Advanced 2D molybdenum disulfide for green hydrogen production: Recent progress and future perspectives

Meng FANG , Yuqin PENG , Puwei WU , Huan WANG , Lixin XING , Ning WANG , Chunmei TANG , Ling MENG , Yuekuan ZHOU , Lei DU , Siyu YE

Front. Energy ›› 2024, Vol. 18 ›› Issue (3) : 308 -329.

PDF (14001KB)
Front. Energy ›› 2024, Vol. 18 ›› Issue (3) : 308 -329. DOI: 10.1007/s11708-024-0916-x
MINI REVIEW

Advanced 2D molybdenum disulfide for green hydrogen production: Recent progress and future perspectives

Author information +
History +
PDF (14001KB)

Abstract

The development of renewable and affordable energy is crucial for building a sustainable society. In this context, establishing a sustainable infrastructure for renewable energy requires the integration of energy storage, specifically use of renewable hydrogen. The hydrogen evolution reaction (HER) of electrochemical water splitting is a promising method for producing green hydrogen. Recently, two-dimensional nanomaterials have shown great promise in promoting the HER in terms of both fundamental research and practical applications due to their high specific surface areas and tunable electronic properties. Among them, molybdenum disulfide (MoS2), a non-noble metal catalyst, has emerged as a promising alternative to replace expensive platinum-based catalysts for the HER because MoS2 has a high inherent activity, low cost, and abundant reserves. At present, greatly improved activity and stability are urgently needed for MoS2 to enable wide deployment of water electrolysis devices. In this regard, efficient strategies for precisely modifying MoS2 are of interest. Herein, the progress made with MoS2 as an HER catalyst is reviewed, with a focus on modification strategies, including phase engineering, morphology design, defect engineering, heteroatom doping, and heterostructure construction. It is believed that these strategies will be helpful in designing and developing high-performance and low-cost MoS2-based catalysts by lowering the charge transfer barrier, increasing the active site density, and optimizing the surface hydrophilicity. In addition, the challenges of MoS2 electrocatalysts and perspectives for future research and development of these catalysts are discussed.

Graphical abstract

Keywords

molybdenum disulfide (MoS2) / hydrogen evolution reaction (HER) / active site / electrocatalyst / modification strategie

Cite this article

Download citation ▾
Meng FANG, Yuqin PENG, Puwei WU, Huan WANG, Lixin XING, Ning WANG, Chunmei TANG, Ling MENG, Yuekuan ZHOU, Lei DU, Siyu YE. Advanced 2D molybdenum disulfide for green hydrogen production: Recent progress and future perspectives. Front. Energy, 2024, 18(3): 308-329 DOI:10.1007/s11708-024-0916-x

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

Conventional fossil fuels can no longer meet the needs of the human society, which has prompted us to develop renewable and green energy sources. Hydrogen, a carbon-free energy source, has attracted tremendous attention as a potential replacement for fossil fuels in the future. Electrocatalysis and photocatalysis are currently two of the most promising hydrogen production technologies. Photocatalytic hydrogen production has the advantages of low cost and simple hydrogen production. Since the introduction of photocatalytic hydrogen production, researchers have developed a variety of photocatalysts for hydrogen formation, such as graphite carbon nitride (g-C3N4) [1], linear conjugated polymers [2], conjugated microporous polymers [3], and covalent organic frameworks (COFs) [4]. However, the current efficiencies for photocatalytic hydrogen production are inadequate for commercial application. Water electrolysis for hydrogen production is a well-established industry with advantages arising from low hydrogen impurity levels and decreased carbon emissions. In particular, water electrolysis comprises the anodic oxygen evolution reaction (OER) and cathodic hydrogen evolution reaction (HER). Therefore, the performance is affected by both the anode and cathode processes. In previous papers, the progress made with the OER [5,6] are carefully reviewed, and in this paper, focus is laid on HER.

To date, water splitting and the HER often necessitate the use of precious metals, such as Pt. However, these metals are both costly and scarce [7]. In this regard, the development of nonnoble metal catalysts with high activity and stability is being pursued [8]. Promising substitutes for the noble metal catalysts used for the HER include transition metal oxides (TMOs), transition metal carbides (TMCs), transition metal phosphides (TMPs), transition metal nitrides (TMNs), and transition metal dichalcogenides (TMDs) [9]. TMDs have become popular and promising non-noble metal electrocatalysts for HER because of their atomic-level thicknesses, complete exposure of active sites, adjustable electronic structures, diverse modulation strategies, and low hydrogen adsorption free energy [10]. In particular, molybdenum disulfide (MoS2) has been intensively investigated. MoS2 is one of two-dimensional (2D) TMDs nanomaterials. MoS2 is a widely-used material for electrocatalytic hydrogen evolution due to its high catalytic activity at edge active site, easy accessibility, and low cost.

Conventionally, it is believed that the unsaturated sulfur atoms located at the edges of MoS2 are the catalytic active centers, while the basal planes make no contribution to HER. The edge structure of MoS2 is very similar to that of the active site of nitrogenase and exhibits ΔGH* 0 [11]. In HER, nitrogenase has a hydrogen formation activity similar to that of Pt. As a result, MoS2 is said to have a Pt-like activity. Jaramillo et al. [12] conducted a point-to-point comparison of the edge active sites of MoS2 with the sites on the Pt (111) face. They found that the exchange current density at the MoS2 edge was significantly higher than that of common metals and only lower than that of Pt group metals. MoS2 has almost no active sites in the basal planes and is prone to stacking, which causes some of the active sites at the edges to disappear. Modifications have been proposed to enhance the activity of MoS2, including phase engineering, morphology design, defect engineering, heteroatom doping, and heterostructure construction (Fig.1). Herein, the recent progress made in modifying MoS2 for use in the HER is reviewed and perspectives on future R&D and beyond are provided.

2 Structural features of MoS2

MoS2 has a layered structure comprising individual S−Mo−S units, the adjacent layers are bonded through van der Waals (VDW) forces, and the S and Mo atoms are bonded via strong ionic-covalent bonds [13]. The bonding mechanism is similar to that of graphite [14]. MoS2 is commonly a black-gray semiconductor in the form of solid powder, which exhibits a metallic luster and a good chemical/thermal stability [15].

MoS2 has four crystal structures, including a distorted tetragonal phase (1T), hexagonal phases (1H and 2H), and a rhombohedral phase (3R), whose structures are shown in Fig.2(a)–2(d) [16]. Among them, 1T-MoS2 and 3R-MoS2 are metastable states, while 2H-MoS2 widely exists under natural conditions. The arrangement of atomic layers (S−Mo−S) in the 1T phase of MoS2 is significantly different from those in the 2H and 3R phases [17]. Particularly, the 1T phase has a regular octahedra with Mo atoms at the centers and coordination by S atoms, and there is only one layer for the repeat units in the crystal unit cell. However, the 2H phase has triangular prisms with Mo atoms at the center, and there are two layers of repeat units in the crystal unit cell. The repeat unit has hexagonal symmetry. Similarly, the 3R phase also has triangular prisms but contains three layers of repeat units in the crystal unit cell. The repeat unit has rhombohedral symmetry.

In MoS2 nanosheets, the crystal phases, sizes, and chemical compositions determine the electrocatalytic properties [18]. Martis et al. [19] used four-dimensional scanning transmission electron microscopy (4D-STEM) (Fig.2(e)) to generate annular dark field (ADF)-STEM (Fig.2(g)) and center of mass images (Fig.2(h) and 2(i)). This reveals that 2H-MoS2 is a 2D semiconductor composed of Mo atoms sandwiched between two S atoms, and it has a direct band gap. The semiconducting properties and direct bandgap make it highly suitable for catalytic applications.

3 Modification of MoS2 for HER

Although the active sites of MoS2 have a hydrogen binding energy comparable to that of Pt, they are mostly located at the edges and are hardly present at the basal planes of MoS2. The blocky structure of MoS2 results in a prevalent concentration of VDW forces within the layers, leading to a lower coordination of the active sites at the edges. Furthermore, the atoms in the central layer of the blocky MoS2 form covalent bonds, which results in fewer exposed active sites. Consequently, the HER performance is diminished [21]. To address this problem, researchers have invested tremendous effort in modifications.

3.1 Phase engineering

Unlike semiconductor-phase MoS2, 1T-MoS2 demonstrates a metal-like electrical conductivity, which is the reason that it is often referred to as metal-phase MoS2. 1T-MoS2 achieves this electrical conductivity due to the presence of unoccupied empty orbitals located near the Fermi energy level [22]. Reasonably, after the transformation from the 2H phase to the 1T phase, the electronic conductivity of MoS2 is increased by approximately 107 times. The phase transition of monolayer MoS2 is accomplished through lateral displacement of S atoms within the S planes (as shown in Fig.3(a)), which is facilitated by intralayer atomic plane gliding [23]. In 2H-MoS2, a layer of S atoms in the S−Mo−S structure undergoes displacement, resulting in the transformation from the original triangular symmetry to an octahedral coordination structure. The electrocatalytic activities of the different MoS2 crystalline phases are completely different. This is attributed to the fact that active sites in 2H-MoS2 are only generated by the edge uncoordinated S atoms, while the basal surface of 1T-MoS2 exhibits a catalytic activity akin to the edge atoms. Moreover, the refined electronic conductivity and hydrophilicity of the 1T phase synergistically improve the electrolyte accessibility during electrolysis. Compared to semiconducting 2H-MoS2, metallic 1T-MoS2 exhibits a larger band gap, which promotes faster electron transfer. Manipulating the proportions of different phases in MoS2 can alter the reaction rates for electrocatalytic reactions [24].

1T-MoS2 is typically obtained via the “top-down” and the “bottom-up” methods: The “top-down” approach employs 2H-MoS2 as the raw material for the production of 1T-MoS2 by chemical, electrochemical, physical or a combination of physical and chemical techniques [17]. For instance, heavy electron doping, such as with Li+, Na+, and K+ intercalation, plasmonic hot electron doping, electron irradiation, and substitutional doping, is employed to introduce foreign species into 2H-MoS2, thereby achieving the 2H → 1T transformation [25]. The “bottom-up” approach is primarily used to synthesize 1T-MoS2 by using Mo sources and S sources and methods such as low-temperature chemical vapor deposition (CVD) and hydrothermal and solvothermal methods.

In “top-down” synthetic methods, the main approach is chemical/electrochemical intercalation with alkali metal ions. In MoS2, the interlayer spacing between the atoms is relatively large, and the VDW forces between them are weak. Therefore, alkali metals are easily intercalated into the atomic layers. In general, researchers chemically exfoliate 2H-MoS2 by inserting ions such as Li+, Na+, and K+ to obtain 1T-MoS2. The resulting 1T-MoS2 material is usually hydrophilic. Liu et al. [26] found that the 2H to 1T phase transition of MoS2 in the lithium intercalation method is induced by insertion of the lithium atoms. The insertion of lithium reduces the energy barrier for movement of the S atoms. In 2H-MoS2, the phase transition is induced by individual transitions of the S atoms rather than collective behavior. Tan et al. [27] proposed an approach that combined ball milling and chemical Li intercalation to achieve a high-yield syntheses of small monolayer MoS2 nanoparticles, a high percentage of the 1T phase, and an exceptionally large specific surface area. Chen et al. [28] demonstrated a simple and direct strategy for intercalating zero-valent metals into bulk MoS2 (Fig.3(b)). Since block-like 2H-MoS2 itself does not have a strong reduction capacity, lithium was first intercalated into bulk MoS2 to obtain 1T'-LixMoS2. It was found that 1T'-LixMoS2 had an increased electron density and a larger reduction potential, and this characteristic was used for in-situ reduction of the metal ions on the surface of LixMoS2 to achieve intercalation of the zero-valent metal atoms. As shown by time-of-flight secondary ion mass spectroscopy (ToF-SIMS) (Fig.3(c)), the zero-valent metal atoms in Pt nanoparticles were successfully intercalated into MoS2. From the stable 2H phase to the unstable 1T' phase, the insertion of Li also increases the interlayer distance of MoS2, which allows for diffusion exchange between MoS2 and the target metal salt ions, and the Li ions inserted into MoS2 can be removed. Inserted metals can improve the stability of the MoS2 1T' phase for use in the HER.

Although the intercalation method is widely used in industrial production, it is limited by the use of hazardous organic solvents and explosive Li+ for embedding the ions, which requires inert atmospheres, as well as the slow kinetics of the intercalation/exfoliation processes. Electron irradiation can be used to optimize the properties and defects of materials by controlling the irradiation conditions, and the irradiation parameters can be used to adjust the structures of materials, thereby changing the chemical and physical properties and obtaining the best catalytic performance. Compared with the intercalation method, electron irradiation achieves a more precise modulation of materials without chemical residues affecting the HER performance. Dong et al. [29] performed electron irradiation on MoS2 in an atmospheric environment with a high-voltage electron accelerator. After electron beam irradiation, the MoS2 nanostructures exhibited a strong desulfurization and formation of the 1T phase in the prepared MoS2, which was controlled by adjusting the energy density of electron irradiation. As the irradiation energy was increased, the content of the 1T phase increased, which generated more active sites and significantly improved the electrocatalytic activity of the inert MoS2 base. Phase transition-induced atomic migration alters the charge distribution on the atomic surfaces, facilitating rapid transfer of H2 at the catalytic interface. However, irradiation of MoS2 with excessive energy can cause severe aggregation, resulting in decreases in the specific surface area and conductivity. It was necessary to control the energy density within a suitable range to obtain a good HER performance.

The “top-down” method of MoS2 phase transitions can also be induced by plasma treatment and atomic doping. Zhu et al. [30] demonstrated that a weak Ar plasma treatment to activate the 2H → 1T phase transition of MoS2 provided a 2H to 1T phase conversion ratio of about 40%. The proportion of the MoS2 2H → 1T phase transition realized by elemental doping has been limited from 40% to 60%, mainly because elemental doping occurs on the surface and cannot adjust the internal lattice. Sun et al. [31] discovered a way to promote the 2H → 1T phase transition of MoS2 nanosheets by utilizing the synergistic effect of Pt atom doping and an N2 plasma, and the specific steps are shown in Fig.4(a). They first directly grew 2H MoS2 on carbon cloth (CC) by the hydrothermal method and then immersed the CC coated with MoS2 nanosheets in H2PtCl6·6H2O (0.5 mol/L, 10 mL) for Pt atom doping. The impregnated material was dried and then bombarded with an N2 plasma for 2 h. The interlayer spacing of a material that was not soaked in the H2PtCl6·6H2O solution increased from 0.64 to 0.67 nm, while the interlayer spacing of the material soaked in the H2PtCl6·6H2O solution increased to 0.69 nm (Fig.4(b)). The larger interlayer spacing indicated that more of the 2H phase in the MoS2 was converted to the 1T phase, suggesting that Pt doping improved the conversion rate of the 1T phase. The proportions of the 1T phase in N-MoS2 and N, Pt-MoS2 nanosheets are 62% and 87%, respectively. Synergistic doping of N and Pt atoms altered the atomic arrangement and electronic structure of MoS2 (Fig.4(e) and Fig.4(f)), which subsequently activated the S vacancies and formed empty 2pz orbitals that facilitated water adsorption and dissociation. N, Pt-MoS2 nanosheets exhibited a lower overpotential (Fig.4(d)) similar to that of 20 wt% of Pt/C, suggesting that N, Pt-MoS2 has a lower charge transfer resistance.

“Bottom-up” synthesis methods can address the complex synthesis limitations in “top-down” methods. Therefore, “bottom-up” one-pot methods such as hydrothermal and solvothermal syntheses are promising and direct strategies for preparing 1T-MoS2. Solvothermal synthesis offers a simple approach to 1T-MoS2. However, the effectiveness of this process in the 2H → 1T phase conversion is restricted, and it may generate unforeseen byproducts such as MoO2 and MoO3. Mai et al. [32] used a direct and effective solvothermal synthesis that optimized and quantified the 1T and 2H phases in MoS2. During the solvothermal synthesis, the S2− ions in thiourea (NH2CSNH2) reduced Mo6+ in the ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O) to Mo4+, and then S2− combined with Mo4+ to form MoS2. Thiourea also produced OH during hydrolysis, which combined with Mo6+ and Mo4+ to form MoxOy. They found that S2− was more reactive at high temperatures, which inhibited the formation of MoxOy. However, an overly high temperature converts the 1T phase to the 2H phase. Mai et al. [32] found that 220 °C was the optimal synthetic temperature, since it inhibited byproduct generation and provided a high proportion of the 1T phase, thus solving the main problem of the solvothermal method in the synthesis of 1T-MoS2. Zhu et al. [33] synthesized 1T-MoS2 with a magneto-hydrothermal method, which provided MoS2 with a pure 1T phase at a magnetic field of 9T. The ΔGH* of 1T-MoS2 was 0.7 eV lower than that of 2H-MoS2 (1.96 → 1.26 eV), and the modification with trace amounts of Ru reduced the ΔGH* of 1T-MoS2 from 1.26 to 0.02 eV. Therefore, it adsorbed hydrogen ions more easily.

Zhang et al. [34] developed a one-step solvothermal method to obtain 1T-MoS2 with extended interlayer spacing derived from Mo-based organic frameworks. Metal-organic frameworks (MOFs) have been used for HER due to their exceptional properties, including high specific surface areas, tunable pore sizes, easily modifiable compositions, and diverse morphological structures. However, most MOFs have low conductivities and stabilities, which significantly limit their application. They used N, N-dimethylformamide (DMF) oxide to increase the surface spacing of MoS2 and successfully synthesized 1T-MoS2 with a 10.87 Å layer spacing. The specific experimental steps for the preparation of 1T-MoS2 are shown in Fig.5(a). The uniform solution yielded Mo-based MOFs (Mo-MOFs) at 120 °C, which were totally transformed into 1T-MoS2 after heating to 200 °C (Fig.5(b)). Scanning electron microscopy (SEM), transmission electron microscopy (TEM), HRTEM, and water contact angle images for 2H-MoS2, 180-1T-MoS2, and 200-1T-MoS2 (Fig.5(d)) showed that their structures became progressively looser and more dispersed. As the samples progressed from 2H-MoS2 to 180-1T-MoS2 and then to 200-1T-MoS2, their interlayer spacings increased, and they became more hydrophilic. This strategy avoided structural collapse of the MOFs during carbonization and coverage of the active sites on the materials. Furthermore, the HER catalytic performance of the MoS2 was improved by increasing the crystal face spacing and reducing the adsorption–desorption potential for protons.

In industrial production, the “top-down” method is the main method used to obtain the 1T-phase of MoS2. Generally, the “top-down” method is used to obtain monolayer 1T-MoS2, but the structural stability of monolayer 1T-MoS2 is poor. Due to the interlayer S-S VDW forces, 1T-MoS2 is gradually converted to 2H-MoS2, and the MoS2 monolayers are stacked on top of each other, which masks the catalytic active sites. The “bottom-up” method is an inexpensive and simple route to 1T-MoS2, but the hydrothermal/solvothermal reaction is not easily observed and difficult to regulate. Phase transformation is an effective strategy for modulating electronic properties and increasing the number of active sites, but there are still many challenges in phase transformation modification. The monolayer dispersion of 1T-MoS2 impedes modification of the microstructure and modulation of the electronic structure, and the 1T-MoS2 basal atoms, although active, are still less reactive than the edge sites, which restricts the catalytic activity. Researchers have introduced S vacancies, doped elements and conductive carbon substrates into MoS2 to obtain highly efficient hydrogen generation with 1T-MoS2. Phase-change engineering still has very broad prospects, and it is worth exploring in depth.

3.2 Morphology designs

Bulk MoS2 usually has a low electronic conductivity. One way to improve the catalytic performance of MoS2 is to adjust the electronic structure by creating an appropriate microstructure. Excellent work was performed to unveil the relationships between the morphology and electrochemical activity of MoS2. In 2005, Hinnemann et al. [11] used density functional theory (DFT) calculations to show that the catalytically active sites of MoS2 were located at its edges, while the basal surface was inert and usually not involved in construction of the catalytic active centers. Subsequently, researchers used various studies to maximize the density of edge active sites and expose more edge areas per unit area by morphological design engineering to modify the MoS2 catalyst. For example, nanospheres, nanowires, nanosheets, vertically aligned nanosheet arrays, and mesoporous structures were constructed (Fig.6(a)).

Jiang et al. [35] synthesized MoS2 nanoparticles via an electrochemical method and uniformly dispersed them on a reduced graphene oxide-modified carbon nanotube/polyimide (PI/CNT-RGO) film. The MoS2 nanoparticles had highly exposed edge sites and exhibited a low overpotential and electrocatalytic activity for the HER. Yi et al. [36] used a solid-state synthesis to prepare MoS2 nanorods in a one-step pyrolysis process with Mo-MOFs as the molybdenum source in the presence of thiourea. Behranginia et al. [37] synthesized MoS2 with a 3D crystal structure and bare Mo-edge atoms by CVD. In this method, the temperature was precisely controlled to maintain the same vaporization rate for S and MoO3 and prevent the generation of undesirable byproducts. Accordingly, the MoS2 grew directly on the defects of glassy carbon (GC) and graphene/GC, forming well-defined 3D-MoS2. The obtained MoS2 had massive active edge sites, which provided a strong synergistic effect for HER by improving the electron transfer rate. Li et al. [38] employed a hydrothermal process with the surfactant polyethylene glycol (PEG) to synthesize MoS2 samples with monodisperse spherical morphologies. This MoS2-PEG sample had a minimum particle diameter of 250 nm with a larger interlayer spacing, shorter plates, and fewer stacked layers, resulting in a higher catalytic activity. Bhimanapati et al. [39] used CVD to deposit MoS2 powder onto high-quality graphite paper and silicon substrates to obtain flower-like MoS2 structures.

Recently, Van Nguyen et al. [40] used (NH4)6Mo7O24·4H2O and thioacetamide (TAA, CH3CSNH2) with heating and annealing in the presence of ammonia and hydrochloric acid, respectively, to obtain MoS2 nanoflowers (MoS2 NFs) and ultrasmall MoS2 nanoflowers (MoS2 SNFs). The step of generating intermediates was controlled by varying the temperature and the concentration of H to obtain nanoflowers of different sizes. Specifically, the sizes of the obtained MoS2 SNFs were only 50–90 nm, which were significantly smaller than conventional MoS2 nanoflowers (900–1500 nm). The decreased sizes exposed more active edge sites. In addition, N-doping enhanced the HER kinetics.

The structural modifications above were designed to increase the densities of edge sites by reducing the thicknesses and sizes of the particles, while the mesoporous structure was mainly intended to enhance the catalytic activity by increasing the ratio of the edges to the basal planes and expose more active sites. Kibsgaard et al. [41] used the previously reported double-gyroid (DG) mesostructure of Pt and SiO2 and synthesized the first fully contiguous large-area mesoporous MoS2 thin film (Fig.6(b)). Compared to traditional 2D mesoporous structures, three-dimensional (3D) mesoporous structures have higher structural stabilities and diffusion rates that are an order of magnitude faster than those of other mesoporous structures. Deng et al. [42] reported a well-defined mesoporous MoS2 foam (mPF-MoS2), as shown in Fig.6(c). The mPF-MoS2 was obtained by attaching Mo precursor to SiO2 nanospheres and reacting with CS2 on the surface to form MoS2, which were self-assembled into vertically aligned structures, followed by SiO2 template etching using hydrofluoric acid solution. The MoS2 nanosheets oriented and grown vertically around the mesopores, providing many edges and active sites. A distinct mesopore can be observed in the SEM image of mPF-MoS2 (Fig.6(d)), while the images of Fig.6(f)–Fig.6(i) reveal the presence of abundant spherical voids resulting from the removal of SiO2 nanospheres, during the formation of a uniform porous framework. Many uniform mesopores in mPF-MoS2 facilitated mass transport and delivery of H3O+ and H2. Moreover, the curved surfaces of MoS2 with abundant mesopores in a 2D plane induced strain, which enhanced the electrocatalytic activity. Notably, mPF-MoS2 with uniform mesopores exhibited a significantly improved HER performance when compared to randomly oriented MoS2 nanosheets.

Among the many nanostructures studied, quantum dots (QDs) have numerous advantages. Ren et al. [44] synthesized MoS2 QDs from sodium molybdate and dibenzyl disulfides raw materials by a hydrothermal approach. These QDs exhibited zero-dimensional structures with high specific surface areas and unique defect-rich configurations, which provided abundant active edge sites to enhance the efficiency of the HER. Additionally, the monolayer structure of the MoS2 QDs enabled smoother electron transfer on the basal plane by eliminating interlayer barriers in the vertical direction. As a result, these MoS2 QDs demonstrated a remarkable electrocatalytic activity, making them highly promising electrocatalysts for hydrogen production. Vikraman et al. [45] synthesized MoS2 directly via chemical bath deposition with (NH4)6Mo7O24·4H2O and CH4N2S as the Mo and S sources, respectively (with the experimental setup shown in Fig.7(a)). This approach provided control of the thicknesses of the MoS2 crystal layers while maintaining large-area uniformity on different substrates. Along with nucleation and growth of the MoS2 crystals, the resulting MoS2 layers exhibited the properties of QDs, including edge effects and improved catalytic performance. SEM images (Fig.7(b)–Fig.7(m)) revealed that the MoS2 layers were composed of crystalline QDs. After 2 min, the deposited MoS2 exhibited a relatively smooth surface, and with increasing deposition time, larger grains uniformly grew on the surface of the MoS2. The growth of MoS2 crystal QDs required uniform nucleation, with larger agglomerates appearing as the layer thickness increased. It has been demonstrated that the thicknesses of MoS2 crystal QD films can be adjusted, and uniformity can be maintained.

Morphological design exposes more edge activity, which is important for improving the hydrogen evolution activity of MoS2 by expanding the exposure of surface atoms while increasing atom utilization (see Tab.1 for electrochemical properties of typical catalysts). A controlled morphology enables fast kinetics for the hydrogen adsorption/desorption process. However, constructing edge-enriched MoS2 is challenging due to the thermodynamic disadvantages of edge sites relative to the basal plane. Enhanced active site exposure originating from morphological design can provide a higher catalytic HER activity. However, morphology design and optimization of MoS2 does little to enhance the intrinsic activity and accelerate electron transfer from the substrate surface. The controlled morphology of MoS2 should be used in conjunction with other modification methods to improve the overall catalytic HER activity and conductivity.

3.3 Defect engineering

Defect engineering is highly effective in creating active sites in 2D MoS2. Defect engineering involves creating defects on MoS2 to manipulate the microstructure. The generation of defects increases the number of active sites, modifies the coordination environments around the active centers, and optimizes the electronic structure. Defect engineering strategies can be divided into edge defects, atomic vacancy defects, and doping defects; the first two are intrinsic defects and the latter describes extrinsic defects.

Intrinsic defects are those arising from the incomplete structure of a crystal itself, without any foreign atoms doped (SEM images of common intrinsic defects are shown in Fig.8(a)‒Fig.8(e) [46]). These defects are usually classified into atomic vacancies and edge defects. Atomic vacancies are usually S atom vacancies in the MoS2 substrate, while edge vacancies are mainly caused by the rich boundaries and pore structure in 2H-MoS2. The process involves the introduction of vacancies on the substrate, flexible adjustment of the local catalytic environments surrounding the adjacent atoms, formation of unsaturated coordination states, and optimization of hydrogen adsorption/desorption, all of which fundamentally optimize the HER activity [47]. While both metallic and nonmetallic sites positioned at the edges of MoS2 exhibit catalytic activity, higher proportions of edge defects enhance the exposure of these active sites to obtain a better catalytic activity.

The catalytic activities of the metal sites in the basal plane of MoS2 are often hindered by the presence of nonmetallic sulfur atoms covering these sites, resulting in a limited catalytic activity. DFT calculations revealed that the S vacancies (Sv) present in the basal plane of MoS2 migrated along the empty conduction band previously occupied by Sv-related S 3p and Mo 4d orbitals. The primary factor contributing to the HER activity of MoS2 defects is the alignment of energy levels between the empty conduction band and the reduction of H+ ions [48]. The presence of Sv in MoS2 has been shown to enhance the kinetics of HER by increasing the exposure of unsaturated S−Mo bonds. This exposure optimizes the ΔGH* for the surface of MoS2. At an optimal ΔGH* value of 0 eV, hydrogen exhibits equilibrium binding that is neither too strong nor too weak. Li et al. [49] used DFT calculations (Fig.8(f) depicts the activation of the basal plane) to show that the introduction of 3.12% Sv (% represents the percentage of S atoms removed) provided stability, and the exposed Mo atoms provide sites for hydrogen bonding, resulting in a decrease in ΔGH* to 0.18 eV (Fig.8(g)). The DFT calculations also showed that the valence band of MoS2 shifted downward when different 3d transition metal (3d-TM) dopants replaced the unsaturated Mo atoms surrounding Sv sites (Fig.8(h)–8(i)) [50]. This indicated that coupling between the inert plane S defects and the surrounding 3d-TM dopants activated the MoS2 basal plane, thereby enhancing the electrical conductivity and electron mobility.

Li et al. [49] obtained MoS2 with abundant S vacancies and an excellent catalytic activity via treatment with an Ar plasma. Xu et al. [51] designed and synthesized a Frenkel-defect monolayer MoS2 (FD-MoS2) catalyst by annealing monolayers of MoS2 in an Ar plasma and generating intrinsic defects, as described in Fig.9(a). FD-MoS2 exhibited a phenomenon in which some of the Mo atoms spontaneously left their lattice sites, which created vacancies, and the atoms lodged into adjacent interstitial sites. This resulted in a distinct surface charge distribution for the MoS2, which enhanced the adsorption of H atoms at the vacancies. An increased proportion of Frenkel defects does not always lead to continuous improvement in HER performance. Monolayer MoS2 is easily damaged and fragmented into small islands during bombardment with the Ar plasma. Furthermore, a strong Ar plasma can disrupt the bonding of MoS2, resulting in small fragments and a loss of structural integrity. This indicates that strong ions may not be ideal for generating Sv sites on MoS2. Cheng et al. [52] found that by treating monolayer MoS2 with an H2 plasma, the atomic ratio of Mo and S was changed from 1:1.97 to 1:1.20. This provided a wider range of atomic ratios than defect generation with an Ar plasma or O2 plasma. They successfully induced the formation of Sv sites on the basal planes of MoS2 monolayers with a remote H2 plasma (Fig.9(b)) and investigated the catalytic performance with different vacancy densities. Fig.9(c) demonstrates that the structures of the monolayer MoS2 nanosheets remained intact and unaffected by treatment with the H2 plasma, i.e., there was no fragmentation. Compared with the initial MoS2, the MoS2 treated with the H2 plasma for 15 min showed a 544 mV decrease in the overpotential at 10 mA/cm2 (Fig.9(d)), as well as a significantly decreased Tafel slope of 77.6 mV/dec (Fig.9(e)). In other performance tests, the MoS2 monolayers treated with the H2 plasma also showed a good performance and a high stability for HER in acidic environments (the electrochemical performance being illustrated in Fig.9(f) and Fig.9(g)).

The introduction of S vacancies on MoS2 substrates with plasma or annealing treatments usually requires high temperature/pressure synthetic conditions, which is not energy efficient. The use of mild electrochemical desulfurization has been effective in generating S vacancies on MoS2 substrates. Tsai et al. [53] showed with DFT calculations that the S vacancies in the basal plane were thermodynamically favored at a strong reducing potential. S vacancies were formed in the basal plane by hydrogenation of S atoms and subsequent elimination as H2S gas. Therefore, the defect concentration could be controlled by changing the desulfurization potential. The active sites remained stable over extended periods of desulfurization. Laser ablation in liquid (LAL) is another environmentally friendly technique for generating very high pressures and temperatures at the moment of laser ablation and rapid quenching between the laser pulses for defect formation. Meng et al. [54] synthesized pristine MoS2 (P-MoS2) through a hydrothermal method, dispersed the P-MoS2 in ethanol under ultrasonic treatment, and then obtained laser-treated MoS2 (L-MoS2) through pulsed laser irradiation. S vacancies were generated on the basal plane of L-MoS2 after the laser treatment, and there were only slight deformations.

Air oxidation etching is commonly used for introducing active defects, which results in the formation of well-oriented triangular pits, also known as antidots, on the surfaces of MoS2 samples. These pits improve the catalytic performance of bulk MoS2. Lv et al. [55] oxidized and etched MoS2 samples by heating them with air in a mini-CVD furnace. To investigate the microscopic oxidation etching process of 2D MoS2 at the atomic level, the researchers used the TEM etching technology, heated the furnace to 300 °C within 10 min and maintained this temperature for 5 min for etching. With ADF-STEM images, they analyzed the types and distributions of the residual edge structures at different positions to understand the mechanism for oxidation etching. Their results revealed a kinetic pathway for oxidation etching of 2D MoS2, in which the process was initiated from the edges of the material, as well as from the atomic defects. The edge sites of MoS2 were accompanied by intrinsic structural defects. By increasing the exposure of these edge active sites, both the density of edge defects and the electrocatalytic performance for the HER were improved. Xie et al. [56] achieved this by inducing controlled cracks on the nanosheet surface to expose active edge sites.

The introduction of atomic vacancy defects optimizes the local electronic configuration of the MoS2 planes, and the introduction of edge defects generates more active sites, both of which modulate the intrinsic properties of MoS2 to produce more active centers. This results in synergistic modulation of the vacancy concentration and distribution and thus provides significant improvement in the catalytic performance. However, intrinsic defects, due to the presence of atomic vacancies and edge defects, make it difficult to distinguish their individual roles in catalytic behavior. Moreover, the defect structures are less stable, as they are subject to remodeling in the presence of strong acid and base. The introduction of defects into bulk MoS2 by dopant atoms is discussed in the next section.

3.4 Heteroatom doping

Heteroatom doping is a simple method used to enhance the catalytic activity of MoS2. It involves embedding dopant atoms into the MoS2 structure to create a uniform and stable composite catalyst. The primary purpose is to modulate and optimize the catalytic performance by combining the 2D structure of the basal plane with the electronic structures of the doped atoms. The doped atoms are firmly incorporated into the MoS2 nanosheets through crystal alignment or strong covalent bonds. Because they remain coordinatively unsaturated, they enhance the electronic structure of the MoS2 surface, increase the intrinsic activities of individual catalytic sites, provide more active sites for catalytic reactions, and reduce the ΔGH*. The changes occurring in chemical bonding during the doping process are essential for optimizing the performance of MoS2. Atomic doping can be divided into two main types: metal (cation) doping and nonmetal (anion) doping [57].

MoS2 has a very high specific surface area and an excellent chemical stability and can be used as a support for Pt-based catalysts. According to both DFT and experimental results [58], strong bonds are formed between S and Pt atoms. This strong electronic coupling between the S and Pt atoms lowers the d-band center of Pt, which facilitates electron transport during HER [59]. Therefore, MoS2 with a large specific surface area and an excellent stability can be used as a support for Pt-based catalysts. The contact interface between the Pt and MoS2 is thus a highly active catalytic site for HER. Shan et al. [60] employed a wet chemical method with sodium borohydride (NaBH4) as the reducing agent to synthesize a series of MoS2 catalysts with different Pt particle coverages and investigated their catalytic properties. In addition, it was reported that modifications of Pt nanoparticles led to lattice strain and defects in MoS2 so that the Pt−S bonds increased the specific surface area, prevented the agglomeration of Pt, and improved the activity and stability.

By introducing single atoms into the catalyst, the electronic structures of the active sites can be strained. Inspired by this approach, Jiang et al. [61] used CVD and chemical etching to fabricate nanoporous MoS2 with a bicontinuous structure, which was termed np-MoS2. Considering the remarkable performance of Ru in HER, they anchored single Ru atoms on MoS2, which was denoted as Ru/np-MoS2, and investigated the synergistic effect between the Ru sites and Sv. Based on a previous report [62] that explored doping with isolated metal atoms to activate the MoS2 substrate, the researchers used theoretical calculations and proposed that the introduction of individual Ru atoms into MoS2 could lead to displacement of the surrounding S atoms. This process would be accompanied by a phase transformation, resulting in the formation of Ru/1T-MoS2. Theoretical results indicated that the incorporation of Ru atoms replaced the Mo atoms, demonstrating the potential of Ru doping in generating Sv sites. DFT has been used to study the synergistic effects of Ru atom doping and the tensile strain on MoS2 (Fig.10(a)). It was observed that the external strain increased the accumulation of OH and H2O within the Sv sites, thereby enhancing the synergistic effect between the sulfur vacancies and Ru sites (Fig.10(b)). Fig.10(c) shows a greater degree of orbital overlap between the Mo 3d orbitals and Oads 2p orbitals below the Fermi level, while the bonding energy between *OH2 and M oSv was stronger, and ΔGH* (Fig.10(d)) was also closer to 0 for Ru/1T-MoS2. Consequently, HER was accelerated at the Ru sites. The resulting catalyst exhibited an exceptional electrochemical performance in alkaline environments for HER.

The two dopants mentioned above were noble metals. In addition to noble metal doping, Shi et al. [63] used non-noble metals (Fe, Co, Ni, Cu, and Zn) and doped MoS2 via a one-pot hydrothermal method. The heteroatoms reduced the hydrogen adsorption barrier and enhanced the HER catalytic activity of MoS2. In their study of atom-doped catalysts, they found that doping of Fe or Co into MoS2 did not improve the catalytic performance. The improvement in HER performance by Ni-doping was found to be negligible. Interestingly, Cu and Zn doping of MoS2 provided improved catalytic activity for HER. The different doping amounts of the Cu and Zn atoms impacted the electrocatalytic performance in HER. For example, 1-fold was the optimal doping amount for Zn−MoS2, while Cu-MoS2 exhibited a better activity when the doping amount was 2-fold. Among them, Zn exhibited the best activity with 1-fold doping; η = 200 mV, and the current density of MoS2 increased by 13 times, indicating a fast electron transfer. The starting potential was −0.13 V vs. RHE, and the Tafel slope was reduced from 101 to 51 mV/dec. Li et al. [64] used a simple one-step hydrothermal method to dope vanadium (V) into MoS2 nanosheets grown vertically on carbon paper. The MoS2 samples doped with 5% vanadium showed the ΔGH* values closest to 0. Vanadium doping increased the number of active sites while causing the MoS2 2H → 1T phase transition. Both doping and the phase transition sharply reduced the resistance, significantly increased the conductivity of the samples, promoted electron transfer, and improved the intrinsic activity of MoS2 while enhancing the HER performance.

In addition to doping MoS2 with various metals, some researchers also reported nonmetallic doping (such as O, C, P, N, F, S, and Cl) to optimize ΔGH* and form noncrystalline structures or lattice distortions. This approach exposed more active centers. Nonmetallic doping changed the electronegativity and produced distorted bond lengths and angles, resulting in a redistribution of charge density at the doped sites. Oxygen incorporation into MoS2 nanosheets provided more unsaturated S atoms as active sites, which facilitated charge transfer during electrocatalysis [65]. Yang et al. [66] used vertical growth and fabricated a nanosheet array of etched O-MoS2 on CC with NH4F etching. The resulting nanosheets exhibited sharp edges and high crystallinities due to their growth on CC. The NH4F etching process introduced disordered structures into the O-MoS2 nanosheets, resulting in an abundance of unsaturated S atoms that served as active sites for electrocatalysis and reduced the charge transfer resistance. The number of unsaturated S atoms obtained in the resulting samples was closely related to the etching temperature.

In conclusion, the introduction of heteroatoms regulates the electronic structure of MoS2 to obtain ΔGH*0. However, agglomeration often occurs during introduction of the metal atoms, which reduces atomic utilization, and the S vacancies provide doping sites for the atoms. Therefore, some researchers created S vacancies in MoS2 prior to, or at the same time as, the introduction of metal atoms to improve atomic utilization. Although there have been many reports of doping MoS2 with heteroatoms, the synergistic effects of diatoms or even triatoms should be further investigated.

3.5 Heterostructure construction

Due to the synergistic effect of the interface, heterostructures exhibit an enhanced electrochemical performance compared to single-component structures. Abundant interfaces between different components can be introduced into mixed catalysts, thereby optimizing the electronic configuration and improving the electrochemical activity [67].

Through synergistic interactions in multiscale electronic structures, the intrinsic activity of each active site and the density of active sites are enhanced, leading to an improved electrochemical accessibility. Theoretical findings suggested that Ru doping activated the inert basal planes of MoS2, enhancing processes such as water adsorption/dissociation and hydrogen adsorption/desorption.

Zhang et al. [68] designed a catalyst with tunable compositions and novel core-shell structures for Ru-doped MoS2 nanosheets, in which the outer shell surrounded multiwalled carbon nanotubes (CNTs) called Ru MoS2/CNTs. The catalyst was characterized by small layers of Ru-doped MoS2 nanosheets tightly wrapped around the CNTs. Theoretical calculations suggested that stable doping of Ru atoms into 2H-MoS2 could be achieved by substituting Mo atoms and coordinating with 6 adjacent S atoms in the substrate. However, due to the low intrinsic reactivities of the Mo edge sites, the effect of Ru edge doping on the HER was minimal [69]. The experimental results suggested that the core-shell structure of Ru-MoS2/CNTs reduced the energy barriers and enhanced mass transfer. The optimal HER activity was achieved with Ru MoS2/CNT and an Ru percentage of 5% (mass fraction).

Nguyen et al. [70] discovered a non-noble metal-doped heterostructure. This structure consisted of Co and Nb incorporated into hierarchical MoS2 ultrathin nanosheets that were then hydrothermal (HT) grown directly on micro-TiO2 hollow spheres (referred to as Co, Nb-MoS2/TiO2 HSs). The specific steps are shown in Fig.11. The surface of the spherical hollow structure was coated with ultrathin and mesoporous MoS2 nanosheets, which increased the contact area of the electrolyte, enhanced diffusion of the electrolyte, and improved the charge transfer capability. The dual doping with Co and Nb changed the electronic structure of the MoS2 host, which enhanced the catalytic performance and increased the electrochemically active surface area (ECSA). Polystyrene microspheres (PS) and titanium dioxide were used to create core-shell microspheres (PS@TiO2) as substrates for deposition. The surfaces of the spheres were rough and provided many nanovoids that served as chambers for loading additional MoS2 nanosheets. This configuration exposed a greater number of active sites at the periphery [71]. Additionally, electrons were easily transferred from TiO2 to the outside Co, Nb-MoS2 NSs, thus providing a good catalytic performance in HER.

Shah et al. [72] combined metal- and nonmetal-containing MoS2 nanosheets with extended interlayer spacings, which were grown vertically on N-doped nickel-carbon (Ni@NC) substrates. This resulted in the formation of Ni@NC@MoS2 submicrospheres. This unique heterostructure exhibited an excellent catalytic performance for HER and a low Tafel slope of 47.5 mV/dec. Furthermore, the Ni@NC@MoS2 catalysts showed an improved HER catalytic activity even after 3000 cycles in acidic conditions.

In addition to metal-doped heterostructures, nonmetal-doped heterostructures are also available. Wang et al. [73] synthesized a porous carbon network confining ultrasmall nanocrystals of N-doped MoS2 (N-MoS2/CN) via a self-templating strategy (Fig.12(a) and 12(j)). The SEM images (Fig.12(b)) show the distinct porous nature of the MoS2/g-C3N4 composite, in contrast to the comparatively bulky structure observed for the g-C3N4 bulk material [74]. The type-II isotherm of N-MoS2/CN with a hysteresis loop is shown in Fig.12(c), indicating a mesoporous structure [75]. The X-ray diffraction (XRD) pattern (Fig.12(d)) for N-MoS2/CN exhibited broad peaks with low intensities, indicating the presence of ultrasmall crystallites of MoS2 and defects induced by the N dopants in the MoS2 structure. As shown in Fig.12(e), the hierarchical interconnected structure of N-MoS2/CN was similar to that of a loofah sponge and exhibited a mesoporous morphology. The construction of N-doped edge sites and heterostructures resulted in an overpotential of 114 mV at 10 mA/cm2 for N-MoS2/CN (Fig.12(f)) and a low Tafel slope of 46.8 mV/dec (Fig.12(g)), suggesting a fast Heyrovsky-dominated Volmer-Heyrovsky mechanism [76]. The strong electronic coupling between the N-MoS2 nanocrystals and the carbon matrix made a great contribution to this improved activity. By extrapolation from the Tafel diagram, the exchange current density for N-MoS2/CN was 140 μA/cm2, which was markedly higher than those of MoS2 (11 μA/cm2), MoS2/CS (1.6 μA) and MoS2/CS (1.6 μA). This catalyst also presents a promising stability (Fig.12(h)).

In heterostructure construction, the bonding interactions at the interfaces between different components often increase the rate of electron transfer. The conductivity, hydrophilicity, chemical stability, and active site density of the heterostructure can be modulated by bonding with different materials. The different energy band structures of different phases enable charge transfer at the interface, which facilitates electron modulation of the heterostructure surfaces. Efficient construction of heterostructures, especially those that combine high specific surface areas with fast mass transfer rates, remains a significant challenge.

4 Summary and outlook

4.1 Summary

The application of MoS2 as an old but promising electrocatalyst toward HER has been widely explored. Even the most recent research works are still focusing on MoS2 [7781]. The continuous interest on this material is due to its tailorability, which is highly important to replace the state-of-the-art Pt/C HER catalyst.

To obtain well-defined MoS2 catalyst, researchers have proposed a series of modification strategies to improve the intrinsic HER activity of MoS2, including phase engineering, morphology design, defect engineering, heteroatom doping, and heterostructure construction. These strategies are efficient to control the active sites on MoS2. For example, the 1T phase of MoS2 is preferred due to its high active site density and fast charge transfer. Conventional synthesis methods lead to low yield of exfoliated 1T-MoS2. Precise phase engineering, e.g., the “top–down” and “bottom–up” strategies as discussed in this paper, is thus important to adjust the electronic structure and obtain a stable 1T phase. Besides, the morphology control is another nonnegligible factor, such as high ratio of edge to plane surface to obtain more available active sites, size reduce to facilitate mass transfer. In addition, atomic vacancy defects can enhance the catalytic performance by modulating the electronic structure and thus altering the local catalytic environment of the surrounding neighboring atoms. Edge defects can enhance catalytic performance by promoting the exposure of edge active sites. Among the various methods of defect engineering, the type of defects is hard to be controlled, so that the effect of specific types of defects on catalytic performance is still not clear. Therefore, methods for preparing a single type of defect need to be explored. The introduction of heteroatoms induced a Fermi energy level shift of MoS2, which introduced excess electrons and holes into MoS2 and increased more active sites. Moreover, the dopant atoms interact with the electrons on the surface of MoS2, changing the MoS2 electron transfer process and thus altering the distribution of the local charges at active sites. Although various heteroatom doping in MoS2 has been investigated, it is still a great challenge to synthesize homogeneously heteroatom doped MoS2 catalysts. Heterostructures improve the activity and durability of electrocatalysts by compositing different components and achieve the redistribution of electrons at the interface to play a cooperativity role, and by modulating the composition and crystalline phases of the materials so that the different interfacial components form a so-called coupling effect. However, the coupling effects between different components are still poorly understood. More work is still needed to investigate the coupling effect of different interfacial components in heterostructures.

4.2 Outlook

Although MoS2 has been intensively and extensively explored since it was discovered, it still attracts wide interest due to its unique properties, particularly toward HER [82]. To promote the application of MoS2 catalyst in HER, there are some perspectives to be noted.

Among the various preparation methods for MoS2-based catalysts, one of the major difficulties is that the existing preparation methods are unable to precisely regulate the structure of active sites of MoS2-based catalysts. Therefore, some improved manufacturing techniques are expected to be used to achieve precise control of the material structure, for example, atomic layer deposition, metalorganic CVD and wet chemical methods of colloidal synthesis.

Significant progress has been made in the modification of MoS2 in recent years, but the recognition of active sites during the electrochemical process as well as their evolution are still unclear, which poses barriers for further activity and stability optimization. The in-situ characterization technique is a good solution to overcome these difficulties. It can not only understand the surface morphology and structure of initial catalysts, but also observe the catalytic active sites under the working conditions and monitor the dynamic change. For example, in-situ X-ray photoelectron spectroscopy (XPS) and in-situ X-ray absorption spectroscopy (XAS) can be used to monitor the electronic state, determine the number of active sites, and judge the initial and intermediate states of a reaction in time. These in-situ characterization techniques can guide the synthesis of highly active and stable MoS2 catalysts by exploring the relationship between the structure and properties of different modified MoS2.

Theoretical calculations are generally used to study the electronic properties of catalysts, electron transfer energy barriers, catalyst activity and stability by calculating the energy band structure of catalysts, adsorption energy, differential charge density and energy profile for the HER process. The minimum energy path during hydrogen evolution can be modeled by calculations. The results obtained experimentally can be interpreted theoretically by calculations. By combining theoretical calculations with practical experimental results, it is possible to gain insights into the underlying reaction mechanisms and degradation mechanisms, which can guide future research and development of MoS2 with excellent properties.

In summary, the modification of MoS2 has been greatly developed, showing that it indeed has a great potential for replacing noble metal toward HER. Although MoS2 is still in its early stages of industrialization, it has the potential to be implemented in membrane-electrode assemblies as an alternative to noble metal catalysts.

References

[1]

Wang H, Wu Y, Feng M. . Visible-light-driven removal of tetracycline antibiotics and reclamation of hydrogen energy from natural water matrices and wastewater by polymeric carbon nitride foam. Water Research, 2018, 144: 215–225

[2]

Sprick R S, Chen Z, Cowan A J. . Water oxidation with cobalt-loaded linear conjugated polymer photocatalysts. Angewandte Chemie International Edition, 2020, 59(42): 18695–18700

[3]

Xu Y, Mao N, Zhang C. . Rational design of donor-π-acceptor conjugated microporous polymers for photocatalytic hydrogen production. Applied Catalysis B: Environmental, 2018, 228: 1–9

[4]

Wang H, Qian C, Liu J. . Integrating suitable linkage of covalent organic frameworks into covalently bridged inorganic/organic hybrids toward efficient photocatalysis. Journal of the American Chemical Society, 2020, 142(10): 4862–4871

[5]

Xie X, Du L, Yan L. . Oxygen evolution reaction in alkaline environment: Material challenges and solutions. Advanced Functional Materials, 2022, 32(21): 2110036

[6]

Du L, Xing L, Zhang G. . Strategies for engineering high-performance PGM-free catalysts toward oxygen reduction and evolution reactions. Small Methods, 2020, 4(6): 2000016

[7]

Wei Z X, Zhu Y T, Liu J Y. . Recent advance in single-atom catalysis. Rare Metals, 2021, 40(4): 767–789

[8]

Yuan F H, Mohammadi M R, Ma L L. . Electrodeposition of self-supported NiMo amorphous coating as an efficient and stable catalyst for hydrogen evolution reaction. Rare Metals, 2022, 41(8): 2624–2632

[9]

Zou X, Zhang Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chemical Society Reviews, 2015, 44(15): 5148–5180

[10]

Liang T, Wang A, Ma D. . Low-dimensional transition metal sulfide-based electrocatalysts for water electrolysis: Overview and perspectives. Nanoscale, 2022, 14(48): 17841–17861

[11]

Hinnemann B, Moses P G, Bonde J. . Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. Journal of the American Chemical Society, 2005, 127(15): 5308–5309

[12]

Jaramillo T F, Jørgensen K P, Bonde J. . Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science, 2007, 317(5834): 100–102

[13]

Wang Q, Lei Y, Wang Y. . Atomic-scale engineering of chemical-vapor-deposition-grown 2D transition metal dichalcogenides for electrocatalysis. Energy & Environmental Science, 2020, 13(6): 1593–1616

[14]

Chhowalla M, Shin H S, Eda G. . The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nature Chemistry, 2013, 5(4): 263–275

[15]

Chen J, Walker W R, Xu L. . Intrinsic capacitance of molybdenum disulfide. ACS Nano, 2020, 14(5): 5636–5648

[16]

Huang X, Zeng Z, Zhang H. Metal dichalcogenide nanosheets: Preparation, properties and applications. Chemical Society Reviews, 2013, 42(5): 1934–1946

[17]

Lei Z, Zhan J, Tang L. . Recent development of metallic (1T) phase of molybdenum disulfide for energy conversion and storage. Advanced Energy Materials, 2018, 8(19): 1703482

[18]

Zhang J, Wang T, Liu P. . Engineering water dissociation sites in MoS2 nanosheets for accelerated electrocatalytic hydrogen production. Energy & Environmental Science, 2016, 9(9): 2789–2793

[19]

Martis J, Susarla S, Rayabharam A. . Imaging the electron charge density in monolayer MoS2 at the Ångstrom scale. Nature Communications, 2023, 14(1): 4363

[20]

Voiry D, Mohite A, Chhowalla M. Phase engineering of transition metal dichalcogenides. Chemical Society Reviews, 2015, 44(9): 2702–2712

[21]

Li S, Sun J, Guan J. Strategies to improve electrocatalytic and photocatalytic performance of two-dimensional materials for hydrogen evolution reaction. Chinese Journal of Catalysis, 2021, 42(4): 511–556

[22]

Zhao W, Pan J, Fang Y. . Metastable MoS2: Crystal structure, electronic band structure, synthetic approach and intriguing physical properties. Chemistry—A European Journal, 2018, 24(60): 15942–15954

[23]

Lin Y C, Dumcenco D O, Huang Y S. . Atomic mechanism of the semiconducting-to-metallic phase transition in single-layered MoS2. Nature Nanotechnology, 2014, 9(5): 391–396

[24]

Liu Z, Wang K, Li Y. . Activation engineering on metallic 1T-MoS2 by constructing in-plane heterostructure for efficient hydrogen generation. Applied Catalysis B: Environmental, 2022, 300: 120696

[25]

Wang D, Li J, Ma H. . Layer-structure adjustable MoS2 catalysts for the slurry-phase hydrogenation of polycyclic aromatic hydrocarbons. Journal of Energy Chemistry, 2021, 63: 294–304

[26]

Liu B, Xu W, Long X. . Atomic mechanism of lithium intercalation induced phase transition in layered MoS2. Physical Chemistry Chemical Physics, 2022, 24(31): 18777–18782

[27]

Tan C, Luo Z, Chaturvedi A. . Preparation of high-percentage 1T-phase transition metal dichalcogenide nanodots for electrochemical hydrogen evolution. Advanced Materials, 2018, 30(9): 1705509

[28]

Chen Z, Leng K, Zhao X. . Interface confined hydrogen evolution reaction in zero valent metal nanoparticles-intercalated molybdenum disulfide. Nature Communications, 2017, 8(1): 14548

[29]

Dong L, Yang J, Yue X. . The effects of the fluence of electron irradiation on the structure and hydrogen evolution reaction performance of molybdenum disulfide. Journal of Materials Chemistry C, 2022, 10(20): 7839–7848

[30]

Zhu J, Wang Z, Yu H. . Argon plasma induced phase transition in monolayer MoS2. Journal of the American Chemical Society, 2017, 139(30): 10216–10219

[31]

Sun Y, Zang Y, Tian W. . Plasma-induced large-area N, Pt-doping and phase engineering of MoS2 nanosheets for alkaline hydrogen evolution. Energy & Environmental Science, 2022, 15(3): 1201–1210

[32]

Mai H D, Jeong S, Bae G N. . Solvothermal temperature-control of active 1T phase in carbon cloth-supported MoS2 and Pt-Ni cluster electrodeposition for hydrogen evolution reaction. Journal of Alloys and Compounds, 2023, 942: 169035

[33]

Zhu L, Wang Z, Li C. . Highly stable 1T-MoS2 by magneto-hydrothermal synthesis with Ru modification for efficient hydrogen evolution reaction. Journal of Materials Chemistry. A, 2022, 10(39): 21013–21020

[34]

Zhang H, Xu H, Wang L. . A metal-organic frameworks derived 1T-MoS2 with expanded layer spacing for enhanced electrocatalytic hydrogen evolution. Small, 2023, 19(4): 2205736

[35]

Jiang Y, Li X, Yu S. . Reduced graphene oxide-modified carbon nanotube/polyimide film supported MoS2 nanoparticles for electrocatalytic hydrogen evolution. Advanced Functional Materials, 2015, 25(18): 2693–2700

[36]

Yi J D, Liu T T, Huang Y B. . Solid-state synthesis of MoS2 nanorod from molybdenum-organic framework for efficient hydrogen evolution reaction. Science China Materials, 2019, 62(7): 965–972

[37]

Behranginia A, Asadi M, Liu C. . Highly efficient hydrogen evolution reaction using crystalline layered three-dimensional molybdenum disulfides grown on graphene film. Chemistry of Materials, 2016, 28(2): 549–555

[38]

Li M, Wang D, Li J. . Surfactant-assisted hydrothermally synthesized MoS2 samples with controllable morphologies and structures for anthracene hydrogenation. Chinese Journal of Catalysis, 2017, 38(3): 597–606

[39]

Bhimanapati G R, Hankins T, Lei Y. . Growth and tunable surface wettability of vertical MoS2 layers for improved hydrogen evolution reactions. ACS Applied Materials & Interfaces, 2016, 8(34): 22190–22195

[40]

Van Nguyen T, Nguyen T P, Van Le Q. . Synthesis of very small molybdenum disulfide nanoflowers for hydrogen evolution reaction. Applied Surface Science, 2023, 607: 154979

[41]

Kibsgaard J, Chen Z, Reinecke B N. . Engineering the surface structure of MoS2 to preferentially expose active edge sites for electrocatalysis. Nature Materials, 2012, 11(11): 963–969

[42]

Deng J, Li H, Wang S. . Multiscale structural and electronic control of molybdenum disulfide foam for highly efficient hydrogen production. Nature Communications, 2017, 8(1): 14430

[43]

Huang H, Chen L, Liu C. . Hierarchically nanostructured MoS2 with rich in-plane edges as a high-performance electrocatalyst for the hydrogen evolution reaction. Journal of Materials Chemistry. A, 2016, 4(38): 14577–14585

[44]

Ren X, Pang L, Zhang Y. . One-step hydrothermal synthesis of monolayer MoS2 quantum dots for highly efficient electrocatalytic hydrogen evolution. Journal of Materials Chemistry. A, 2015, 3(20): 10693–10697

[45]

Vikraman D, Akbar K, Hussain S. . Direct synthesis of thickness-tunable MoS2 quantum dot thin layers: Optical, structural and electrical properties and their application to hydrogen evolution. Nano Energy, 2017, 35: 101–114

[46]

Trainer D J, Nieminen J, Bobba F. . Visualization of defect induced in-gap states in monolayer MoS2. npj 2D Materials and Applications, 2022, 6(1): 13

[47]

Su H, Pan X, Li S. . Defect-engineered two-dimensional transition metal dichalcogenides towards electrocatalytic hydrogen evolution reaction. Carbon Energy, 2023, 5(6): e296

[48]

Lau T H, Lu X, Kulhavý J. . Transition metal atom doping of the basal plane of MoS2 monolayer nanosheets for electrochemical hydrogen evolution. Chemical Science, 2018, 9(21): 4769–4776

[49]

Li H, Tsai C, Koh A L. . Activating and optimizing MoS2 basal planes for hydrogen evolution through the formation of strained sulphur vacancies. Nature Materials, 2016, 15(1): 48–53

[50]

Ma Y, Leng D, Zhang X. . Enhanced activities in alkaline hydrogen and oxygen evolution reactions on MoS2 electrocatalysts by in-plane sulfur defects coupled with transition metal doping. Small, 2022, 18(39): 2203173

[51]

Xu J, Shao G, Tang X. . Frenkel-defected monolayer MoS2 catalysts for efficient hydrogen evolution. Nature Communications, 2022, 13(1): 2193

[52]

Cheng C C, Lu A Y, Tseng C C. . Activating basal-plane catalytic activity of two-dimensional MoS2 monolayer with remote hydrogen plasma. Nano Energy, 2016, 30: 846–852

[53]

Tsai C, Li H, Park S. . Electrochemical generation of sulfur vacancies in the basal plane of MoS2 for hydrogen evolution. Nature Communications, 2017, 8(1): 15113

[54]

Meng C, Lin M C, Du X W. . Molybdenum disulfide modified by laser irradiation for catalyzing hydrogen evolution. ACS Sustainable Chemistry & Engineering, 2019, 7(7): 6999–7003

[55]

Lv D, Wang H, Zhu D. . Atomic process of oxidative etching in monolayer molybdenum disulfide. Science Bulletin, 2017, 62(12): 846–851

[56]

Xie J, Zhang H, Li S. . Defect-rich MoS2 ultrathin nanosheets with additional active edge sites for enhanced electrocatalytic hydrogen evolution. Advanced Materials, 2013, 25(40): 5807–5813

[57]

Chen J, Lin Y, Wang H. . 2D molybdenum compounds for electrocatalytic energy conversion. Advanced Functional Materials, 2023, 33(4): 2210236

[58]

Yang C L, Wang L N, Yin P. . Sulfur-anchoring synthesis of platinum intermetallic nanoparticle catalysts for fuel cells. Science, 2021, 374(6566): 459–464

[59]

Chen I W P, Chen Y X, Wu C W. . Large-scale fabrication of a flexible, highly conductive composite paper based on molybdenum disulfide-Pt nanoparticle-single-walled carbon nanotubes for efficient hydrogen production. Chemical Communications, 2017, 53(2): 380–383

[60]

Shan A, Teng X, Zhang Y. . Interfacial electronic structure modulation of Pt-MoS2 heterostructure for enhancing electrocatalytic hydrogen evolution reaction. Nano Energy, 2022, 94: 106913

[61]

Jiang K, Luo M, Liu Z. . Rational strain engineering of single-atom ruthenium on nanoporous MoS2 for highly efficient hydrogen evolution. Nature Communications, 2021, 12(1): 1687

[62]

Luo Z, Ouyang Y, Zhang H. . Chemically activating MoS2 via spontaneous atomic palladium interfacial doping towards efficient hydrogen evolution. Nature Communications, 2018, 9(1): 2120

[63]

Shi Y, Zhou Y, Yang D R. . Energy level engineering of MoS2 by transition-metal doping for accelerating hydrogen evolution reaction. Journal of the American Chemical Society, 2017, 139(43): 15479–15485

[64]

Li M, Cai B, Tian R. . Vanadium doped 1T MoS2 nanosheets for highly efficient electrocatalytic hydrogen evolution in both acidic and alkaline solutions. Chemical Engineering Journal, 2021, 409: 128158

[65]

Xie J, Zhang J, Li S. . Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution. Journal of the American Chemical Society, 2013, 135(47): 17881–17888

[66]

Yang X, Li X, Wang Y. . Efficient etching of oxygen-incorporated molybdenum disulfide nanosheet arrays for excellent electrocatalytic hydrogen evolution. Applied Surface Science, 2019, 491: 245–255

[67]

Chen H X, Xu H, Song Z R. . Pressure-induced bimetallic carbon nanotubes from metal-organic frameworks as optimized bifunctional electrocatalysts for water splitting. Rare Metals, 2023, 42(1): 155–164

[68]

Zhang X, Zhou F, Zhang S. . Engineering MoS2 basal planes for hydrogen evolution via synergistic ruthenium doping and nanocarbon hybridization. Advanced Science, 2019, 6(10): 1900090

[69]

Wang H, Tsai C, Kong D. . Transition-metal doped edge sites in vertically aligned MoS2 catalysts for enhanced hydrogen evolution. Nano Research, 2015, 8(2): 566–575

[70]

Nguyen D C, Luyen Doan T L, Prabhakaran S. . Hierarchical Co and Nb dual-doped MoS2 nanosheets shelled micro-TiO2 hollow spheres as effective multifunctional electrocatalysts for HER, OER, and ORR. Nano Energy, 2021, 82: 105750

[71]

Pei J, Geng H, Ang E H. . Controlled synthesis of hollow C@TiO2@MoS2 hierarchical nanospheres for high-performance lithium-ion batteries. Nanoscale, 2018, 10(36): 17327–17334

[72]

Shah S A, Shen X, Xie M. . Nickel@nitrogen-doped carbon@MoS2 nanosheets: An efficient electrocatalyst for hydrogen evolution reaction. Small, 2019, 15(9): 1804545

[73]

Wang H, Xiao X, Liu S. . Structural and electronic optimization of MoS2 edges for hydrogen evolution. Journal of the American Chemical Society, 2019, 141(46): 18578–18584

[74]

Yu H, Shang L, Bian T. . Nitrogen-doped porous carbon nanosheets templated from g-C3N4 as metal-free electrocatalysts for efficient oxygen reduction reaction. Advanced Materials, 2016, 28(25): 5080–5086

[75]

Su J, Yang Y, Xia G. . Ruthenium-cobalt nanoalloys encapsulated in nitrogen-doped graphene as active electrocatalysts for producing hydrogen in alkaline media. Nature Communications, 2017, 8(1): 14969

[76]

Li Y, Wang H, Xie L. . MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction. Journal of the American Chemical Society, 2011, 133(19): 7296–7299

[77]

Liu H J, Zhang S, Chai Y M. . Ligand modulation of active sites to promote Co doped 1T-MoS2 electrocatalytic hydrogen evolution in alkaline media. Angewandte Chemie, 2023, 62(48): e202313845

[78]

Han W, Ning J, Long Y. . Unlocking the ultrahigh-current-density hydrogen evolution on 2H-MoS2 via simultaneous structural control across seven orders of magnitude. Advanced Energy Materials, 2023, 13(16): 2300145

[79]

Gao B, Zhao Y, Du X. . Modulating trinary-heterostructure of MoS2 via controllably carbon doping for enhanced electrocatalytic hydrogen evolution reaction. Advanced Functional Materials, 2023, 33(22): 2214085

[80]

Guo X, Song E, Zhao W. . Charge self-regulation in 1T-MoS2 structure with rich S vacancies for enhanced hydrogen evolution activity. Nature Communications, 2022, 13(1): 5954

[81]

Hu B, Huang K, Tang B. . Graphene quantum dot-mediated atom-layer semiconductor electrocatalyst for hydrogen evolution. Nano-Micro Letters, 2023, 15(1): 217

[82]

Shi Z, Zhang X, Lin X. . Phase-dependent growth of Pt on MoS2 for highly efficient H2 evolution. Nature, 2023, 621(7978): 300–305

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (14001KB)

2306

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/