoplates on carbon cloth for the HER. For sulfurization and to improve material properties, the sample was calcinated at different temperatures, including 800 °C (WS2TN/CC-800), 900 °C (WS2TN/CC-900), and 1000 °C (WS2TN/CC-1000). Among them, WS2TN/CC-900 exhibited higher surface roughness, more active sites, better kinetics, and good stability. It required overpotentials of 196, 150, and 193 mV at a current density of 10 mA cm−2 in acidic, alkaline, and neutral media, respectively [78]. Huang et al. synthesized phosphorus-doped Co3S4/NiS2 heterostructures embedded in N-doped carbon nanoboxes via a pyrolysis–sulfidation–phosphorization strategy as an electrocatalyst for water splitting. The core–shell structure of the material provided a large inner surface area, which increased the number of active sites and facilitated charge transfer. Due to the synergistic effect between NiS2 and Co3S4, the material showed excellent performance for both the HER (overpotential of 150 mV at 10 mA cm−2 and Tafel slope of 115.91 mV dec−1) and OER (overpotential of 257 mV at 10 mA cm−2 and Tafel slope of 75.8 mV dec−1) [114].
Fang et al. designed Fe-doped CoS2 nanocages with nitrogen-doped carbon wrapping (CN/Fe-CoS2) as an electrocatalyst for overall water splitting. Fe doping in COS2 enhanced the charge transfer, making a reduced band gap, expanded the electrochemical surface area, and regulated the adsorption of intermediates during the reaction. These facilitated the exposure of a more active site and improved the intrinsic catalytic activity of CoS2. In these composites, CoS2 aggregation was prevented by the protective layer of the nanocage architecture and nitrogen-doped carbon wrapping. These results showed superior electrochemical activity and stability towards the HER and OER. The material displayed overpotentials of 186 and 304 mV to drive the current density of 10 mA cm−2 in 1.0 M KOH media for the HER and OER, respectively [115]. Xia et al. developed WS2-decorated Co-N-co-doped carbon hollow nanocages (WS2-3/Co-N-CN) using a one-pot liquid phase mixing and calcination process for the HER. At the higher calcination temperature (750 °C), Co particles precipitated on the surface of the nanocage, improving electron transport efficiency. The microstructure and pore structure of the hybrid materials were influenced by the calcination temperature; thus, WS2-3/Co-N-CN-750 showed better pore volume and average pore size compared to WS2-3/Co-N-CN-650 and WS2-3/Co-N-CN-850. In addition, electrochemical testing was conducted in different electrolytes, such as 0.5 M H2SO4, 0.05 M H2SO4, 1 M KOH, 0.1 M KOH, and 1 M PBS. Among these, WS2-3/Co-N-CN-750 exhibited good electrochemical performance in 0.05 M H2SO4, as proved by an overpotential of 204 mV at a current density of 10 mA cm−2 and a small Tafel slope of 79.1 mV dec−1 [116].
Researchers make composites with carbonaceous materials, including graphene, CNTs, and activated carbon, for the enhancement of catalytic efficiency because these carbonaceous materials possess a high surface area, high electrical conductivity, and stability for the HER. Therefore, introducing the carbonaceous material highlighted that the material resulted in two-fold increased utility for the HER process. He et al. synthesized 3D MoS2 nanosheets coated with rGO on a Mo mesh to improve the Volmer reaction of the HER. Amorphous MoS2 nanosheets exposed more active sites, and the interaction with r-GO enhanced the surface area by the space confinement effect for the stacked layers of graphene. MoS2-rGO@Mo nanohybrids exhibited superior catalytic performance for the HER in 1.0 M H2SO4 with a low overpotential of 123 mV to achieve the current density of 10 mA cm−2, a small Tafel slope of 62 mV dec−1, and excellent stability for more than 15 h [79]. Huang et al. developed a layer-by-layer 3D MoS2-rGO-CNT structure for the HER. CNTs facilitated the non-restacking of the 2D MoS2-rGO nanosheets and the bridge between the fragmented MoS2-rGO nanosheets for electron transport. Thus, the MoS2-rGO-CNT composite showed an overpotential of 400 mV at the high current density of 1200 mA cm−2 with a Tafel slope of 39 mV dec−1 [80]. Liu et al. designed CoS2 nanospheres anchored on MWCNTs using a hydrothermal route for the HER. CoS2 were deposited on the surface of MWCNTs to fix the CoS2 particles and to avoid their excessive growth, which would lead to the increased surface area and the number of electrochemical active sites. This composite enhanced the HER activity in 0.5 M H2SO4 with an overpotential of 257 mV at 10 mA cm−2 and a Tafel slope of 83 mV dec−1 [81]. Haung et al. explained the synergistic effect of proton intercalation and electron transfer via electro-activated molybdenum disulfide/graphite felt (MoS2/GF) for the HER. This synergic effect is mainly due to an active functional group in the GF, the crystalline structure of MoS2 nanoflowers, electron and proton transfer between MoS2 and GF, intrinsic conductivity, and high catalytic activity with good stability (Figure 9). The MoS2/GF material showed the smallest overpotential of 82 mV at 10 mA cm−2 and the smallest Tafel slope of 48 mV dec−1 [82]. Peng et al. prepared highly stable 3D vertically orientated NbS2 nanosheets on the CNT film substrate by the atmospheric pressure chemical vapor deposition (APCVD) route for the HER. In addition, they examined the thickness and nanosheet size, which were controlled by fine-tuning the growth parameters, including temperature, type, and ratio of carrier gases and growth time. In a unique 3D hierarchical structure, CNTs provided high structural stability and conductivity, while 2H NbS2 nanosheets exposed more active sites. The composite exhibited remarkably low overpotentials of ≈55 mV at 10 mA cm−2, a small Tafel slope of 92.1 mV dec−1, and superior stability after 200 h [117].
Table 2 provides recent reports on several TMDs and their composites for the HER. It depicts the overpotential, the Tafel slope, and the various electrolytes employed in the electrochemical reactions. The TMD morphologies varied, and several synthesis methods were discussed. The majority of the papers used a hydrothermal method to synthesize TMD. The investigations commonly utilized 0.5 M H2SO4 and 1.0 M KOH electrolytes for electrochemical reactions. The table reveals that the superior overpotential reported in the Ru MIs-doped MoS2 nanosheet material was 17 mV at a current density of 10 mA cm−2. The MnS2/MnO2-CC material has remarkable kinetics and provides an extremely small Tafel slope of 26.72 mV dec−1 in 0.5 M H2SO4. Pt/ReS2 was synthesized as nanoflowers using microwave-assisted hydrothermal and wet-impregnation techniques, which demonstrate outstanding catalytic performance with a low overpotential of 20 mV at 10 mA cm−2 and a very small Tafel slope of 31.5 mV dec−1 in Ar-saturated 0.1 M HClO4 electrolyte. Some reports describe electrochemical testing in both acidic (0.5 M H2SO4) and alkaline (1.0 M KOH) conditions. MSOR1 (Ru, O-co-doped MoS2) and NiCo2S4/ReS2 materials perform well in alkaline solutions, whereas CoS2@NHCs-800 material shows good performance in acidic solutions. In the investigations, 2D materials, such as nanosheets, demonstrate notable electrochemical performance.

4. Current Challenges and Future Perspectives

Developing TMDs as efficient electrocatalysts for the HER presents challenges. These problems include the synthesis of materials with controlled morphologies; this is critical for enhancing catalytic activity because they have a large surface area where the HER occurs. Noble metals, such as Pt and Ru, decorated the TMDs and showed excellent electrochemical performance towards the HER because of their high intrinsic activity. However, these materials present cost and scaling challenges. Scaling up the catalyst is a major challenge, and most synthesis methods are restricted to the laboratory level due to the scaling up of materials with similarity and reproducibility, and these syntheses modify the electrochemical properties and cost of the materials. Furthermore, intrinsic HER activity in TMDs is constrained since most active sites are associated with the edges, whereas basal planes are normally inactive for the HER. This constraint reduces the overall effectiveness of TMDs and needs a high overpotential for the HER in comparison to noble metals like platinum. While the integration of defects and doping materials in TMDs improves the HER’s electrochemical performance, an improper quantity of these components might degrade the catalyst’s structure and properties. Stability is also important for the HER, but many catalysts deteriorate quickly and are therefore unsuitable for commercial use.
Future advancements will necessitate enhancing active materials and electron transport through the use of hybrid structures, such as core–shell structures, which improve the performance of the HER. They also improve the conductivity, stability, and availability of active sites. The development of ternary and quaternary TMDs is a good alternative for noble metal-based electrocatalysts because they have excellent redox reversibility and high electronic conductivity. Compositing TMDs with carbonaceous materials shows promise for the HER since it improves electrical conductivity and texture. It reveals more active sites because of the tight interaction between the metal ions and carbon base; the resulting metal nanoparticles are smaller and more widely distributed than the bulk metal. The active metal phase and carbon substrates have a synergistic coupling effect that makes charge transfer easier and boosts catalytic activity. Moreover, their scalability and suitability for industrial applications further enhance their potential in the HER. Heteroatom doping and phase engineering are effective strategies for improving electrochemical properties, such as electrical conductivity, charge transfer kinetics, and stability. The electrocatalytic HER activity can be increased through phase engineering since the 1T phase often shows higher HER activity than the 2H phase. We will have to use proper element doping and alloying to change the electrical structure in order to achieve good electrochemical characteristics for HER applications. Moreover, heteroatom doping facilitates and optimizes the chemical-adsorption energies of reactive intermediates during hydrogen evolution. Co-doping, which combines heteroatoms with other elements, is also a viable method for improving material catalysis activity. Another strategy is to develop effective catalysts with specific sites for the adsorption of hydroxide and hydrogen in the HER process. These strategies developed efficient catalysts; future developments will grow high-performing catalytic materials and will provide an intrinsic understanding of the alkaline HER. When employing TMDs in industrial HER applications, we must use clean and sustainable synthetic approaches. However, the cost optimization of TMDs will be exclusively determined by the development of solutions, and their implementation into commercial devices will be vital to the widespread adoption of TMDs in the energy industry. Finally, the development of new TMD materials is needed; this would most likely be guided by increasingly advanced computational approaches. To predict the impacts of doping, defect formation, and nanostructuring on HER activity, we must trust in the previously mentioned approaches, which will aid in understanding and expediting their practical implementation in renewable energy systems.

5. Conclusions

In conclusion, the advancement of sustainable energy is crucial in energy conversion and storage technologies. Developing research works have shown that TMDs have better conductivity, lower band gap energy, and lower overpotential. The crystal structure of TMDs plays an important role in enhancing their catalytic activities with respect to band structure, electron mobility, and electrical conductivity. A discussion of the electrochemical properties of various TMD materials according to their morphology for the HER is provided here. Thus, nanosheet TMDs have been recognized as good 2D nanomaterials for the HER because of their large surface area, higher number of active sites, and exceptional quantum confinement effect. In addition, their higher density of reactive edge sites and their tendency to intercalate between layers raise their activity. The properties of doped or composite materials that change the electrical and structural properties of TMDs are also described in this review. Doping the metals improves the binding energies of intermediates with TMDs and adjusts the electrical structure to obtain quick electron transfer kinetics. In order to reveal more active sites and accelerate electron transmission, the composite material exhibits a synergistic impact between its components as well as the activation of additional interfaces and defects. Finally, several future developments that can contribute to the improvement of the efficiency and life of TMD-based HER catalysts are pointed out.

Author Contributions

Conceptualization, H.M.Y., S.S.J. and H.M.Y. writing—original draft preparation, H.M.Y. and T.S.B. writing—review and editing, S.S.J., V.D.C. and D.-K.K. literature review and editing; H.M.Y. supervision, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Anusandhan National Research Foundation (ANRF) and Science and Engineering Research Board (SERB), Government of India under the grant number RJF/2020/000077. We gratefully acknowledge MDPI for waiving the Article Processing Charge (APC).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shiraz, H.G.; Crispin, X.; Berggren, M. Transition metal sulfides for electrochemical hydrogen evolution. Int. J. Hydrogen Energy 2021, 46, 24060–24077. [Google Scholar] [CrossRef]
  2. Muthu, S.; Balaji, M.; Saravanan, S.; Sekar, S.; Kumar, S.A.; Ilanchezhiyan, P.; Lee, S.; Sridharan, M.B. Hierarchical nanocomposite of cobalt sulfide nanoflakes-decorated reduced graphene oxide for enhancing the electrocatalytic hydrogen evolution reaction. Int. J. Hydrogen Energy 2023, 52, 1405–1414. [Google Scholar] [CrossRef]
  3. Nasser, M.; Megahed, T.F.; Ookawara, S.; Hassan, H. A review of water electrolysis–based systems for hydrogen production using hybrid/solar/wind energy systems. Environ. Sci. Pollut. Res. 2022, 29, 86994–87018. [Google Scholar] [CrossRef] [PubMed]
  4. Karapekmez, A.; Dincer, I. Development of a multigenerational energy system for clean hydrogen generation. J. Clean. Prod. 2021, 299, 126909. [Google Scholar] [CrossRef]
  5. Pareek, A.; Dom, R.; Gupta, J.; Chandran, J.; Adepu, V.; Borse, P.H. Insights into renewable hydrogen energy: Recent advances and prospects. Mater. Sci. Energy Technol. 2020, 3, 319–327. [Google Scholar] [CrossRef]
  6. You, B.; Sun, Y. Innovative Strategies for Electrocatalytic Water Splitting. Acc. Chem. Res. 2018, 51, 1571–1580. [Google Scholar] [CrossRef]
  7. Farhan, A.; Murad, M.; Qayyum, W.; Nawaz, A.; Sajid, M.; Shahid, S.; Qamar, M.A. Transition-metal sulfides with excellent hydrogen and oxygen reactions: A mini-review. J. Solid State Chem. 2024, 329, 124445. [Google Scholar] [CrossRef]
  8. Zahra, R.; Pervaiz, E.; Yang, M.; Rabi, O.; Saleem, Z.; Ali, M.; Farrukh, S. A review on nickel cobalt sulphide and their hybrids: Earth abundant, pH stable electro-catalyst for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 24518–24543. [Google Scholar] [CrossRef]
  9. Guo, C.; Liu, X.; Gao, L.; Ma, X.; Zhao, M.; Zhou, J.; Kuang, X.; Deng, W.; Sun, X.; Wei, Q. Oxygen defect engineering in cobalt iron oxide nanosheets for promoted overall water splitting. J. Mater. Chem. A 2019, 7, 21704–21710. [Google Scholar] [CrossRef]
  10. Chen, P.; Ye, J.; Wang, H.; Ouyang, L.; Zhu, M. Recent progress of transition metal carbides/nitrides for electrocatalytic water splitting. J. Alloys Compd. 2021, 883, 160833. [Google Scholar] [CrossRef]
  11. Babar, P.; Lokhande, A.; Karade, V.; Pawar, B.; Gang, M.G.; Pawar, S.; Kim, J.H.B. ifunctional 2D Electrocatalysts of Transition Metal Hydroxide Nanosheet Arrays for Water Splitting and Urea Electrolysis. ACS Sustain. Chem. Eng. 2019, 7, 10035–10043. [Google Scholar] [CrossRef]
  12. Zhang, H.; Maijenburg, A.W.; Li, X.; Schweizer, S.L.; Wehrspohn, R.B. Bifunctional Heterostructured Transition Metal Phosphides for Efficient Electrochemical Water Splitting. Adv. Funct. Mater. 2020, 30, 2003261. [Google Scholar] [CrossRef]
  13. Irshad, A.; Munichandraiah, N. Electrodeposited Nickel-Cobalt-Sulfide Catalyst for the Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2017, 9, 19746–19755. [Google Scholar] [CrossRef]
  14. Zhang, B.; Li, X.; Liu, M. Effect of nickel sulfide catalysts with different morphologies on high efficiency alkaline hydrogen evolution reaction. Int. J. Electrochem. Sci. 2024, 19, 100579. [Google Scholar] [CrossRef]
  15. Wu, H.; Feng, C.; Zhang, L.; Zhang, J.; Wilkinson, D.P. Non-Noble Metal Electrocatalysts for the Hydrogen Evolution Reaction in Water Electrolysis; Springer: Gateway East, Singapore, 2021; Volume 4. [Google Scholar] [CrossRef]
  16. Dutta, B.; Wu, Y.; Chen, J.; Wang, J.; He, J.; Sharafeldin, M.; Suib, S.L. Partial Surface Selenization of Cobalt Sulfide Microspheres for Enhancing the Hydrogen Evolution Reaction. ACS Catal. 2019, 9, 456–465. [Google Scholar] [CrossRef]
  17. Ding, X.; Liu, D.; Zhao, P.; Chen, X.; Wang, H.; Oropeza, F.E.; Gorni, G.; Barawi, M.; García-Tecedor, M.; de la Peña O’Shea, V.A.; et al. Dynamic restructuring of nickel sulfides for electrocatalytic hydrogen evolution reaction. Nat. Commun. 2024, 15, 5336. [Google Scholar] [CrossRef]
  18. Yang, J.; Mohmad, A.R.; Wang, Y.; Fullon, R.; Song, X.; Zhao, F.; Bozkurt, I.; Augustin, M.; Santos, E.J.G.; Shin, H.S.; et al. Ultrahigh-current-density niobium disulfide catalysts for hydrogen evolution. Nat. Mater. 2019, 18, 1309–1314. [Google Scholar] [CrossRef]
  19. Yu, Q.; Luo, Y.; Qiu, S.; Li, Q.; Cai, Z.; Zhang, Z.; Liu, S.; Sun, C.; Liu, B. Tuning the hydrogen evolution performance of metallic 2d tantalum disulfide by interfacial engineering. ACS Nano 2019, 13, 11874–11881. [Google Scholar] [CrossRef]
  20. Shi, M.; Jiang, Z.; Mei, B.; Li, Y.; Sun, F.; Yu, H.; Xu, Y. Tuning the hydrogen evolution performance of 2D tungsten disulfide by interfacial engineering. J. Mater. Chem. A 2021, 9, 7059–7067. [Google Scholar] [CrossRef]
  21. Pan, H. Tension-Enhanced Hydrogen Evolution Reaction on Vanadium Disulfide Monolayer. Nanoscale Res. Lett. 2016, 11, 113. [Google Scholar] [CrossRef]
  22. Liu, X.; Hu, J.; Yue, C.; Della Fera, N.; Ling, Y.; Mao, Z.; Wei, J. High performance field-effect transistor based on multilayer tungsten disulfide. ACS Nano 2014, 8, 10396–10402. [Google Scholar] [CrossRef] [PubMed]
  23. Li, M.; Zhou, Q.; Cao, M.; Zhou, Z.; Liu, X. High-temperature solid lubrication applications of Transition Metal Dichalcogenides(TMDCs) MX2: A review. Nano Mater. Sci. 2024. [Google Scholar] [CrossRef]
  24. Meerbach, C.; Klemmed, B.; Spittel, D.; Bauer, C.; Park, Y.J.; Hübner, R.; Jeong, H.Y.; Erb, D.; Shin, H.S.; Lesnyak, V.; et al. General Colloidal Synthesis of Transition-Metal Disulfide Nanomaterials as Electrocatalysts for Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 13148–13155. [Google Scholar] [CrossRef]
  25. Guo, H.; Zhang, Z.; Huang, B.; Wang, X.; Niu, H.; Guo, Y.; Li, B.; Zheng, R.; Wu, H. Theoretical study on the photocatalytic properties of 2D InX(X = S, Se)/transition metal disulfide (MoS2 and WS2) van der Waals heterostructures. Nanoscale 2020, 12, 20025–20032. [Google Scholar] [CrossRef]
  26. Zhao, M.; Yang, M.; Huang, W.; Liao, W.; Bian, H.; Chen, D.; Wang, L.; Tang, J.; Liu, C. Synergism on Electronic Structures and Active Edges of Metallic Vanadium Disulfide Nanosheets via Co Doping for Efficient Hydrogen Evolution Reaction in Seawater. ChemCatChem 2021, 13, 2138–2144. [Google Scholar] [CrossRef]
  27. Huang, C.; Yu, L.; Zhang, W.; Xiao, Q.; Zhou, J.; Zhang, Y.; An, P.; Zhang, J.; Yu, Y. N-doped Ni-Mo based sulfides for high-efficiency and stable hydrogen evolution reaction. Appl. Catal. B 2020, 276, 119137. [Google Scholar] [CrossRef]
  28. Bhimanapati, G.R.; Lin, Z.; Meunier, V.; Jung, Y.; Cha, J.; Das, S.; Xiao, D.; Son, Y.; Strano, M.S.; Cooper, V.R.; et al. Recent Advances in Two-Dimensional Materials beyond Graphene. ACS Nano 2015, 9, 11509–11539. [Google Scholar] [CrossRef]
  29. Tekalgne, M.A.; Nguyen, K.V.; Nguyen, D.L.T.; Nguyen, V.H.; Nguyen, T.P.; Vo, D.V.N.; Trinh, T.P.; Vo, N.D.; Trinh, Q.T.; Hasani, A.; et al. Hierarchical molybdenum disulfide on carbon nanotube–reduced graphene oxide composite paper as efficient catalysts for hydrogen evolution reaction. J. Alloys Compd. 2020, 823, 153897. [Google Scholar] [CrossRef]
  30. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 nanoparticles grown on graphene: An advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef]
  31. Chen, R.; Song, Y.; Wang, Z.; Gao, Y.; Sheng, Y.; Shu, Z.; Zhang, J.; Li, X. Porous nickel disulfide/reduced graphene oxide nanohybrids with improved electrocatalytic performance for hydrogen evolution. Catal. Commun. 2016, 85, 26–29. [Google Scholar] [CrossRef]
  32. Li, Y.; Yin, X.; Huang, X.; Liu, X.; Wu, W. Efficient and scalable preparation of MoS2 nanosheet/carbon nanotube composites for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 16489–16499. [Google Scholar] [CrossRef]
  33. Tiwari, J.N.; Tiwari, R.N.; Kim, K.S. Zero-dimensional, one-dimensional, two-dimensional and three-dimensional nanostructured materials for advanced electrochemical energy devices. Prog. Mater. Sci. 2012, 57, 724–803. [Google Scholar] [CrossRef]
  34. Long, X.; Li, G.; Wang, Z.; Zhu, H.; Zhang, T.; Xiao, S.; Gua, W.; Yang, S. Metallic Iron-Nickel Sulfide Ultrathin Nanosheets As a Highly Active Electrocatalyst for Hydrogen Evolution Reaction in Acidic Media. J. Am. Chem. Soc. 2015, 137, 11900–11903. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, H.; Si, J.; Lyu, S.; Zhang, T.; Li, Z.; Lei, C.; Lei, L.; Yuan, C.; Gao, L.; Hou, Y. Highly Effective Electrochemical Exfoliation of Ultrathin Tantalum Disulfide Nanosheets for Energy-Efficient Hydrogen Evolution Electrocatalysis. ACS Appl. Mater. Interfaces 2020, 12, 24675–24682. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, X.; Jia, F.; Song, S. Recent advances in structural engineering of molybdenum disulfide for electrocatalytic hydrogen evolution reaction. Chem. Eng. J. 2021, 405, 127013. [Google Scholar] [CrossRef]
  37. Lasia, A. Mechanism and kinetics of the hydrogen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 19484–19518. [Google Scholar] [CrossRef]
  38. Shajaripour, J.S.Y.; Ghaffarinejad, A.; Khajehsaeidi, Z. The effect of annealing temperature, reaction time, and cobalt precursor on the structural properties and catalytic performance of CoS2 for hydrogen evolution reaction. Int. J. Hydrogen Energy 2021, 46, 3922–3932. [Google Scholar] [CrossRef]
  39. Wang, M.; Zhang, L.; He, Y.; Zhu, H. Recent advances in transition-metal-sulfide-based bifunctional electrocatalysts for overall water splitting. J. Mater. Chem. A 2021, 9, 5320–5363. [Google Scholar] [CrossRef]
  40. Safizadeh, F.; Ghali, E.; Houlachi, G. Electrocatalysis developments for hydrogen evolution reaction in alkaline solutions—A Review. Int. J. Hydrogen Energy 2015, 40, 256–274. [Google Scholar] [CrossRef]
  41. Wang, J.; Liu, J.; Zhang, B.; Ji, X.; Xu, K.; Chen, C.; Miao, L.; Jiang, J. The mechanism of hydrogen adsorption on transition metal dichalcogenides as hydrogen evolution reaction catalyst. Phys. Chem. Chem. Phys. 2017, 19, 10125–10132. [Google Scholar] [CrossRef]
  42. Mondal, A.; Vomiero, A. 2D Transition Metal Dichalcogenides-Based Electrocatalysts for Hydrogen Evolution Reaction. Adv. Funct. Mater. 2022, 32, 2208994. [Google Scholar] [CrossRef]
  43. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.; Asefa, T.; Chhowalla, M. Conducting MoS2 nanosheets as catalysts for hydrogen evolution reaction. Nano Lett. 2013, 13, 6222–6227. [Google Scholar] [CrossRef] [PubMed]
  44. Ran, N.; Qiu, W.; Song, E.; Wang, Y.; Zhao, X.; Liu, Z.; Liu, J. Bond Electronegativity as Hydrogen Evolution Reaction Catalyst Descriptor for Transition Metal (TM = Mo, W) Dichalcogenides. Chem. Mater. 2020, 32, 1224–1234. [Google Scholar] [CrossRef]
  45. Yu, H.; Zhang, M.; Cai, Y.; Zhuang, Y.; Wang, L. The Advanced Progress of MoS2 and WS2 for Multi-Catalytic Hydrogen Evolution Reaction Systems. Catalysts 2023, 13, 1148. [Google Scholar] [CrossRef]
  46. Yin, Y.; Han, J.; Zhang, Y.; Zhang, X.; Xu, P.; Yuan, Q.; Samad, L.; Wang, X.; Wang, Y.; Zang, Z.; et al. Contributions of Phase, Sulfur Vacancies, and Edges to the Hydrogen Evolution Reaction Catalytic Activity of Porous Molybdenum Disulfide Nanosheets. J. Am. Chem. Soc. 2016, 138, 7965–7972. [Google Scholar] [CrossRef]
  47. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  48. Wang, R.; Shao, Q.; Yuan, Q.; Sun, P.; Nie, R.; Wang, X. Direct growth of high-content 1T phase MoS2 film by pulsed laser deposition for hydrogen evolution reaction. Appl. Surf. Sci. 2020, 504, 144320. [Google Scholar] [CrossRef]
  49. Li, Z.; Yue, Y.; Peng, J.; Luo, Z. Phase engineering two-dimensional nanostructures for electrocatalytic hydrogen evolution reaction. Chin. Chem. Lett. 2023, 34, 107119. [Google Scholar] [CrossRef]
  50. Kashif, M.; Thangarasu, S.; Oh, T.H. Enriching the active sites of nanosheets assembled as flower-like MoS2 microstructure by controllable morphology for electrochemical hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 82, 47–52. [Google Scholar] [CrossRef]
  51. Chen, X.; Yu, Y.; Li, J.; Deng, P.; Wang, C.; Hua, Y.; Shen, Y.; Tian, X. Recent advances in cobalt disulfide for electrochemical hydrogen evolution reaction. Int. J. Hydrogen Energy 2023, 48, 9231–9243. [Google Scholar] [CrossRef]
  52. Xie, W.; Liu, K.; Shi, G.; Fu, X.; Chen, X.; Fan, Z.; Liu, M.; Yuan, M.; Wang, M. CoS2 nanowires supported graphdiyne for highly efficient hydrogen evolution reaction. J. Energy Chem. 2021, 60, 272–278. [Google Scholar] [CrossRef]
  53. Zhang, J.; Liu, Y.; Sun, C.; Xi, P.; Peng, S.; Gao, D.; Xue, D. Accelerated Hydrogen Evolution Reaction in CoS2 by Transition-Metal Doping. ACS Energy Lett. 2018, 3, 779–786. [Google Scholar] [CrossRef]
  54. Hu, Y.; Qi, Y.; Pi, M.; Qiao, Y.; Hao, L.; Wang, Y.; Guan, H.; Feng, J. Nanostructure and stoichiometry tailoring of CoS2 for high performance hydrogen evolution reaction. Int. J. Hydrogen Energy 2023, 48, 16279–16285. [Google Scholar] [CrossRef]
  55. Hedayati, M.A.; Ahangar, A.M.; Maleki, M.; Ghanbari, H. Low-temperature hydrothermal growth of MoS2 nanostructures on carbon foam for hydrogen evolution reaction. Diam. Relat. Mater. 2023, 139, 110342. [Google Scholar] [CrossRef]
  56. Mahanthappa, M.; Bhat, M.P.; Alshoaibi, A. Exploring the role of redox-active species on NiS-NiS2 incorporated sulfur-doped graphitic carbon nitride nanohybrid as a bifunctional electrocatalyst for overall water splitting. Int. J. Hydrogen Energy 2024, 84, 641–649. [Google Scholar] [CrossRef]
  57. Yue, H.; Guo, Z.; Fan, J.; Wang, P.; Zhen, S.; Yuan, W. Cu intercalation and defect engineering realize an atomic-scale hydrogen spillover effect in NbS2 to boost acidic hydrogen evolution. Inorg. Chem. Front. 2024, 11, 1899–1911. [Google Scholar] [CrossRef]
  58. Shen, W.; Qiao, L.; Ding, J.; Sui, Y. Constructing 1 T-2 H TaS2 nanosheets with architecture and defect engineering for enhance hydrogen evolution reaction. J. Alloys Compd. 2023, 935, 167877. [Google Scholar] [CrossRef]
  59. Hanan, A.; Lakhan, M.N.; Solangi, M.Y.; AlSalhi, M.S.; Kumar, V.; Devanesan, S.; Shar, M.A.; Abro, M.I.; Aftab, U. MXene based electrocatalyst: CoS2@Ti3C2Tx composite for hydrogen evolution reaction in alkaline media. Mater. Today Sustain. 2023, 24, 100585. [Google Scholar] [CrossRef]
  60. Cui, W.; Sun, X.; Xu, S.; Li, C.; Bai, J. CoS2-CoSe2 hybrid nanoparticles grown on carbon nanofibers as electrode for supercapacitor and hydrogen evolution reaction. J. Alloys Compd. 2024, 990, 174366. [Google Scholar] [CrossRef]
  61. Van Nguyen, T.; Nguyen, T.P.; Van Le, Q.; Van Dao, D.; Ahn, S.H.; Kim, S.Y. Facile route for synthesizing WS2/W2C nano hollow flowers and their application for hydrogen evolution reaction. J. Alloys Compd. 2023, 955, 170231. [Google Scholar] [CrossRef]
  62. Sun, F.; Tian, X.; Zang, J.; Zhu, R.; Hou, Z.; Zheng, Y.; Wang, Y.; Dong, L. Constructing composite active centers optimized with Cr-doped NiS/NiS2 heterostructure for efficiently catalyzing alkaline hydrogen evolution reaction. Fuel 2024, 363, 130999. [Google Scholar] [CrossRef]
  63. Fu, Y.G.; Liu, H.Q.; Liu, C.; Lü, Q.F. Ultralight porous carbon loaded Co-doped MoS2 as an efficient electrocatalyst for hydrogen evolution reaction in acidic and alkaline media. J. Alloys Compd. 2023, 967, 171748. [Google Scholar] [CrossRef]
  64. Liu, Z.; Li, N.; Su, C.; Zhao, H.; Xu, L.; Yin, Z.; Li, J.; Du, Y. Colloidal synthesis of 1T’ phase dominated WS2 towards endurable electrocatalysis. Nano Energy 2018, 50, 176–181. [Google Scholar] [CrossRef]
  65. Gong, X.; Kong, D.; Pam, M.E.; Guo, L.; Fan, S.; Huang, S.; Wang, Y.; Gao, Y.; Shi, Y.; Yang, H.Y. Polypyrrole coated niobium disulfide nanowires as high performance electrocatalysts for hydrogen evolution reaction. Nanotechnology 2019, 30, 405601. [Google Scholar] [CrossRef]
  66. Wang, H.; Lu, Z.; Kong, D.; Sun, J.; Hymel, T.M.; Cui, Y. Electrochemical tuning of MoS2 nanoparticles on three-dimensional substrate for efficient hydrogen evolution. ACS Nano 2014, 8, 4940–4947. [Google Scholar] [CrossRef]
  67. Cui, Y.; Zhang, R.; Zhang, J.; Wang, Z.; Xue, H.; Mao, W.; Huang, W. Highly active and stable electrocatalytic hydrogen evolution catalyzed by nickel, iron doped cobalt disulfide@reduced graphene oxide nanohybrid electrocatalysts. Mater. Today Energy 2018, 7, 44–50. [Google Scholar] [CrossRef]
  68. Salarizadeh, P.; Askari, M.B.; Seifi, M.; Rozati, S.M. MoS2 coating on different carbonaceous materials: Comparison of electrochemical properties and hydrogen evolution reaction performance. J. Electroanal. Chem. 2019, 847, 113198. [Google Scholar] [CrossRef]
  69. He, G.; Zhang, W.; Deng, Y.; Zhong, C.; Hu, W.; Han, X. Engineering pyrite-type bimetallic Ni-doped CoS2 nanoneedle arrays over awide compositional range for enhanced oxygen and hydrogen electrocatalysis with flexible property. Catalysts 2017, 7, 366. [Google Scholar] [CrossRef]
  70. Chen, P.; Zhou, T.; Chen, M.; Tong, Y.; Zhang, N.; Peng, X.; Chu, W.; Wu, X.; Wu, C.; Xie, Y. Enhanced Catalytic Activity in Nitrogen-Anion Modified Metallic Cobalt Disulfide Porous Nanowire Arrays for Hydrogen Evolution. ACS Catal. 2017, 7, 7405–7411. [Google Scholar] [CrossRef]
  71. Wang, C.; Lu, H.; Tang, K.; Mao, Z.; Li, Q.; Wang, X.; Yan, C. Atom removal on the basal plane of layered MoS2 leading to extraordinarily enhanced electrocatalytic performance. Electrochim. Acta 2020, 336, 135740. [Google Scholar] [CrossRef]
  72. Nguyen, T.P.; Kim, S.Y.; Lee, T.H.; Jang, H.W.; Le, Q.; Van Kim, I.T. Facile synthesis of W2C@WS2 alloy nanoflowers and their hydrogen generation performance. Appl. Surf. Sci. 2020, 504, 144389. [Google Scholar] [CrossRef]
  73. Dinh, K.N.; Sun, Y.; Pei, Z.; Yuan, Z.; Suwardi, A.; Huang, Q.; Yan, Q. Electronic Modulation of Nickel Disulfide toward Efficient Water Electrolysis. Small 2020, 16, 1905885. [Google Scholar] [CrossRef]
  74. Sumesh, C.K. Zinc oxide functionalized molybdenum disulfide heterostructures as efficient electrocatalysts for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 619–628. [Google Scholar] [CrossRef]
  75. Zhu, J.; Cai, L.; Yin, X.; Wang, Z.; Zhang, L.; Ma, H.; Ke, Y.; Du, Y.; Xi, S.; Wee, A.T.S.; et al. Enhanced Electrocatalytic Hydrogen Evolution Activity in Single-Atom Pt-Decorated VS2 Nanosheets. ACS Nano 2020, 14, 5600–5608. [Google Scholar] [CrossRef]
  76. Zainal, S.N.; Sookhakian, M.; Woi, P.M.; Alias, Y. Ternary molybdenum disulfide nanosheets-cobalt oxide nanocubes-platinum composite as efficient electrocatalyst for hydrogen evolution reaction. Electrochim. Acta 2020, 345, 136255. [Google Scholar] [CrossRef]
  77. Wu, A.; Gu, Y.; Xie, Y.; Yan, H.; Jiao, Y.; Wang, D.; Tian, C. Interfacial engineering of MoS2/MoN heterostructures as efficient electrocatalyst for pH-universal hydrogen evolution reaction. J. Alloys Compd. 2021, 867, 159066. [Google Scholar] [CrossRef]
  78. Liu, M.; Geng, A.; Yan, J. Construction of WS2 triangular nanoplates array for hydrogen evolution reaction over a wide pH range. Int. J. Hydrogen Energy 2020, 45, 2909–2916. [Google Scholar] [CrossRef]
  79. He, B.; Chen, L.; Jing, M.; Zhou, M.; Hou, Z.; Chen, X. 3D MoS2-rGO@Mo nanohybrids for enhanced hydrogen evolution: The importance of the synergy on the Volmer reaction. Electrochim. Acta 2018, 283, 357–365. [Google Scholar] [CrossRef]
  80. Huang, J.; Chen, M.; Li, X.; Zhang, X.; Lin, L.; Liu, W.; Liu, Y. A facile layer-by-layer fabrication of three dimensional MoS2-rGO-CNTs with high performance for hydrogen evolution reaction. Electrochim. Acta 2019, 300, 235–241. [Google Scholar] [CrossRef]
  81. Liu, H.; He, P.; Jia, L.; He, M.; Zhang, X.; Wang, S.; Zhang, X.; Li, C.; Zhang, Y.; Dong, F. Cobalt disulfide nanosphere dispersed on multi-walled carbon nanotubes: An efficient and stable electrocatalyst for hydrogen evolution reaction. Ionics 2018, 24, 3591–3599. [Google Scholar] [CrossRef]
  82. Huang, D.; Lei, L.; Deng, R.; Chen, S.; Li, Z.; Zhang, C.; Wang, W.; Wang, K. The synergistic effect of proton intercalation and electron transfer via electro-activated molybdenum disulfide/graphite felt toward hydrogen evolution reaction. J. Catal. 2020, 381, 175–185. [Google Scholar] [CrossRef]
  83. Adam, A.; Adamu, H.; AlMansour, S.M.; Turkestani, F.A.; Obot, I.; Qamar, M. Comparative electrocatalytic behavior of mesoporous FeP and FeS2 for the hydrogen evolution reaction. Mater. Lett. 2024, 377, 137312. [Google Scholar] [CrossRef]
  84. Jin, H.; Sun, Y.; Sun, Z.; Yang, M.; Gui, R. Zero-dimensional sulfur nanomaterials: Synthesis, modifications and applications. Coord. Chem. Rev. 2021, 438, 213913. [Google Scholar] [CrossRef]
  85. Wang, X.; You, W.; Yang, L.; Chen, G.; Wu, Z.; Zhang, C.; Chen, Q.; Che, R. Enhanced electrocatalytic hydrogen evolution by molybdenum disulfide nanodots anchored on MXene under alkaline conditions. Nanoscale Adv. 2022, 4, 3398–33406. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, L.; Liang, J.; Zhang, Q.; Hu, X.; Peng, W.; Li, Y.; Zhang, F.; Fan, X. Quasi zero-dimensional MoS2 quantum dots decorated 2D Ti3C2Tx MXene as advanced electrocatalysts for hydrogen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 10583–10593. [Google Scholar] [CrossRef]
  87. Lin, Y.; Cui, X.; Liu, P.; Wang, Z.; Zhang, X.; Yang, X.; Feng, C.; Pan, Y. CoS2 Nanoparticles Embedded N-doped Carbon Nanotubes Hollow Polyhedron for Hydrogen Evolution and Al−S Batteries. ChemCatChem 2024, 202401155. [Google Scholar] [CrossRef]
  88. Ji, X.; Xu, B.; Zhang, H.; Zhang, X.; Yang, P. NiS2 nanoparticles anchored on Co-carbon nanotubes for supercapacitor and overall water splitting. J. Alloys Compd. 2023, 968, 172192. [Google Scholar] [CrossRef]
  89. Sun, W.; Li, P.; Liu, X.; Shi, J.; Sun, H.; Tao, Z.; Li, F.; Chen, J. Size-controlled MoS2 nanodots supported on reduced graphene oxide for hydrogen evolution reaction and sodium-ion batteries. Nano Res. 2017, 10, 2210–2222. [Google Scholar] [CrossRef]
  90. Li, J.; Zheng, G. One-dimensional earth-abundant nanomaterials for water-splitting electrocatalysts. Adv. Sci. 2017, 4, 1600380. [Google Scholar] [CrossRef]
  91. Li, P.; Chen, W. Recent advances in one-dimensional nanostructures for energy electrocatalysis. Chin. J. Catal. 2019, 40, 4–22. [Google Scholar] [CrossRef]
  92. Yang, Y.; Li, C.; Ma, X.; Li, S.; Li, X.; Du, X.; Hao, X.; Abudula, A.; Guan, G. One-dimensional CoMoS4 nanorod arrays as an efficient electrocatalyst for hydrogen evolution reaction. J. Alloys Compd. 2020, 821, 153245. [Google Scholar] [CrossRef]
  93. Zhai, S.; Feng, Y.; Yuan, Z.Y. Engineering one-dimensional transition metal compound nanostructures for electrocatalytic hydrogen evolution: Overview and perspectives. Int. J. Hydrogen Energy 2023, 48, 34677–34699. [Google Scholar] [CrossRef]
  94. Dutta, A.; Krishnappa, M.; Porat, H.; Lavi, R.; Lal, A.; Yadav, M.K.; Mandic, V.; Makrinich, G.; Lakhiman, A.; Zak, A.; et al. Plasma-treated 1D transition metal dichalcogenides for efficient electrocatalytic hydrogen evolution reaction. J. Mater. Chem. A 2024, 12, 25176–25185. [Google Scholar] [CrossRef]
  95. Van Nguyen, T.; Huynh, K.A.; Le, Q.V.; Ahn, S.H.; Kim, S.Y. Facile Synthesis of WS2/WO3Materials in a Batch Reactor for the Hydrogen Evolution Reaction. Int. J. Energy Res. 2024, 2024, 4878549. [Google Scholar] [CrossRef]
  96. Abdullah, M.; Alyousef, H.A.; Alotaibi, B.M.; Alrowaily, A.W.; Al-Harbi, N.; Henaish, A.M.A.; Dahshan, A. Hydrothermal synthesis of NiS2@SnS2 nanohybrid for the water splitting application. Mater. Chem. Phys. 2024, 319, 129361. [Google Scholar] [CrossRef]
  97. Xue, Y.; Wang, H.; Zhang, X.; Zhao, C.; Yao, J.; Yan, Q.; Zhan, K.; Cao, D.; Wang, G. Construction of tubular carbon matrix-supported NiCoP-NiS2 nanowires with heterointerfaces for overall water splitting. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130516. [Google Scholar] [CrossRef]
  98. Liu, F.; Wang, F.; Sun, X.; Wang, Y.; Liu, Y.; Zhang, S.; Li, Y.; Xue, Y.; Zhang, J.; Tang, C. Plasma-induced vacancies in CoS2 electrocatalysts to activate sulfur sites for hydrogen evolution reaction. Int. J. Hydrogen Energy 2024, 58, 941–947. [Google Scholar] [CrossRef]
  99. Nan, J.; Ye, B.; Peng, S.; Zhang, W.; Liu, H.; Zhang, Y. Self-support P-doped CoS2 as efficient hybrid nano-electrocatalysts for hydrogen evolution in acid and alkaline solutions. Mater. Lett. 2024, 354, 2–5. [Google Scholar] [CrossRef]
  100. Chen, Y.; Zhao, G.; Sun, W. Strategies of engineering 2D nanomaterial-based electrocatalysts toward hydrogen evolution reaction. Mater. Renew. Sustain. Energy 2020, 9, 10. [Google Scholar] [CrossRef]
  101. Chen, Y.; Yang, K.; Jiang, B.; Li, J.; Zeng, M.; Fu, L. Emerging two-dimensional nanomaterials for electrochemical hydrogen evolution. J. Mater. Chem. A 2017, 5, 8187–8208. [Google Scholar] [CrossRef]
  102. Yang, L.; Liu, P.; Li, J.; Xiang, B. Two-dimensional material molybdenum disulfides as electrocatalysts for hydrogen evolution. Catalysts 2017, 7, 285. [Google Scholar] [CrossRef]
  103. Zhu, Q.; Qu, Y.; Liu, D.; Ng, K.W.; Pan, H. Two-Dimensional Layered Materials: High-Efficient Electrocatalysts for Hydrogen Evolution Reaction. ACS Appl. Nano Mater. 2020, 3, 6270–6296. [Google Scholar] [CrossRef]
  104. Duc, D.A.; The, N.D.; Ha, D.H.; Phuong, N.T.T.; Manh, L.D. Comparative study on various morphologies of WS2 for electrochemical hydrogen evolution reaction. J. Sci. Technol. 2024, 6, 7. [Google Scholar] [CrossRef]
  105. Homayounfard, A.M.; Maleki, M.; Ghanbari, H.; Kahnamouei, M.H.; Safaei, B. Growth of few-layer flower-like MoS2 on heteroatom-doped activated carbon as a hydrogen evolution reaction electrode. Int. J. Hydrogen Energy 2024, 55, 1360–1370. [Google Scholar] [CrossRef]
  106. Li, P.Z.; Chen, N.; Al-Hamry, A.; Sheremet, E.; Lu, R.; Yang, Y.; Kanoun, O.; Baumann, R.R.; Rodriguez, R.D.; Chen, J.J. Inkjet-printed MoS2-based 3D-structured electrocatalysts on Cu films for ultra-efficient hydrogen evolution reaction. Chem. Eng. J. 2023, 457, 141289. [Google Scholar] [CrossRef]
  107. Wang, R.; Zhang, D.; Luo, S.; Wang, Q.; Jiang, L.; Wei, G.F.; Wang, X. Incorporation of tungsten nanoparticles on WS2 film for the enhanced hydrogen evolution. Int. J. Hydrogen Energy 2024, 59, 645–653. [Google Scholar] [CrossRef]
  108. Li, C.; Zhu, L.; Wu, Z.; Chen, Q.; Zheng, R.; Huan, J.; Huang, Y.; Zhu, X.; Sun, Y. Phase Engineering of W-Doped MoS2 by Magneto-Hydrothermal Synthesis for Hydrogen Evolution Reaction. Small 2023, 19, 2303646. [Google Scholar] [CrossRef] [PubMed]
  109. Wang, R.; Wan, L.; Liu, X.; Cao, L.; Hu, Y.; Dong, B. Anchoring Pt nanoparticle onto monolayer VS2 nanosheets boost efficient acidic hydrogen evolution. Int. J. Hydrogen Energy 2024, 74, 384–391. [Google Scholar] [CrossRef]
  110. Zhao, H.; Zhu, Y.P.; Yuan, Z.Y. Three-Dimensional Electrocatalysts for Sustainable Water Splitting Reactions. Eur. J. Inorg. Chem. 2016, 2016, 1916–1923. [Google Scholar] [CrossRef]
  111. Yun, Q.; Lu, Q.; Zhang, X.; Tan, C.; Zhang, H. Three-Dimensional Architectures Constructed from Transition-Metal Dichalcogenide Nanomaterials for Electrochemical Energy Storage and Conversion. Angew. Chem. Int. Ed. 2018, 57, 626–646. [Google Scholar] [CrossRef]
  112. Dong, H.; Liu, C.; Ye, H.; Hu, L.; Fugetsu, B.; Dai, W.; Cao, Y.; Qi, X.; Lu, H.; Zhang, X. Three-dimensional nitrogen-doped graphene supported molybdenum disulfide nanoparticles as an advanced catalyst for hydrogen evolution reaction. Sci. Rep. 2015, 5, 2–11. [Google Scholar] [CrossRef] [PubMed]
  113. Kumari, R.; Sammi, A.; Shubhangi; Srivastava, A.; Azad, U.P.; Chandra, P. Emerging 3D nanomaterials as electrocatalysts for water splitting reactions. Int. J. Hydrogen Energy 2024, 74, 214–231. [Google Scholar] [CrossRef]
  114. Huang, S.; Ma, S.; Liu, L.; Jin, Z.; Gao, P.; Peng, K.; Jiang, Y.; Naseri, A.; Hu, Z.; Zhang, J. P-doped Co3S4/NiS2 heterostructures embedded in N-doped carbon nanoboxes: Synergistical electronic structure regulation for overall water splitting. J. Colloid Interface Sci. 2023, 652, 369–379. [Google Scholar] [CrossRef] [PubMed]
  115. Fang, B.; Li, Y.; Yang, J.; Lu, T.; Liu, X.; Chen, X.; Pan, L.; Zhao, Z. Fe-Doped CoS2 Nanocages as Bifunctional Electrocatalysts for Water Splitting. ACS Appl. Nano Mater. 2024, 7, 9685–9695. [Google Scholar] [CrossRef]
  116. Xia, L.; Pan, K.; Wu, H.; Wang, F.; Zhao, N.; Yu, H.; Ren, Y.; Chang, M.; Dong, Z.; Wei, S. Ultra-thin WS2 decorated Co-based Metal-organic framework derived materials for synergistic electrocatalysis towards pH-Universal hydrogen evolution reaction. J. Phys. Chem. Solids 2024, 184, 111708. [Google Scholar] [CrossRef]
  117. Peng, Y.; Zhu, L.; Li, C.; Hu, J.; Lu, Y.; Fu, J.; Cui, F.; Wang, X.; Cao, A.; Ji, Q.; et al. Highly Stable Vertically Oriented 2H-NbS2 Nanosheets on Carbon Nanotube Films toward Superior Electrocatalytic Activity. Adv. Energy Mater. 2024, 14, 2302510. [Google Scholar] [CrossRef]
  118. Wang, T.; Zhang, X.; Yu, X.; Liu, Y.; Li, J.; Liu, Z.; Zho, N.; Zhang, J.; Niu, J.; Feng, Q. Modulating the electronic structure of VS2via Ru decoration for an efficient pH-universal electrocatalytic hydrogen evolution reaction. Nanoscale 2024, 16, 11250–11261. [Google Scholar] [CrossRef]
  119. Zavala, L.A.; Kumar, K.; Martin, V.; Maillard, F.; Maugé, F.; Portier, X.; Oliviero, L.; Dubua, L. Direct Evidence of the Role of Co or Pt, Co Single-Atom Promoters on the Performance of MoS2 Nanoclusters for the Hydrogen Evolution Reaction. ACS Catal. 2023, 13, 1221–1229. [Google Scholar] [CrossRef]
  120. Zhang, Y.; Yang, T.; Li, J.; Zhang, Q.; Li, B.; Gao, M. Construction of Ru, O Co-Doping MoS2 for Hydrogen Evolution Reaction Electrocatalyst and Surface-Enhanced Raman Scattering Substrate: High-Performance, Recyclable, and Durability Improvement. Adv. Funct. Mater. 2023, 33, 2210939. [Google Scholar] [CrossRef]
  121. Xu, Y.; Cheng, J.; Lv, H.; Zhang, K.; Hu, A.; Yang, X. Cactus-like amorphous MoS2-CoFeLDO heterostructures for alkaline hydrogen evolution reaction. Chem. Eng. J. 2023, 470, 144344. [Google Scholar] [CrossRef]
  122. Wang, R.; Chen, Q.; Liu, X.; Hu, Y.; Cao, L.; Dong, B. Synergistic Effects of Dual-Doping with Ni and Ru in Monolayer VS2 Nanosheet: Unleashing Enhanced Performance for Acidic HER through Defects and Strain. Small 2024, 20, 2311217. [Google Scholar] [CrossRef] [PubMed]
  123. Gupta, S. Electrocatalytic Water Splitting. In Photo-and Electro-Catalytic Processes: Water Splitting, N 2022; John Willy Sons: Hoboken, NJ, USA, 2022; pp. 123–158. [Google Scholar] [CrossRef]
  124. Pei, C.; Kim, M.C.; Li, Y.; Xia, C.; Kim, J.; So, W.; Yu, X.; Park, H.S.; Kim, J.K. Electron Transfer-Induced Metal Spin-Crossover at NiCo2S4/ReS2 2D–2D Interfaces for Promoting pH-universal Hydrogen Evolution Reaction. Adv. Funct. Mater. 2023, 33, 2210072. [Google Scholar] [CrossRef]
  125. Kim, G.; Jung, S.M.; Giri, A.; Kim, J.S.; Kim, Y.W.; Kim, K.S.; Choi, Y.; Lee, B.; Kim, Y.T.; Jeong, U. Enhanced hydrogen desorption via charge transfer in Pt Nanoclusters/ReS2 hybrid electrocatalyst for efficient hydrogen evolution reaction. J. Power Sources 2023, 579, 233287. [Google Scholar] [CrossRef]
  126. Liu, F.; Zhang, X.; Zong, H.; Xu, H.Z.; 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]
  127. Liu, Y.; Zhang, H.; Song, W.; Zhang, Y.; Hou, Z.; Zhou, G.; Zhou, 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, 138905. [Google Scholar] [CrossRef]
  128. Yao, C.; Wang, Q.; Peng, C.; Wang, R.; Liu, J.; Tsidaeva, N.; Wang, W. MOF-derived CoS2/WS2 electrocatalysts with sulfurized interface for high-efficiency hydrogen evolution reaction: Synthesis, characterization and DFT calculations. Chem. Eng. J. 2024, 479, 147924. [Google Scholar] [CrossRef]
  129. Trivedi, N.A.; Sharma, P.J.; Joshi, K.K.; Patel, V.; Sumesh, C.K.; Pataniya, P.M. ReS2/CoS heterostructure catalysts for electrochemical application in hydrogen evolution and supercapacitor. Int. J. Hydrogen Energy 2024, 61, 1212–1219. [Google Scholar] [CrossRef]
  130. Li, X.; Han, S.; Qiao, Z.; Zeng, X.; Cao, D.; Chen, J. Ru monolayer island doped MoS2 catalysts for efficient hydrogen evolution reaction. Chem. Eng. J. 2023, 453, 139803. [Google Scholar] [CrossRef]
  131. Ghising, R.B.; Pan, U.N.; Kandel, M.R.; Dhakal, P.P.; Sidra, S.; Kim, D.H.; Kim, N.H.; Lee, J.H. Ruthenium single atoms implanted on NiS2-FeS2 nanosheet heterostructures for efficacious water electrolysis. J. Mater. Chem. A 2024, 12, 3489–3500. [Google Scholar] [CrossRef]
  132. Yang, Y.; Chen, Y.; Xiong, Y.; He, Y.; Sun, Q.; Xu, D.; Hu, Z. Self-supported monometallic FeS2/FeOOH-ZnO@NF with abundant oxygen vacancies as efficient and stable electrocatalysts for the OER and HER. J. Alloys Compd. 2024, 991, 174525. [Google Scholar] [CrossRef]
  133. Zhan, W.; Liu, G.; Lang, L.; Gao, X.S.; Zheng, B.; Zhang, J.; Luo, G.; Hu, L.; Chen, W. Sulfur vacancy riched pure 2H phase VS2 for efficient electrocatalytic hydrogen evolution. J. Electroanal. Chem. 2024, 971, 118609. [Google Scholar] [CrossRef]
  134. Tang, J.; Huang, J.; Zhang, S.; Liu, Z.; Deng, X. Synergistic Low-Pt Modification and Ni Doping Facilitate the MoS2 Phase Transition for Accelerating the Hydrogen Evolution Reaction. Energy Fuels 2023, 37, 14914–14921. [Google Scholar] [CrossRef]
  135. Di, F.; Wang, X.; Farid, S.; Ren, S. CoS2 with carbon shell for efficient hydrogen evolution reaction. Int. J. Hydrogen Energy 2023, 48, 17758–17768. [Google Scholar] [CrossRef]
  136. Zhao, W.K.; Ma, Z.Q.; Zheng, J.Y.; Han, C.B.; Zhou, K.L.; Hao, M.Y.; Fang, D.C.; Xia, Y.; Yan, X. Dual-functional MnS2/MnO2 heterostructure catalyst for efficient acidic hydrogen evolution reaction and assisted degradation of organic wastewater. J. Energy Chem. 2023, 87, 215–224. [Google Scholar] [CrossRef]
  137. Mariappan, A.; Dharman, R.K.; Oh, T.H. Metal-organic frameworks derived FeS2@CoS2 heterostructure for efficient and stable bifunctional electrocatalytic water splitting. Ceram. Int. 2023, 49, 29984–29990. [Google Scholar] [CrossRef]
Figure 1. (a,b) Structures of TMDs. Schematic representations of a typical TMD structure with trigonal prismatic. (c) Coordination from c-axis (upper) and section view (middle). (d,e) Schematic illustration of the advantages of 2D TMDs as the electrocatalyst [42] Reproduced with permission from ref. [42]. © 2022 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.4. Different TMDs for HER.
Figure 1. (a,b) Structures of TMDs. Schematic representations of a typical TMD structure with trigonal prismatic. (c) Coordination from c-axis (upper) and section view (middle). (d,e) Schematic illustration of the advantages of 2D TMDs as the electrocatalyst [42] Reproduced with permission from ref. [42]. © 2022 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.4. Different TMDs for HER.
Hydrogen 05 00041 g001
Figure 2. SEM images of (a) NF, (b) MoO3/MXene/NF, and (c) MoS2/MXene/NF. The insets of (b,c) are images at high magnification. (d,e) TEM images and (f) HRTEM image of MoO3/MXene removed from MoO3/MXene/NF. (g,h) TEM images, (i) HRTEM image, (j) HAADF image, and (k) EDX elemental distributions of MoS2/MXene removed from MoS2/MXene/NF [85]. Reproduced with permission from ref. [85]. © 2022 the Royal Society of Chemistry.
Figure 2. SEM images of (a) NF, (b) MoO3/MXene/NF, and (c) MoS2/MXene/NF. The insets of (b,c) are images at high magnification. (d,e) TEM images and (f) HRTEM image of MoO3/MXene removed from MoO3/MXene/NF. (g,h) TEM images, (i) HRTEM image, (j) HAADF image, and (k) EDX elemental distributions of MoS2/MXene removed from MoS2/MXene/NF [85]. Reproduced with permission from ref. [85]. © 2022 the Royal Society of Chemistry.
Hydrogen 05 00041 g002
Figure 3. FESEM images of CoS2 (a), CNFS (b), and CNFS@rGO (c), EDS spectra recorded from CNFS@rGO (d) and ICP analysis results of the as-synthesized CNFS (e) [67]. Reproduced with permission form ref. [67]. © 2017 Elsevier.
Figure 3. FESEM images of CoS2 (a), CNFS (b), and CNFS@rGO (c), EDS spectra recorded from CNFS@rGO (d) and ICP analysis results of the as-synthesized CNFS (e) [67]. Reproduced with permission form ref. [67]. © 2017 Elsevier.
Hydrogen 05 00041 g003
Figure 4. (a) SEM image of carbon cloth (CC). (b) SEM image of Co(OH)F/CC. (c,d) Low- and high-magnification SEM image of CoS2. (eg) STEM elemental mapping of Co and S. (hj) SEM images of GDY/CoS2/CC. (k) HRTEM images of GDY/CoS2 nanowire. (lo) The STEM elemental mapping of Co, S, and C for the GDY/CoS2 nanowire [52]. Reproduced with permission from ref. [52]. © 2021 Elsevier.
Figure 4. (a) SEM image of carbon cloth (CC). (b) SEM image of Co(OH)F/CC. (c,d) Low- and high-magnification SEM image of CoS2. (eg) STEM elemental mapping of Co and S. (hj) SEM images of GDY/CoS2/CC. (k) HRTEM images of GDY/CoS2 nanowire. (lo) The STEM elemental mapping of Co, S, and C for the GDY/CoS2 nanowire [52]. Reproduced with permission from ref. [52]. © 2021 Elsevier.
Hydrogen 05 00041 g004
Figure 5. SEM images of the as-synthesized Ni–Co precursors (ac), NixCo1−xS2 (df), and EDS spectra. The corresponding Ni/Co molar ratios were 1:2 (a,d,g), 1:1 (b,e,h), and 2:1 (c,f,i), respectively [69]. Reproduced with permission from ref. [69]. © 2017 MDPI, Basel, Switzerland.
Figure 5. SEM images of the as-synthesized Ni–Co precursors (ac), NixCo1−xS2 (df), and EDS spectra. The corresponding Ni/Co molar ratios were 1:2 (a,d,g), 1:1 (b,e,h), and 2:1 (c,f,i), respectively [69]. Reproduced with permission from ref. [69]. © 2017 MDPI, Basel, Switzerland.
Hydrogen 05 00041 g005
Figure 6. SEM images of (a) NiCu-LDHs and (b,c) Cu-NiS2. (d) Low-magnification TEM and (e) HRTEM images of Cu-NiS2 (inset: SAED pattern). (f) HAADF image of Cu-NiS2 and the corresponding STEM–EDX elemental mappings of Ni, Cu, and S [73]. Reproduced with permission from ref. [73]. © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 6. SEM images of (a) NiCu-LDHs and (b,c) Cu-NiS2. (d) Low-magnification TEM and (e) HRTEM images of Cu-NiS2 (inset: SAED pattern). (f) HAADF image of Cu-NiS2 and the corresponding STEM–EDX elemental mappings of Ni, Cu, and S [73]. Reproduced with permission from ref. [73]. © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Hydrogen 05 00041 g006
Figure 7. (a) TEM image of pure MoS2 nanosheets, (b) TEM image of Co3O4-Pt composite, (c,d) Low and high magnification TEM image of Mo-Co-Pt ternary composite [76]. Reproduced with permission from ref. [76]. © 2020 Elsevier.
Figure 7. (a) TEM image of pure MoS2 nanosheets, (b) TEM image of Co3O4-Pt composite, (c,d) Low and high magnification TEM image of Mo-Co-Pt ternary composite [76]. Reproduced with permission from ref. [76]. © 2020 Elsevier.
Hydrogen 05 00041 g007
Figure 8. XRD patterns of flower-like MoS2 and MoS2/MoN-8 (a); SEM images of (b,c) MoS2 and (d,e) MoS2/MoN-8 and TEM and HRTEM images of MoS2/MoN-8 (fi) [77]. Reproduced with permission from ref. [77]. © 2021 Elsevier.
Figure 8. XRD patterns of flower-like MoS2 and MoS2/MoN-8 (a); SEM images of (b,c) MoS2 and (d,e) MoS2/MoN-8 and TEM and HRTEM images of MoS2/MoN-8 (fi) [77]. Reproduced with permission from ref. [77]. © 2021 Elsevier.
Hydrogen 05 00041 g008
Figure 9. SEM images of (a,b) MoS2/GF and (c,d) a-MoS2/GF. (eh) Corresponding EDS mappings of C, O, S, and Mo elements of MoS2/GF [82]. Reproduced with permission from ref. [82]. © 2019 Elsevier.
Figure 9. SEM images of (a,b) MoS2/GF and (c,d) a-MoS2/GF. (eh) Corresponding EDS mappings of C, O, S, and Mo elements of MoS2/GF [82]. Reproduced with permission from ref. [82]. © 2019 Elsevier.
Hydrogen 05 00041 g009
Table 1. Summary of different TMDs and their electrochemical measurements for HER with different morphologies.
Table 1. Summary of different TMDs and their electrochemical measurements for HER with different morphologies.
Sr No.MaterialsMorphology of TMDsSynthesis MethodOverpotential η [mV]
at Current Density 10 mA cm−2
Tafel Slope
[mV dec−1]
Electrolyte Ref.
1.1T’ phase WS2Nanoparticles Colloidal
synthesis
200 50.40.5 M H2SO4[64]
2. MoS2NanoparticlesPyrolysis
and sulfurization methods
200620.5 M H2SO4[66]
3.CNFS@rGONanoparticlesHydrothermal138 490.5 M H2SO4[67]
4.MoS2/rGONanoparticlesHydrothermal70 440.5 M H2SO4[68]
MoS2/MWCNT19073
5.Graphdiyne-CoS2Nanowires-97561 M KOH[52]
6.NbS2-PPyNanowiresChemical
vapor deposition
21956-[65]
7.Ni-doped CoS2 (Ni0.33Co0.67S2 NN/SS)NanoneedleHydrothermal350761 M KOH[69]
8.N−CoS2NanowiresHydrothermal152580.5 M H2SO4[70]
9.MoS2Nanosheets-200 65-[71]
10.W2C@WS2NanoflowersHydrothermal17055.40.5 M H2SO4[72]
11.Cu-doped NiS2Nanosheets-139771 M KOH[73]
12.MoS2.ZnONanoflakes Microwave-assisted method239620.5 M H2SO4[74]
13.Pt-decorated VS2NanosheetsHydrothermal7754.270.5 M H2SO4[75]
14.Mo-Co-PtNanosheets
nanocubes
Hydrothermal20340.5 M H2SO4[76]
15.MoS2/MoNNanospheresHydrothermal117870.5 M H2SO4[77]
132981 M KOH
16.WS2TN/CCTriangular nanoplatesHydrothermal196870.5 M H2SO4[78]
1502001.0 M PBS
1931701 M KOH
17.MoS2-rGO@MoNanosheetsHummers method
Hydrothermal
123621 M KOH[79]
18.VF-MoS2-rGO-CNTsHybridHydrothermal400390.5 M H2SO4[80]
19.CoS2/MWCNTsNanospheresHydrothermal257830.5 M H2SO4[81]
20.MoS2/GFFlower-like structureHydrothermal82480.5 M H2SO4[82]
Table 2. Recent reports on TMDs and their composites for HER.
Table 2. Recent reports on TMDs and their composites for HER.
Sr No.MaterialsMorphology of TMDsSynthesis MethodOverpotential η [mV]
at Current Density 10 mA cm−2
Tafel Slope
[mV dec −1]
Electrolyte Ref.
1.Ru–VS2/CCNanoparticlesHydrothermal8963 0.5 M H2SO4[118]
2.Pt1%−CoMoS2/CNanoclustersIncipient wetness impregnation method11868 0.5 M H2SO4[119]
3.MSOR1
(Ru co-doped MoS2)
Flower-like nanosphereHydrothermal19763.1 0.5 M H2SO4[120]
4375.8 1.0 M KOH
4.MoS2-CoFeLDHNanoneedle (cactus-like spherical)Hydrothermal,
Electrodeposition
36 235 1.0 M KOH [121]
5.SL-Ni-Ru-VS2NanosheetsColloidal20 34.14 0.5 M H2SO4[122]
6.NiFe LDH/NiS2/VS2NanosheetsHydrothermal7679 1.0 M KOH [123]
7.NiCo2S4/ReS2NanosheetsHydrothermal 126 67.8 0.5 M H2SO4[124]
85 78.3 1.0 M KOH
8.Pt/ReS2NanoflowersMicrowave-assisted hydrothermal method,
Wet-impregnation method
2031.5 Ar-saturated 0.1 M HClO4[125]
9.Co-NiS2/CeO2/NFNanosheetsHydrothermal84115.3 1.0 M KOH [126]
10.ReS2/NiSNanosheetsHydrothermal7876 1.0 M KOH [127]
11.CoS2/WS2NanosheetsHydrothermal7952 0.5 M H2SO4[128]
12.ReS2/CoSNanosheetsHydrothermal18761 1.0 M KOH [129]
13.Ru MIs-doped MoS2NanosheetsHydrothermal1763 1.0 M KOH [130]
14.RuSA-NiS2-FeS2NanosheetsHydrothermal5759.4 1.0 M KOH [131]
15.FeS2/FeOOH-ZnO@NFNanoplatesHydrothermal74105.79 1.0 M KOH [132]
16.2H–VS2NanoflowerHydrothermal18152 0.5 M H2SO4[133]
17.Pt@1T-MoS2−NiPetal-like nanosheetsHydrothermal7646 1.0 M KOH [134]
18.CoS2@NHCs-800MicrospheresHydrothermal 98 85 0.5 M H2SO4[135]
118 157 1.0 M KOH
19.MnS2/MnO2-CCHeterostructureElectrodeposition6626.72 0.5 M H2SO4[136]
20.FeS2@CoS2HeterostructureHydrothermal13658 1.0 M KOH [137]
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

Jakkanawar, S.S.; Chavan, V.D.; Kim, D.-K.; Bhat, T.S.; Yadav, H.M. Tuning the Morphology of Transition Metal Disulfides: Advances in Electrocatalysts for Hydrogen Evolution Reaction. Hydrogen 2024, 5, 776-799. https://doi.org/10.3390/hydrogen5040041

AMA Style

Jakkanawar SS, Chavan VD, Kim D-K, Bhat TS, Yadav HM. Tuning the Morphology of Transition Metal Disulfides: Advances in Electrocatalysts for Hydrogen Evolution Reaction. Hydrogen. 2024; 5(4):776-799. https://doi.org/10.3390/hydrogen5040041

Chicago/Turabian Style

Jakkanawar, Shravani S., Vijay D. Chavan, Deok-Kee Kim, Tejasvinee S. Bhat, and Hemraj M. Yadav. 2024. "Tuning the Morphology of Transition Metal Disulfides: Advances in Electrocatalysts for Hydrogen Evolution Reaction" Hydrogen 5, no. 4: 776-799. https://doi.org/10.3390/hydrogen5040041

APA Style

Jakkanawar, S. S., Chavan, V. D., Kim, D. -K., Bhat, T. S., & Yadav, H. M. (2024). Tuning the Morphology of Transition Metal Disulfides: Advances in Electrocatalysts for Hydrogen Evolution Reaction. Hydrogen, 5(4), 776-799. https://doi.org/10.3390/hydrogen5040041

Article Metrics

Back to TopTop