Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Jul 12;8(29):25649–25673. doi: 10.1021/acsomega.3c02084

MoS2-Based Nanocomposites for Photocatalytic Hydrogen Evolution and Carbon Dioxide Reduction

Bhagyalakshmi Balan , Marilyn Mary Xavier , Suresh Mathew †,‡,*
PMCID: PMC10373465  PMID: 37521597

Abstract

graphic file with name ao3c02084_0014.jpg

Photocatalysis is a facile and sustainable approach for energy conversion and environmental remediation by generating solar fuels from water splitting. Due to their two-dimensional (2D) layered structure and excellent physicochemical properties, molybdenum disulfide (MoS2) has been effectively utilized in photocatalytic H2 evolution reaction (HER) and CO2 reduction. The photocatalytic efficiency of MoS2 greatly depends on the active edge sites present in their layered structure. Modifications like reducing the layer numbers, creating defective structures, and adopting different morphologies produce more unsaturated S atoms as active edge sites. Hence, MoS2 acts as a cocatalyst in nanocomposites/heterojunctions to facilitate the photogenerated electron transfer. This review highlights the role of MoS2 as a cocatalyst for nanocomposites in H2 evolution reaction and CO2 reduction. The H2 evolution activity has been described comprehensively as binary (with metal oxide, carbonaceous materials, metal sulfides, and metal–organic frameworks) and ternary composites of MoS2. Photocatalytic CO2 reduction is a more complex and challenging process that demands an efficient light-responsive semiconductor catalyst to tackle the thermodynamic and kinetic factors. Photocatalytic reduction of CO2 using MoS2 is an emerging topic and would be a cost-effective substitute for noble catalysts. Herein, we also exclusively envisioned the possibility of layered MoS2 and its composites in this area. This review is expected to furnish an understanding of the diverse roles of MoS2 in solar fuel generation, thus endorsing an interest in utilizing this unique layered structure to create nanostructures for future energy applications.

1. Introduction

Overusing fossil fuels, growing environmental pollution, and global climate change are the most significant challenges in this century. This crisis has emphasized developing renewable and clean energy sources for future purposes. One of the potential alternatives for this dilemma is solar energy. Semiconductor photocatalysis is a process that effectively converts solar energy to chemical energy. It is considered one of the most widely explored solar-fuel strategies for the photocatalytic splitting of water into H2 and O2 and the photocatalytic reduction of CO2 into hydrocarbon fuels.14 So far, noble metals, viz., Au, Pd, and Pt, are the most efficient catalysts for either H2 evolution or CO2 reduction reactions. However, over the past two decades, significant progress has been achieved in creating an efficient catalyst for H2 evolution and CO2 reduction utilizing non-noble metals, including metal oxides,510 metal sulfides,11,12 chalcogenides,13,14 carbonaceous materials,15,16 and carbides.1719 Among these, the discovery of materials having two-dimensional (2D) structures has been a significant triumph in developing and expanding the energy industry. Primarily, such a journey started with the invention of graphene in 2004 by Andre Geim and Kostya Novoselov.20,21 Their outstanding properties significantly influence energy-related fields, such as solar cells, photo and electrocatalysis, light-emitting diodes (LEDs), transistors, and photodetectors.2227 Like graphene, layered transition metal dichalcogenides (TMDCs) were introduced as another class of 2D materials with intriguing properties that could potentially replace noble metals in catalytic applications. Molybdenum disulfide (MoS2) mainly shows wide applications in various energy-related fields.2830 Most importantly, MoS2-based nanocomposites have developed significantly in photocatalysis, such as pollutant degradation,31,32 H2 production,33,34 and CO2 reduction.35,36 Therefore, we explore the advancement of nanocomposites based on MoS2 in photocatalytic H2 evolution and CO2 reduction.

2. Photocatalytic Hydrogen Evolution Reaction (HER)

2.1. Basic Principles of HER

The overall photocatalytic splitting of H2O to produce H2 follows three basic steps. The first step is the electron–hole pair generation through photon absorption, which occurs inside the photocatalyst. Second-step involves the separation of charge carriers and their transport to the surface of the photocatalyst, and the third step constitutes the subsequent surface reduction of photoexcited electrons to produce hydrogen (Figure 1). Water splitting to produce H2 is a thermodynamically uphill reaction, and the free energy associated with this process is 237.2 kJ mol–1, corresponding to 1.23 eV. Therefore, the photocatalyst has a minimum band gap energy of 1.23 eV for water splitting, and the band gap should be lower than 3 eV for maximum solar energy utilization. Another criterion is related to the positions of the valence band (VB) and conduction band (CB) of the photocatalyst; the conduction band minimum (CBM) of the photocatalyst should be above the water reduction potential, and the valence band minimum (VBM) should be below the water oxidation potential.37 Thus, this limitation of band gap is a significant problem associated with the low yield of H2 production in most photocatalysts.

Figure 1.

Figure 1

Schematic illustration of the sequences of photocatalytic H2O splitting in a photocatalyst: (a) irradiation of light and photon absorption, (b) generation and separation of charge carriers, (c) migration of electron and hole to the surface of the photocatalyst, and (d) oxidation–reduction reactions on the surface to produce H2 and O2.

2.2. Structure and Properties of MoS2

Hydrogen (H2), having an energy density of 140 MJ kg–1, is an ideal solution for future energy and environmental crises.3840 Since the pioneering discovery by Fujishima and Honda in 1972 through the photoelectrochemical splitting of H2O, H2 production via solar energy using semiconductor photocatalysts has received greater attention from researchers.41,42 Further research is ongoing to produce highly active photocatalysts for photocatalytic H2 evolution with high catalytic activity and lower overpotential. The development of nanocomposites of MoS2 for photocatalytic H2 generation is of great interest due to its excellent chemical stability, low cost, nontoxicity, and 2D-layered structure.4345

It is well-known that bulk MoS2 has a bandgap of 1.2 eV and has interestingly been explored as a lubricant, 2D transistor, and hydrodesulfurization catalyst. As shown in Figure 2a, it is composed of vertically stacked, weakly interacting layers held together by van der Waals interactions. In monolayer MoS2, the Mo atoms are sandwiched between two layers of sulfur atoms with a thickness of 0.65 nm. On the basis of the atomic stacking order, MoS2 has three structural polymorphs: 2H (two layers per repeat unit, hexagonal symmetry), 3R (three layers per repeat unit, rhombohedral symmetry), and 1T (one layer per repeat unit, tetragonal symmetry)46,47 (Figure 2b). MoS2 has numerous intriguing properties due to its unique layered structure. Because of the antibonding state created by the interaction between molybdenum dz2 and sulfur pz orbital at the top of the valence band, MoS2 exhibits excellent stability versus photocorrosion in solution. Previous research on bulk MoS2 revealed that it would not be an effective HER catalyst during light illumination.48,49 However, studies have demonstrated that the fabrication of nanostructured MoS2 materials can considerably enhance HER activity, which is driven by several factors. For bulk MoS2, the CB edge potential is positioned at −0.16 V, more than the H+/H2 redox couple (−0.41 V vs SHE, pH 7). As a result, it fails to satisfy the condition of water splitting. Alternatively, the CB edge potential tends to be −0.53 eV in monolayers, indicating the ability to reduce protons to generate H2.50,51 This kind of indirect-to-direct bandgap transition of MoS2 was accomplished by decreasing the layer numbers, which resulted in an increase in bandgap from 1.2 eV (bulk MoS2) to 1.9 eV (monolayer MoS2), eliminating the potential barrier, and MoS2 could be used as appropriate material for the water-splitting reaction.47,52 Hinnemann et al. employed DFT calculations to determine the free energy of hydrogen adsorption (GH* on the MoS2 edge) to find out the active edge sites of MoS2 for the H2 evolution reaction. The results demonstrated that GH* is significantly close to thermoneutrality, suggesting that MoS2 could be an effective HER catalyst. This result was also confirmed experimentally by the electrochemical measurements of MoS2 particles grown on graphite.53 Bonde et al. made similar calculations and determined the free energy of adsorption on MoS2 edges.54 No̷rskov’s group has recently modified these values. On the basis of their findings, the suitable edge configurations correspond to Mo edges covered with 50% S having GH* = 0.06 eV instead of S being covered with 100% S having GH* = −0.45 eV.55 Meanwhile, Jaramillo et al. provided experimental evidence for the role of MoS2 edges in the HER reaction. They systematically produced different sizes of MoS2 nanoclusters and correlated the electrochemical behavior and edge length of MoS2 nanoclusters by STM analysis. The results show that the HER activity was associated linearly with many edge sites on the MoS2 catalyst. Thus, the edge sites of MoS2 were identified as active sites for the HER reaction.56 Hence, the first part of this review focuses on 2D layered MoS2-based nanocomposites as photocatalysts for H2 evolution reactions. It envisages the unique role of MoS2 in increasing the photocatalytic activity of different composites for future energy applications.

Figure 2.

Figure 2

(a) Three-dimensional representation of the structure of a MoS2 monolayer. Reproduced from ref (46). Copyright 2014 American Chemical Society (b) Schematic structural polytypes: including 2H, 3R, and 1T MoS2 polytypes. Reproduced with permission from ref (44). Copyright 2016 Elsevier.

2.3. MoS2-Based Composites for Photocatalytic Hydrogen Evolution Reaction

2.3.1. MoS2/Metal Oxide Composites

The fundamental issues associated with metal oxide as photocatalysts are the ultrafast charge recombination and photocorrosion within the photocatalysts. Various heterostructures have been developed to overcome these issues by combining metal oxides with MoS2 and providing promising, low-cost, efficient catalysts.57,58 The edges of 2D MoS2 nanosheets have more catalytically active sites that show superior photocatalytic activity. This edge site activity can be further enhanced efficiently by suppressing the stacking of MoS2 flakes. This modified edge-enriched ultrathin MoS2 nanosheets uniformly embedded into yolk–shell structured TiO2 exhibited excellent photocatalytic H2 evolution. The composite shows an H2-evolution rate of 2443 μmol g–1 h–1 than pure TiO2 (247 μmol g–1 h–1) with a PL lifetime of 5.17 ns. Since the energy differences between the CB of TiO2 and MoS2 are comparable, the electrons are more easily transferred from the excited TiO2 to MoS2 nanosheets. Because of their higher active sites and enriched edges, ultrathin MoS2 nanoflakes potentially act as an electron acceptor rather than bulk MoS2.59 Using a plasma sputtering method, a thin layer (0.5 to 10 nm) of MoS2 is selectively deposited on the top of anodic anatase TiO2 nanotubes (TiNTs). It has been tested in an open-circuit for photocatalytic H2 evolution reaction. Using solar simulator (100 mW/cm2), a significant increase in H2-evolution can be observed using a nominal 1 nm thick MoS2 on the top of a 6 μm thick TiNTs layer. The band gap positions of the TiNTs/MoS2 structure were tuned from 1.2, 1.3, and 1.8 eV for the deposition of 10, 3, and 1 nm MoS2, respectively. Here, MoS2 serves as a photosensitizer and as an H2 evolution cocatalyst.60

Several studies have shown that engineered defects at the 2D MoS2 nanosheet edges can enhance the photocatalytic H2 evolution efficiency. Metal oxides can be combined with this defect-rich ultrathin form of MoS2 nanosheets to improve light-harvesting efficiency and increase the photocatalytic H2 evolution rate.61 Zhu et al., for example, developed a series of MoS2-TiO2 catalysts for photocatalytic H2-evolution via a simple ball-milling process using MoS2 as a precursor. The recombination of photogenerated charge carriers decreased the photocatalytic performance of pure TiO2. The H2 evolution activity increases considerably with MoS2 loading, reaching a maximum photocatalytic activity of 150.7 μmol h–1 for the optimum 4.0 wt % MoS2-TiO2 catalyst. The too-quick ball-milling process produces significantly more defective MoS2 structures, which is essential to the practical application of this simple, cost-effective, and sustainable method.62 In addition to TiO2, the coupling of MoS2 with other metal oxide photocatalysts has also been investigated. Defect-rich MoS2-nanosheets combined with nitrogen-doped ZnO nanorod composites has recently been developed.63 Under solar light irradiation, an optimized heterostructure consisting of 15 wt % defect-rich MoS2 nanosheets coated on N-ZnO (N-ZnO-MoS2) exhibited the maximum H2 evolution, 17.3 mmol h–1 g–1 (Figure 3a). The effective encapsulation of defect-rich MoS2 nanosheets over N-ZnO nanorods resulted in a defect-induced interfacial contact region, contributing to increased photocatalytic activity. Figure 3b illustrates the pathway of enhanced photocatalytic H2 evolution reaction. Under sunlight irradiation, the electron transfer occurs from the in situ generated ZnS to the conduction band of N-ZnO, followed by the defect-rich MoS2 nanosheets. The defect-rich structure has several active edge sites with many exposed unsaturated “S” atoms that show a significant affinity for H+ ions to produce H2 during this reaction and boost visible light absorption. Using a simple hydrothermal approach, Yuan et al. investigated the effect of the interaction of MoS2-nanosheet with ZnO nanoparticles.64 Noble metals have replaced MoS2 as a cocatalyst for many photocatalytic H2 evolution reactions. For comparison, composites of ZnO with noble metals were also made.

Figure 3.

Figure 3

(a) Volume of H2 generated using the N-ZnO-MoS2 photocatalysts. (b) Schematic illustration of photocatalytic H2 generation on the N-ZnO-MoS2 heterostructure under natural sun light. Reproduced from ref (63). Copyright 2019 American Chemical Society.

Among these, Rh exhibited the maximum photocatalytic H2 evolution rate of 153 μmol h–1 g–1, around one-fifth of the 1 wt % MoS2-ZnO composite photocatalyst (768 μmol g–1 h–1). Hence, layered MoS2 could be used as a substitute for noble metals in photocatalytic H2 evolution reactions. Similarly, a noble metal-free visible light H2 evolution was achieved through Zn-5, 10, 15, 20-tetrakis (4-carboxyphenyl)-porphyrin dye (Zn TCPP) sensitized MoS2/ZnO heterostructures as a photocatalyst. The effect of MoS2 loading on the photocatalytic activities of the Zn TCPP-MoS2/ZnO hybrid system was investigated using triethanolamine (TEOA) aqueous solution as a sacrificial agent. The results suggested that, even without any noble metals, the as-prepared composite loaded with 0.5 wt % MoS2 delivered an H2 evolution rate of 75 μmol h–1 g–1. The amount of MoS2 loaded on the ZnO surface plays an essential role in trapping the conduction band electrons injected by the zinc porphyrin dye (ZnTCPP*), thereby suppressing the recombination of photogenerated charge carriers. This study provides insight into developing noble metal-free visible-light responsive ZnO-based photocatalysts using zinc porphyrin dyes as a light-harvesting material and MoS2 as a cocatalyst.65

A photocatalytic H2 evolution of ß-Tm3+, Yb3+: NaYF4/MoS2-Ta2O5 photocatalyst was reported by Shu et al. MoS2 was deposited onto the surface of ß-Tm3+, Yb3+: NaYF4/Ta2O5 nanocomposite as a cocatalyst. The presence of an up-conversion luminescence agent and MoS2 cocatalyst could significantly increase the photocatalytic H2 evolution rate of the Ta2O5 catalyst. Because of its outstanding electronic properties, MoS2 provides more active sites to accept photogenerated electrons from the conduction band (CB) of Ta2O5 for H2 evolution, reducing electron–hole recombination and increasing photocatalytic H2 evolution activity of ß-Tm3+, Yb3+: NaYF4/MoS2-Ta2O5 catalysts.66 A MoS2/Bi2O3 Z-scheme heterojunction catalyst with rich oxygen vacancies was prepared in another study. The superior photocatalytic activity stemmed from oxygen vacancies that enhance the interfacial interaction between MoS2 and Bi2O3.67 Monolayer metal sulfides are promising for photocatalysis due to their excellent activity and high surface area. Especially their excellent charge carrier mobility shows high promise in photocatalysis.68

For the first time, catalytic H2 production and pollutant degradation catalyzed by MoS2 monolayers loaded on the m-BiVO4 surface were studied using hybrid density functional calculations and long-range dispersion correction methods. Interfacial adhesion energy, binding energy, and equilibrium distance indicate that the association of MoS2 monolayer with m-BiVO4 (010) is a van der Waals interaction, resulting in a high reduction capacity toward H2 generation. With a high optical absorption coefficient, the monolayer of MoS2 serves as a suitable electron acceptor, and more active sites can accept reactive species during the photocatalytic process.69 Semiconductor photocatalysis through p-n heterojunctions is essential for improving future hydrogen storage. Swain et al. discovered a p-n heterojunction between p-type MoS2 and n-type CeO2 in a p-n heterojunction. The suggested p-n heterojunction method provides photoinduced charge transfer and separation. Following optimization of the overall experiment, it was discovered that the 2 wt % MoS2/CeO2 heterojunction nanocomposite has the most remarkable rate of H2 evolution of 508.44 μmol h–1.70Table 1 summarizes the different MoS2/metal oxide composites, the synthesis method, and photocatalytic parameters to produce H2 from the photocatalytic H2O splitting.

Table 1. Summarized List of Photocatalytic Hydrogen Evolution of MoS2 with Metal Oxides and Carbonaceous Materials.
photocatalyst synthesis method light source amount of catalyst sacrificial reagent H2 evolution activity ref
yolk shell TiO2/ultrathin MoS2 hydrothermal 300 W Xe lamp 50 mg 80 mL aqueous solution contains 20 mL methanol 2443 μmol g–1 h–1 (59)
TiO2/MoS2 ball-milling 300 W Xe lamp 0.2 g 100 mL of 15% mixed methanol-H2O solution 150.7 μmol h–1 (62)
N-ZnO/defect rich MoS2 hydrothermal natural sunlight 10 mg 50 mL aqueous solution contains 0.3 M Na2S and Na2SO4 17.3 mmol g–1 h–1 (63)
ZnO/MoS2 hydrothermal 300 W Xe lamp 100 mg 100 mL aqueous solution contains 0.1 M Na2S and 0.1 M Na2SO3 768 μmol g–1 h–1 (64)
Zn-porphyrin MoS2/ZnO hydrothermal 300 W LED lamp 100 mg 100 mL aqueous solution of 0.02 M TEOA 75 μmol g–1 h–1 (65)
Bi2O3/MoS2 hydrothermal 300 W Xe lamp 25 mg 50 mL aqueous solution contains 0.02 mol L–1 TEOA 3075.21 μmol g–1 h–1 (67)
CeO2/MoS2 hydrothermal 150 W Xe lamp 20 mg 20 mL aqueous solution of methanol 508.44 μmol h–1 (70)
activated carbon/MoS2 liquid phase reaction 30 W white light LED lamp 0.1 g 200 mL 10 vol % TEOA aqueous solution 872.3 μmol h–1 (72)
g-C3N4/MoS2 hydrothermal 300 W Xe lamp 20 mg 20 mL of H2O-methanol (4:1 v/v) solution 1497 μmol g–1 h–1 (73)
hollow g-C3N4/MoS2 impregnation and sulfidation 300 W Xe lamp 20 mg 100 mL aqueous solution contains 10 vol % lactic acid 26.8 μmol h–1 (75)
g-C3N4/MoS2 impregnation method 300 W Xe lamp 0.1 g 120 mL of aqueous solution contains 25% methanol 23.10 μmol h–1 (76)
g-C3N4/MoS2 hydrothermal 300 W Xe lamp 50 mg 120 mL of aqueous solution contains 25% methanol 867.6 μmol g–1 h–1 (77)
g-C3N4/MoS2 hydrothermal 300 W Xe lamp 50 mg 250 mL 0.1 M TEOA aqueous solution 1155 μmol g–1 h–1 (78)

2.3.2. MoS2/Carbonaceous Materials

To enhance the catalytic efficiency of MoS2 nanosheets, different layered carbon-based materials, including graphene, activated carbon, carbon nanofibers, and carbon nanotubes, have been incorporated. The resulting 2D/2D heterostructures with large contact areas will significantly benefit catalysis. Interfacial engineering between the layered heterojunctions as interface electron channels increases the HER performance.71 Recently, Yin et al. reported an activated carbon/MoS2 composite via a facile one-pot liquid phase reaction. The activity was mainly attributed to the in situ incorporation of activated carbon, resulting in an increased electrical conductivity due to the formation of thinner and smaller MoS2 nanosheets.72 A hydrothermal sulfurization and liquid phase exfoliation method was used to synthesize the modified gC3N4-MoS2 heterojunction for the first time. An excellent system for exposing active edge sites to photochemical processes was obtained by adjusting the lateral diameters of MoS2 nanosheets from 18 to 52 nm. The optimized structure with 20 wt % MoS2 loading and a lateral size of 39 nm had the maximum photocatalytic H2 activity of 1497 μmol h–1 g–1 and a quantum efficiency of 3.3% at 410 nm.

The density of MoS2 edge sites created at the interface of the MoS2/gC3N4 heterojunction is responsible for this increased HER activity. The MoS2 nanosheets with a large fraction of surface edge sites enhance the electron transfer.73 Inorganic molten salts offer a suitable platform for high 2D/2D heterostructures. A 2D/2D gC3N4-MoS2 visible-light-driven photocatalyst is developed via simple pyrolysis of melamine and (NH4)2MoS4 in a ternary molten LiCl-NaCl-KCl solution at 550 °C. The surface area of the composite was about 57 cm2 g–1, more than that of pure gC3N4. MoS2 assembly suppresses the overgrowth of gC3N4-nuclei and enhances the synthesis of a hybrid structure with a greater porosity and H2 evolution efficiency.74

A highly stable hollow carbon nitride nanosheet (HCNS) integrated MoS2 was recently constructed by Zheng et al. for an H2 evolution reaction. Applying lactic acid as a hole acceptor, the photocatalytic performance of MoS2/HCNS nanocomposites for H2 evolution was investigated. Upon 0.5 wt % MoS2 loading, HCNS indicated an increase in the H2 evolution rate of 26.8 μmol h–1. The improved photocatalytic efficiency was ascribed to developing surface heterostructures across MoS2 and HCNS, facilitating charge carrier separation and migration.75 Ge et al. used a simple impregnation technique to produce MoS2-gC3N4 composite photocatalysts. The maximum H2 evolution rate of 23.10 μmol h–1 was obtained for the 0.5 wt % MoS2-gC3N4 composite.76 Using a simple hydrothermal process, a unique 3D flowerlike hexagonal 2H-MoS2 was produced as thin stacked nanosheets, which were then combined with gC3N4 nanosheets of the same layered structure. The synthesized samples exhibited excellent photocatalytic and electrocatalytic activity with lower overpotential and substantial current densities. The SEM micrographs of pristine MoS2 demonstrated 3D flowerlike nanostructures having a diameter of 400–600 nm and thicknesses of 10 nm. Remarkably, the nano platelike subunits are linked with one another to develop 3D flowerlike networks. It overcomes disordered stacking of MoS2 layers by disclosing many active edge sites and increasing the specific surface area to decrease the diffusion pathway for both reactants and surface charges. The MoS2/gC3N4 composites have a thinner and fluffier microstructure than pure gC3N4, showing the introduction and deposition of MoS2 nanoflowers within the framework of gC3N4 nanosheets (Figure 4a,b). The 0.5% composition has the maximum photocatalytic activity of 867.6 μmol h–1 g–1.

Figure 4.

Figure 4

(a) SEM images of MoS2. (b) 0.5% of MoS2/gC3N4 photocatalyst. Reproduced with permission from ref (77). Copyright 2018 Elsevier.

The photocatalytic activity was elevated to 2.8 times higher than pure gC3N4. Introducing a unique flowerlike MoS2 structure boosted more increased and stable H2 evolution activity. Several H+ ions in the solution effectively bond to unsaturated active “S” atoms along the edges, converting them into H2.77 Yuan et al. recently reported a 2D-2D MoS2/gC3N4 photocatalyst having an H2 evolution rate of 1155 μmol g–1 h–1. The photoluminescence lifetime (PL) decreased as the concentration of MoS2 increased, confirming that MoS2 plays a crucial role in the effective separation of charge carriers.78 The chemical vapor deposition (CVD) process was used to synthesize 3D edge-oriented graphene with an edge-rich MoS2 nanoarray corresponding to gC3N4. The high conductivity and transparency of the 3D graphene layer provide a large exposed surface area for incorporating many photocatalysts, and developed 3D graphene/E-MoS2 structures exhibit superior optical and electrical characteristics. The maximum H2 evolution activity of 2232.7 μmol g–1 h–1 was achieved under white light irradiation.79 Binary composites of MoS2-carbon nitride (MoS2-CN) have recently been found as a substrate for piezo-photocatalytic hydrogen evolution reactions. The structure of the MoS2 on CN significantly impacts the piezoelectric activity of the binary composites. The MoS2 nanosheet acts as a piezo potential generator to induce an electric field on CN; thus, the recombination of the charge carriers is inhibited under low-frequency vortex vibration.80Table 1 outlines the various MoS2/carbonaceous composites and the synthesis methods and photocatalytic parameters used to produce H2 by photocatalytic water splitting.

2.3.3. MoS2/Metal Sulfides

Many metal chalcogenides, such as CdS, In2S3, Bi2S3, and CuS, have narrow band gaps and excellent absorption in the visible light range, making them suitable candidates for photocatalytic H2O splitting to release hydrogen.8183 However, metal sulfides also had significant drawbacks, such as fast electron–hole pair recombination and serious photocorrosion in photocatalysis. MoS2, as a catalyst in association with typical metal chalcogenides, can increase solar energy absorption in the visible range. For example, Li et al. demonstrated cost-effective and efficient photocatalysis based on MoS2 and CdS nanospheres.84 MoS2 is intimately formed on CdS and has an unusually high H2-evolution activity of 37.31 mmol g–1 h–1.

The agglomeration-resistant structure, low overpotential, and the large number of active sites of MoS2 contribute to the exceptional photocatalytic efficiency of these heterostructures.84 The HER activity of MoS2 nanosheets is linked to the active “S” atoms on the exposed edges rather than the S atoms presented on the basal planes, and this can be increased further by reducing the layer numbers of MoS2.85 In particular, CdS-MoS2 hybrids have attracted more attention, and several heterogeneous photocatalysts have been generated from CdS nanocrystals by integrating with MoS2 cocatalysts. For example, few-layered MoS2 nanosheets loaded with CdS nanorods produce a high H2 evolution rate (1009 mmol h–1 g–1) with an apparent quantum efficiency (AQE) of 71% at 420 nm. The few-layer MoS2 edges attached to CdS contain the more active sites, minimizing the distance and resistance of the transfer of electrons from the basal plane for the H2 evolution process.86

Ma et al. also developed a 2D-2D CdS/MoS2 heterojunction for photocatalytic H2 evolution reaction. Under visible light, the heterojunctions could attain hydrogen evolution activity of 1.75 mmol g–1 h–1, using an aqueous solution of sulfide and sulfate, and this became 2.03 times greater than pure CdS NSs. The increased activity can be due to the loading of MoS2 NSs, which effectively promotes the separation of charge carriers.87 A possibility of photocatalytic H2 evolution using biomass-derived sacrificial reagents such as glucose has been applied to a series of flowerlike CdS/MoS2 binary heterostructure composites. After 900 min of continuous testing, the 40% CdS/MoS2 catalyst achieved the maximum H2 evolution rate (55.0 mmol g–1 h–1) and maintained high stability. The synergistic interaction of CdS and MoS2 facilitates effective charge carrier separation. Furthermore, glucose performed as a better sacrificial agent, creating more protons during photocatalysis.88 Doping of MoS2 with heteroatoms is a simple and effective approach for producing an excellent catalyst for H2 production from water splitting. A Zn-doped few-layered MoS2 on the surface of CdS was reported. The doped MoS2 material was demonstrated to be a good cocatalyst for enhancing the photocatalytic activity of CdS. It also facilitates photoinduced charge transfer by providing many active edge sites for catalytic reactions and increasing long-term stability for useful applications. The optimized structure (6 wt % Zn-MoS2) shows the maximum catalytic activity of 241 mmol h–1 g–1, which is ∼75 times more than pure CdS.89

Similarly, Guo et al. synthesized the nickel subsulfide Ni3S2/MoS2 heterostructure. The Ni3S2 nanostructures were developed on Ni foam via a sulfuration approach by the CVD method, and MoS2 was formed along with the Ni3S2 nanostructures via a hydrothermal process. After synthesizing, the MoS2 achieved direct interaction with the Ni3S2 structure, resulting in effective interfacial interaction. The hybrid enhances broadband absorption covering 300 to 800 nm, yielding a significant increase in hydrogen evolution (540.75 μmol g–1 h–1).90 Zhao et al. synthesized CdS nanorods with MoS2 nanosheets, and the composite achieved an H2 production rate of 63.71 and 71.24 mmol g–1 h–1 in visible light and stimulated sunlight irradiation, respectively. The resulting heterojunction effectively separates photogenerated holes and electrons. Decorating CdS and MoS2 helped to reduce the photocorrosion of CdS nanorods by stabilizing the sulphions.91 Similar results were obtained in another CdS/MoS2 heterojunction.92

Multicomponent sulfides form heterostructures with MoS2 received considerable attention for applications in photocatalytic hydrogen production.93,94 By adding a transition metal into the CdS solid solution, photocorrosion of CdS can be eliminated and transition metal doped CdS can be formed. Among these, CdxZn1–xS takes more attention due to its excellent photocatalytic activity.

By combining with MoS2, a new CdxZn1–xS/MoS2 heterojunction was synthesized by an electrostatic self-assembly method.95 Similar results were also obtained for a 2D/2D ZnIn2S4/MoS2 composite, achieving a H2 evolution rate of 221.71 μmol h–1 was achieved under visible light irradiation. The close contact of ZnIn2S4 with MoS2 offers several channels for carrier migration. Furthermore, the active S atoms at exposed edges of MoS2 give more active sites for proton reduction.96 Whereas Chand et al. adopted a two-step process including hydrothermal and simple chemical methods to fabricate MoS2-ZnIn2S4 composites. The H2 production rate was also increased by adding MoS2 as a cocatalyst. MoS2 is used as an oxidation catalyst to strengthen the oxidation half-cell reaction and synchronously strengthen the reduction half-cell reaction; thus, the overall photocatalytic H2 evolution is enhanced.97

Many other reports have been published incorporating layered MoS2 with ternary chalcogenides. Hexagonal ZnIn2S4 has received more attention due to its structural similarity with MoS2. It is facile to develop an intimate 2D/2D ZnIn2S4/MoS2 heterojunction by growing a layer of MoS2 on the hexagonal ZnInS4 surface. The CB position of MoS2 was less negative than that of ZnIn2S4, resulting from the direct migration of photogenerated electrons possible through ZnIn2S4 to MoS2.

The optimized 5% MoS2/ZnIn2S4 composition exhibits a hydrogen production rate of 3891.6 μmol g–1 h–1.98 Huang et al. adopted a two-step hydrothermal method to construct the ZnIn2S4/MoS2 composite. The interaction of MoS2 can effectively promote electron–hole separation and extends the photocatalytic lifetime.99 For the first time, a solvothermal process was adopted for the controlled growth of ZnIn2S4 nanosheets on few-layered MoS2 slices to construct the MoS2/ZnIn2S4 hybrid. The exfoliated 2D platform framework of MoS2 slices can be an excellent supporting medium for the in situ growth of ZnIn2S4 nanosheets, enhancing interfacial charge transfer and minimizing self-agglomeration.100 Other reports also showed the highly enhanced photocatalytic performance for the H2-evolution of ZnIn2S4/MoS2 composites.101,102

The ZnIn2S4/MoS2 composite was synthesized using an environmentally friendly biomolecule-assisted microwave heating method. Using visible light irradiation, a ZnIn2S4/MoS2 composite having an optimized weight percentage of MoS2 was used to reduce harmful Cr (VI) ion to Cr (III) ion.

Furthermore, even following three cyclic tests, this catalyst demonstrated high stability. The ZnIn2S4 and MoS2 combine to produce a type II heterostructure responsible for the increased activity.103 Similarly, flowerlike p-MoS2/n-MgIn2S4 hybrids were synthesized and studied for photocatalytic efficiency in H2-evolution and N2-reduction. By inducing an internal electric current at the interface of two photocatalysts, the p-n heterojunction provides an excellent potential for separating photogenerated charge carriers. MoS2 nanosheets provide many exposed active sites through S–S bonding on the edges and offer efficient excitons separation.104 Swain et al. developed another type II heterostructure with excellent charge separation efficiency. A set of roselike p-MoS2 nanoflowers are decorated with n-type cubic CaIn2S4 (c-CIS) micro flowers incorporating different amounts of MoS2. Highly dispersed MoS2 nanoflowers with more active edge sites provided the platform for growing CaIn2S4, resulting in a hierarchical network (MoS2/c-CIS) with intimate contact. The p-n heterojunction character of the composite is revealed by electrochemical studies such as Mott–Schottky analysis and LSV curves. The lower slope of 0.5 wt % p-MoS2/n-CaIn2S4 compared to pure CaIn2S4 indicated the increased donor density capacity, which was the primary source of the photocatalytic performance (Figure 5a). When the potential was applied, pure c-CIS showed a very low photocurrent, i.e., 0.09 mA, although all composites showed a significant photocurrent. The plot is asymmetrical in forward and reversed biasing (Figure 5b). The modified photocatalysts comprising 0.5 wt % p-MoS2/n-CaIn2S4 had a higher H2 generation rate of 602.35 μmol h–1, having photocurrent densities of 0.743 mA cm–2. The type II heterojunction photocatalyst was developed to improve charge separation effectiveness by shifting electrons from the upper Fermi level of n-type c-CIS to the lower Fermi level p-type MoS2 (Figure 5c).105Table 2 outlines the various MoS2/metal sulfides composites and the synthesis methods and photocatalytic parameters used to produce H2 by photocatalytic water splitting.

Figure 5.

Figure 5

(a) Mott–Schottky plots of pure c-CIS and 0.5% MoS2/c-CIS. (b) LSV curves of pure c-CIS and different MoS2/c-CIS composites. (c) Possible charge transfer pathway for H2 evolution by p–n junction mechanism. Reproduced from ref (105). Copyright 2018 American Chemical Society.

Table 2. Summarized List of Photocatalytic Hydrogen Evolution of MoS2/Metal Sulfide Composites and MoS2/MOFs Composites.
photocatalyst synthesis method light source amount of catalyst sacrificial reagent H2 evolution activity ref
CdS/MoS2 hydrothermal 300 W Xe lamp 20 mg 80 mL aqueous solution contains 8 mL lactic acid 37.31 mmol g–1 h–1 (84)
CdS/MoS2 hydrothermal 300 W Xe lamp 20 mg 100 mL aqueous solution contains lactic acid 1009 mmol g–1 h–1 (86)
CdS/MoS2 hydrothermal 300 W Xe lamp 50 mg 80 mL aqueous solution contains 0.5 mol L–1 Na2S and 0.5 mol L–1 Na2SO3 1.75 mmol g–1 h–1 (87)
CdS/MoS2 hydrothermal 300 W Xe lamp 0.1 g 100 mL aqueous solution contains 0.1 M glucose 55 mmol g–1 h–1 (88)
Ni3S2/MoS2 CVD and hydrothermal 150 W Xe lamp 0.05 g 40 mL distilled H2O contains 0.1 mol L–1 Na2S and 0.1 mol L–1 Na2SO3 540.75 μmol g–1 h–1 (90)
ZnIn2S4/MoS2 electrostatic self-assembly visible light 36 mL deionized H2O + 4 mL lactic acid 4.97 mmol g–1 h–1 (93)
ZnIn2S4/MoS2 wet chemical method 300 W Xe lamp 25 mg 90 mL DIW H2O + 10 mL TEA 8898 μmol g–1 h–1 (94)
ZnIn2S4/MoS2 solvothermal 300 W Xe lamp 80 mg 100 mL aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 3891.6 μmol g–1 h–1 (98)
p-MoS2/n-ZnIn2S4 hydrothermal 150 W Xe lamp 20 mg 100 mL aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 320.2 μmol h–1 (102)
ZnIn2S4/MoS2 hydrothermal 150 W Xe lamp 0.1 g 100 mL aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 200.1 μmol g–1 h–1 (103)
MgIn2S4/MoS2 hydrothermal 150 W Xe lamp 20 mg 20 mL aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 570.8 μmol h–1 (104)
CaIn2S4/MoS2 hydrothermal 150 W Xe lamp 20 mg 20 mL aqueous solution contains 0.025 M Na2S and 0.025 M Na2SO3 602.35 μmol h–1 (105)
MoS2-MOF/Co3O4 hydrothermal 300 W Xe lamp 10 mg 15% (v/v) TEOA (30 mL) aqueous solution 629.75 μmol (106)
MoS2@Cd-MOF solvothermal 300 W Xe lamp 30 mg 150 mL aqueous solution contains 35 mM Na2S and 18 M Na2SO3 5587 μmol g–1 h–1 (108)
Cu0.9Co2.1S4@MoS2 hydrothermal 300 W Xe lamp aqueous solution contains TEOA 40156 μmol g–1 h–1 (110)

2.3.4. MoS2/Metal–Organic Frameworks (MOFs)

Metal–organic frameworks (MOFs), a new class of porous and crystalline materials exhibiting massive coordination centers, have found significant applications in photocatalysis.106,107 The photocatalytic activity of MoS2 in water splitting for H2 evolution can be enhanced by constructing heterogeneous interfaces with MOFs or MOF-derived metal oxides or sulfides. The sulfurization of the core–shell MoS2@Cd-MOF constructed a new CdS/MoS2 heterojunction. The CdS/MoS2 presents uniform loading of CdS nanoparticles on the MoS2 flowers. The composite exhibited a higher specific surface area of 78 m2 g–1, and the optimized CdS/MoS2 heterojunction showed an average H2 evolution rate of 5587 μmol g–1 h–1. In the heterojunction, the conduction band minimum of MoS2 is lower than that of CdS, so the photoexcited electrons transferring from CdS to MoS2 become possible. The MoS2 was a cocatalyst to trap photoexcited electrons and holes and improve photocatalytic activity.108

MoS2 combined with terephthalic acid to form MoS2-MOF and Co3O4 was deposited on these catalytic surfaces in methanol solution under light irradiation to form a MoS2-MOF/Co3O4 photocatalyst. MoS2-MOF constitutes a new type of catalyst that could provide a large specific surface area, pore size, and active sites to improve electron transport efficiency. The deposition of Co3O4 further accelerates the charge separation. The MoS2-MOF structure appears as a leaf, and Co3O4 is structured into this system as a venation. This catalytic system has an excellent fluorescence quenching rate and lifetime, and photocatalytic hydrogen evolution reached 629.75 μmol using an eosin Y-sensitized system; this is a successful example of preparing an organic framework composed of metal sulfides as an H2 evolution photocatalyst.106 A simple hydrothermal approach developed a unique 3D nanofibrous aerogel-based MoS2@Co3S4 composites. The heterojunction showed an exceptional H2-evolution activity of 228.2 μmol g–1 h–1 and delivered excellent photocatalytic dye degradation toward Cr (VI), sulfamethoxazole, and bacteria.109 A bimetallic metal–organic frameworks catalyst was coupled with MoS2 forming a heterojunction of the composition CuxCo3-xS4@MoS2. After optimization, the flowerlike Cu0.9Co2.1S4@MoS2 catalyst delivered 40156 μmol g–1 h–1 of visible light H2 production. These structures of Cu0.9Co2.1S4@MoS2 hybrid can fully expose the catalytic sites and maintain structural stability over a long period. The photocatalytic mechanism sensitized by eosin Y (EY) proceeds via a reductive electron transfer process. TEOA reductively quenches the excited EY molecules to produce EY*. Subsequently, electrons were transferred from the EY* molecule to the conduction band of Cu0.9Co2.1S4@MoS2, where the protons are reduced to form molecular H2 (Figure 6a). The stability of the photocatalyst was retained steadily after three consecutive cycles with an apparent quantum efficiency of about 3.2% (Figure 6b)110 (Table 2). A MOF-199 modified MoS2 heterojunction was reported by Qiao et al. to achieve high H2 evolution via water splitting (626.3 μmol g–1 h–1). The synthesized MOF/MoS2 composites show the combined properties of MoS2 and MOF; thus, the active sites can be significantly enhanced. Here the MOF-199 acts as a photosensitizer to assist MoS2 in harvesting visible light by transferring photoexcited electrons to MoS2.111

Figure 6.

Figure 6

(a) Photocatalytic process for H2 production on Cu0.9Co2.1S4@MoS2 photocatalyst using visible-light irradiation. (b) Stability tests of Cu0.9Co2.1S4@MoS2 in hydrogen evolution for three cycles under visible irradiation. Reproduced from ref (110). Copyright 2019 American Chemical Society.

2.3.5. Ternary Composites of MoS2

For photocatalytic H2 evolution processes, MoS2-based ternary materials have received considerable interest. Because of its numerous active sites and efficient separation of charge carriers, the MoS2 ternary systems produce more H2 than binary composites.112,113 Zhu et al. has synthesized a ternary CeO2@MoS2/gC3N4 catalyst with excellent photocatalytic performance for H2 production. The optimized 0D/2D composite produced a short and efficient multielectron transfer mechanism. The H2 evolution rate achieved 65.4 μmol/h, roughly 8.3 and 17.5-fold higher than pure gC3N4 and CeO2. The composite also exhibited an extended quantum efficiency of 10.35% at a wavelength of 420 nm.114 Transition metal doping (Co) of MoS2 effectively tunes the geometry and electronic structure for enhancing HER activity. A ternary CoMoS2/rGO/gC3N4 was constructed by a solvothermal method. The synergistic effect between the nanosheets of gC3N4, rGO, and Co-doped MoS2 enhances the HER activity rate to 684 μmol g–1 h–1. The enriched electrons on CoMoS2 with abundant active sites and lower atomic hydrogen absorption energy demonstrated high charge separation efficiency for reducing the protons to H2.115 In another study, a multifunctional ternary photocatalyst of MoS2 was introduced by Kumar et al. They developed a photocatalytic and electrocatalytic H2 production reaction for the ternary ZnO/MoS2/RGO heterojunction. As shown in Figure 7a, H2 evolution increases with increasing the MoS2-RGO content. The optimized composition achieved the highest H2 evolution (28.616 mmol g–1 h–1). Because it comprises a layered structure containing exposed S atoms at its edge sites with a high affinity toward H+ ions, the nanosized MoS2 facilitates H2 evolution (Figure 7b). During HER reaction, ZnO dissolves in an alkaline sulfide solution and the in situ-generated ZnS enhances interfacial charge transport to MoS2 and RGO. The photocatalytic process also indicates that the synergistic effect of RGO-MoS2 nanosheets and in situ formed ZnS efficiently reduce charge recombination in ZnO and significantly accelerates the H2 evolution reaction (Figure 7c).116 In another system, a noble metal-free heterostructured WS2-MoS2 integrated on CdS nanorods served as active sites for H2 production. By accelerating the photoinduced charge transfer and reducing electron–hole recombination, the WS2-MoS2 heterostructure ultrathin nanosheets became an efficient cocatalyst for enhancing the photocatalytic hydrogen evolution rate.117 CdS nanospheres with MoS2 and NiP ternary nanohybrids were synthesized for the first time by using a hydrothermal process and a MOF-templated approach. The composite has a high photocatalytic H2 evolution capacity of 72.76 mmol h–1 g–1 under optimum conditions.118

Figure 7.

Figure 7

(a) H2 volume evolved in various ternary ZnO/MoS2/RGO photocatalysts. (b) Layered structure of nanosized MoS2 showing exposed edges of sulfur atoms for H+ ions in aqueous solution for HER performance. (c) A plausible mechanism for enhanced electron transfer across the ZnO-MoS2-RGO heterojunction under solar irradiation using Na2S-Na2SO3 as a sacrificial reagent. Reproduced with permission from ref (116). Copyright 2017 Wiley.

Yu et al. described a biomolecule-assisted synthesis of CdS/MoS2/graphene hollow spheres. l-Cysteine worked as a sulfur source and was important for the self-assembly of the graphene structure to produce the ternary heterostructure. The CdS/MoS2 heterojunction developed by in situ MoS2 growth offers better contact. It allows more photogenerated electrons in CdS to move to the MoS2 layer and combine with H+ to produce H2 at considerably lower overpotentials.119 Similar results were obtained when MoS2 was modified with CdS and PANI.120 A ternary structure recently comprised of MoS2/SiO2/GO was reported as a photocatalytic degradation and H2 evolution catalyst. The as-synthesized lotuslike ternary photocatalyst has a suitable energy band structure for removing cotton pulp black liquor (CPBL) and simultaneously produces H2 (233.4 μmol). The loading of MoS2 onto this SiO2/GO structure dramatically improves the dispersion and specific surface area of MoS2, endowing its superior photocatalytic activity.121 As an excellent noble metal-free photocatalyst for H2 generation, ternary MoS2/ZnCdS/ZnS dual nanocomposites were fabricated. The material produces H2 at a 79.3 mmol g–1 h–1 rate and an apparent quantum yield of 47.9 at 420 nm. MoS2 has a higher specific perimeter and many active sites and easily separates and transports electron–hole pairs with enhanced photocatalytic performance.122 A chemical vapor deposition (CVD) method was adopted to construct a novel MoO2/MoS2/TiO2 ternary structure. Here few-layered MoS2 nanoflakes were deposited on TiO2 nanotubes, enabling the directional transfer of photogenerated charge carriers and showing nearly full spectrum absorption with excellent photocatalytic activity for H2 evolution.123 Feng et al. described a binary Mn0.2Cd0.8S/Co3O4 transformed to ternary Mn0.2Cd0.8S/MoS2/Co3O4. By producing a type-II heterojunction, the synergistic interaction of the Mn0.2Cd0.8S/Co3O4 p–n junction and MoS2 enhance photocatalytic H2 production. The few-layered MoS2 served as a cocatalyst with a low CB position, providing electron transport from the CB of Mn0.2Cd0.8S to the CB of MoS2 and Co3O4, respectively.124

Systematic incorporation of the different constituents in a ternary composite was essential for increasing solar absorption and H2 evolution reactions. Liu et al. constructed an Au-Cu nanoalloy/TiO2/MoS2 ternary composite by combining the Au-Cu nanoalloy and TiO2/MoS2 nanosheets. The high surface area of the TiO2/MoS2 hybrid improves sunlight absorption, and the Au-Cu nanoparticles serve as a bridge to connect the MoS2 and TiO2 via the surface plasmon resonance phenomenon.125 Chen et al. synthesized a CdS/MoS2/Mxenes ternary hybrid for the first time. When irradiated to visible light, the compound exhibits remarkable photocatalytic activity, with an H2 production rate of 9679 μmol g–1 h–1.126 Another MoS2/CQDs/ZnIn2S4 nanocomposites exhibit enhanced H2 evolution activity with an apparent quantum efficiency of 25.6%.127 The separation of photogenerated charge carriers and the resulting photocatalytic activity was considerably enhanced by the fabrication of 0D/2D heterostructures. Guan et al. developed a ZnIn2S4/MoS2-RGO nanosheets 0D/2D nanocomposite through a fast, low-temperature hydrothermal process. The MoS2/RGO nanosheets support inhibiting 0D ZnIn2S4 nanoparticles from aggregation in the ternary heterostructure, resulting in a significantly higher H2 generation than pure ZnIn2S4. The ZnIn2S4/MoS2-RGO composite benefited from the superior visible light absorption characteristics of MoS2 and RGO.128 The sheetlike MoS2/CdS/TiO2 ternary structure exhibited exceptional stability and a high H2 evolution rate of 4146 μmol g–1 h–1 under visible light. In this case, the conduction band of MoS2 enhances the transfer of photoexcited electrons from sunlight on CdS to porous TiO2 for photocatalytic H2 generation. Conducting MoS2 can increase the conductivity of the photocatalyst and reduce the photocorrosion of CdS as a hole collector.129 Another ternary composite with similar results includes CuInS2QDs/TiO2/MoS2 and gC3N4/red phosphorus/MoS2. The H2 evolution activity in these structures was 1034 μmol g–1 h–1 and 257.9 μmol g–1 h–1, respectively.130,131 The effective electrospinning and photodeposition process developed a unique ternary structure, MoS2/CdS-TiO2 nanofiber. The electro-spun TiO2 nanofibers have the advantages of improved CdS and MoS2 dispersion. The influence of MoS2 deposition on H2 production was investigated. When the concentration of MoS2 increases from 0.5 to 1%, the H2 evolution rate also increases. The photocatalytic performance for H2 evolution in the 1% MoS2/CdS-TiO2 sample is significantly higher (280.0 μmol h–1) than others (Figure 8a). The active hydrogen sites of MoS2 can minimize the overpotential and enhance photocatalytic H2 evolution. On the basis of the process (Figure 8b), photoexcited electrons can instantly be transported to the H2 evolution active sites of MoS2 and indirectly migrated to MoS2 through TiO2 as a bridge, enhancing the H2 evolution rate. The unique structure of the heterojunction accomplished this and close intrafacial contact.132 Ternary composites of CdS/MoS2/MWCNTs composite photocatalyst were reported by Jo et al. The few-layered MoS2 nanosheets exhibit high absorption and photoluminescence. A transient photocurrent density of 11.12 mA cm2 illustrates the charge separation ability of the hybrid, confirming the transport of electrons from photoexcited CdS to MoS2 edges along MWCNTs.133 Similar results were obtained for the bifunctional ternary heterostructure In2S3/MoS2/CdS composite.134 The defect-rich MoS2 in Cd1–xZnxS/MoS2/graphene hollow spheres performs as a low-cost cocatalyst to accelerate the photocatalytic H2 evolution reaction. Few-layered MoS2 having exposed active edges can take electrons directly from CdS or as an electron transfer medium via graphene. This heterostructure obtained an optimal H2 production activity of 2.97 mmol h–1 g–1 (Table 2).135 Furthermore, photocatalysts for HER over ZnIn2S4 were synthesized using MoS2 nanosheets on the surface of hollow carbon spheres (C/MoS2) and Co3O4 nanoparticles. The optimized C/MoS2@ZnIn2S4/Co3O4 composite shows a photocatalytic H2 evolution rate of 6.7 mmol g–1 h–1. The enhanced photocatalytic activity is due to a dual photo generated process in which C/MoS2 produces a Schottky heterojunction involving ZnIn2S4 and ZnIn2S4 which makes an S-scheme heterojunction with Co2O4. Additionally, the core–shell structure of the composite encourages intimate contact between constituents, enabling the separation of charge carriers.136 Similar observations were reported in another ZnIn2S4 ternary structure containing hollow carbon spheres (HCSs) with MoS2. The synergistic effect of MoS2 and HCSs effectively facilitates H2 production.137 Ran et al. fabricated a microporous “S” doped NH2-UiO-66 bridged ZnIn2S4/MoS2 sheet structure (MS/ZIS) photocatalyst. The synthesized MS/ZIS photocatalyst exhibits a relatively high H2 production rate of 5.69 mmol h–1 g–1. The “S” doped NU66 creates a new electron transport path, which benefits electron transport to the surface and active edge sites of MoS2.138

Figure 8.

Figure 8

(Left) Photocatalytic H2 evolution rate on MoS2/CdS-TiO2 nanofibers. (Right) Schematic representation of MoS2/CdS-TiO2 nanofibers for photocatalytic H2 evolution in the presence of visible light irradiation (λ ≥ 420 nm). Reproduced with permission from ref (132). Copyright 2017 Elsevier.

A two-step solvothermal approach followed by an in situ deposition technique was used to generate the oxygen-incorporated MoS2 (O-MoS2) coated on 1D Cd1–xZnxS with NiOx ternary nanostructures. NiOx nanoparticles were effectively grown on Cd1–xZnxS@O-MoS2 to construct unique Cd1–xZnxS@O-MoS2/NiOx composites. By exchanging the S-sites, oxygen is integrated into the MoS2 lattice. The longitudinal dimension of 1D CdZnS nanocrystals may serve as an effective channel. If the produced O-MoS2 contains many structural defects, the photoexcited electrons in CdZnS and O-MoS2 may also transfer to NiOx, extending the recombination procedure considerably (Figure 9).139 For photocatalytic H2 production, a ternary heterostructure of Ti3C2/MoS2/CdS was produced. In this report, MoS2 was formed in situ on Ti3C2 via electrostatic self-assembly, and hydrothermal techniques exhibited a vertically aligned framework suitable for more active edge sites for H2 production. This binary structure served as a cocatalyst for H2 evolution. It increases electron transport and prevents the recombination of electron–hole pairs formed from CdS by availing of the synergistic interactions of the MoS2/Ti3C2 structures.140

Figure 9.

Figure 9

Schematic showing the photocatalytic HER over CZ0.15S@O-MoS2/NiOx. Reproduced with permission from ref (139). Copyright 2019 Wiley.

Noble metals are generally recognized as excellent cocatalysts for photocatalytic H2O splitting reactions. A substantial surface plasmon resonance (SPR) phenomenon can contribute to a greater absorption cross-section with higher energy density of trapped electrons, resulting in rapid H2 production in photocatalytic splitting of water.141,142 Ternary composites of MoS2 containing noble metals have attracted much attention due to its high H2 evolution activity. An Au nanoparticle deposited ternary Au-MoS2/Ti3C2 photocatalyst was prepared using a hydrothermal method. The synthesized composite can modify electron transport paths. Due to the surface plasmon resonance effect of Au nanoparticles, the deposited Au nanoparticles on the surface of MoS2 were excited, thereby improving the photocatalytic hydrogen evolution rate.143 Similarly, the Pt-decorated ternary system, MoS2/Pt-TiO2, was efficiently developed by depositing MoS2 nanosheets to Pt-TiO2 frameworks. Appropriate ratios of MoS2 and Pt inhibited the recombination of photogenerated charge carriers. The rate of H2 evolution reaches 4.18 mmol h–1 g–1, which is around 41.8 times higher than pure TiO2. This unique method provides a path for transferring excited electrons from TiO2 particles using MoS2 and Pt nanoparticles as a link to facilitate proton reduction into H2.144 The sequential attachment of Au nanocrystals and CdS onto two-dimensional exfoliated MoS2 nanosheets (e-MoS2) was successfully fabricated by forming internally coupled plasmonic metal semiconductor hybrids. Due to their distinct band structure, the CdS and exfoliated MoS2 will have a wide optical absorption range. Moreover, its 2D morphology allows e-MoS2 to construct hybrids effectively with exposed surfaces.145 Similar results were observed for the compound AgInZnS/MoS2-GO, which exhibited the maximum hydrogen evolution activity of 1302 μmol h–1. The higher H2 evolution activity of the composite arises due to the synergetic effect of graphene and MoS2, where graphene facilitates the charge transfer, and MoS2 provides more active sites for H2 evolution.146 The Au nanoparticle-incorporated ternary CdS-Au/MoS2 core–shell hollow structure demonstrated the strongest visible-light photocatalytic H2 evolution. Electron and hole recombination was efficiently reduced after the formation of heterostructures by CdS, MoS2 nanosheets, and Au NPs, resulting in increased H2 generation. Figure 10 illustrates the reaction pathway for H2 evolution. The SPR effects of Au NPs stimulate photon absorption and accelerate electron transport, acting as an electron mediator from CdS to MoS2. During visible light excitation, the electrons produced by CdS are transported into the CB of MoS2 nanosheets via the Au NPs. MoS2 nanosheets are electronic sinks and provide active sites for the photocatalytic H2 evolution reaction.147

Figure 10.

Figure 10

Schematic illustration for the charge separation and enhanced H2 evolution activity in CdS-Au/MoS2 composites. Reproduced from ref (147). Copyright 2018 American Chemical Society

In another study, AuNPs@MoS2/ZnO nanorod hybrid structure has been successfully synthesized and shows excellent stability for H2 production. The MoS2 seemed like a monodispersed sphere with a diameter of ∼400 nm and composed of tangled MoS2 nanosheets. Introducing Au NPs into MoS2 microspheres results in a core–shell hybrid with improved plasmonic absorption. Adding ZnO nanorods into this system further enhances optical absorption. Mutual interactions between Au NPs (or ZnO nanorods) and MoS2 spheres efficiently prevent MoS2 nanosheets from peeling off from the spheres. Thus, after a 32 h test, the hybrid exhibited promising H2 production activity of 13.56 μmol g–1 min–1 with increased stability of 91.9%. This is one of the highest photocatalytic activities among the MoS2-based photocatalysts.148Table 3 summarizes the different MoS2/noble metal composites, the synthesis method, and the photocatalytic parameters for producing H2 from the photocatalytic water splitting.

Table 3. Summarized List of Photocatalytic Hydrogen Evolution of Ternary Composites of MoS2.
photocatalyst synthesis method light source amount of catalyst sacrificial reagent H2 evolution activity ref
CeO2/MoS2/g-C3N4 hydrothermal 3 W UV LEDs 50 mg 80 mL aqueous solution containing 0.5 M Na2SO3 and 0.5 M Na2S 65.4 μmol h–1 (114)
CoMoS2/rGO/C3N4 solvothermal 300 W Xe lamp 100 mg mixed solution of TEOA and H2O (volume ratio equal to 1/5) 684 μmol g–1 h–1 (115)
ZnO/MoS2/RGO hydrothermal natural sunlight 5.0 mg 45 mL aqueous solution contains 0.1 M Na2S and 0.1 M Na2SO3 28.61 mmol g–1 h–1 (116)
CdS/WS2/MoS2 hydrothermal 150 W Xe lamp 1.0 mg 20 vol % lactic acid 209.79 mmol g–1 h–1 (117)
CdS/Ni2P/MoS2 hydrothermal 150 W Xe lamp 1.0 mg 15 mL aqueous solution containing 3 mL lactic acid 72.76 mmol g–1 h–1 (118)
CdS/MoS2/graphene hydrothermal 300 W Xe lamp 30 mg 45 mL aqueous solution + 5 mL lactic acid 1913 μmol g–1 h–1 (119)
MoO2/MoS2/TiO2 CVD 300 W Xe lamp 100 mL aqueous solution contains 0.5 M ethylene glycol and 0.5 M Na2SO4 110 μmol h–1 cm–2 (123)
Mn0.2Cd0.8S/MoS2/CoO4 solvothermal 300 W Xe lamp 0.05 g 100 mL aqueous solution contains 0.5 mol L–1 Na2S and 0.5 mol L–1 Na2SO3 16.45 mmol g–1 h–1 (124)
Au-Cu nanoalloy-TiO2/MoS2 hydrothermal and annealing 300 W Xe lamp 5 mg aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 0.75 mmol g–1 h–1 (125)
MXene/CdS/MoS2 hydrothermal 300 W Xe lamp 5 mg 50 mL aqueous solution contains 0.25 M Na2S and 0.35 M Na2SO3 9679 μmol g–1 h–1 (126)
ZnIn2S4/CQDs/MoS2 hydrothermal 300 W Xe lamp 50 mg 100 mL aqueous solution contains 10 mL TEOA 133 μmol h–1 (127)
ZnIn2S4/MoS2/RGO hydrothermal 300 W Xe lamp 50 mg 100 mL 20% v/v lactic acid solution 425.1 μmol g–1 h–1 (128)
CuInS2 QDs/TiO2/MoS2 ultrasonic dispersion and annealing 300 W Xe lamp 50 mg 250 mL aqueous solution contains 0.1 M Na2S and 0.1 M Na2SO3 1034 μmol g–1 h–1 (130)
g-C3N4/red phosphorus/MoS2 in situ photodeposition 300 W Xe lamp 10 mg 9 mL aqueous solution contains 1 mL TEOA 257.9 μmol g–1 h–1 (131)
CdS/TiO2/MoS2 electrospinning andphotodeposition 300 W Xe lamp 20 mg 80 mL aqueous solution contains 10 vol % lactic acid 280 μmol h–1 (132)
CdS/MoS2/MWCNTs hydrothermal 300 W Xe lamp 20 mg 45 mL H2O + 5 mL lactic acid 676.0 μmol g–1 h–1 (133)
Cd0.8Zn0.2S/MoS2/graphene hydrothermal 300 W Xe lamp 50 mg 96 mL H2O + 4 mL lactic acid 2.97 mmol g–1 h–1 (135)
NH2-UiO-66 bridged ZnIn2S4/MoS2 hydrothermal 300 W Xe lamp 40 mg 80 mL aqueous solution contains 10% TEOA 5.69 mmol g–1 h–1 (138)
Cd1–xZnxS@OMoS2/NiOx hydrothermal 300 W Xe lamp 10 mg 100 mL aqueous solution contains 0.35 M Na2S and 0.25 M Na2SO3 66.08 mmol g–1 h–1 (139)
Au-MoS2/Ti3C2 hydrothermal 300 W Xe lamp 30 mg 50 mL methanol solution (deionized H2O: 70%, methanol: 30%) 12000 μmol g–1 h–1 (143)
Pt-TiO2/MoS2 hydrothermal 300 W Xe lamp 20 mg 100 mL solution contains 0.25 M Na2S and 0.5 M Na2SO3 4.18 mmol g–1 h–1 (144)
CdS-Au/MoS2 hydrothermal 300 W Xe lamp 10 mg 50 mL aqueous solution contains 10% lactic acid 237.90 μmol g–1 h–1 (147)
Au-ZnO/MoS2 hydrothermal 300 W Xe lamp 0.05 g 50 mL deionized H2O contains 0.3 M Na2S and 0.3 M Na2SO3 13.56 μmol g–1 min–1 (148)

3. Photocatalytic Carbon Dioxide Reduction Reaction (PCO2RR)

Carbon dioxide (CO2), the most abundant greenhouse gas, contributed significantly to global warming. Different techniques have been developed to decrease the amount of CO2 in the atmosphere, including absorption, bioconversion, and photocatalytic reduction.149151 Compared to photocatalytic H2 evolution, CO2 reduction by photocatalysis is a more complicated and challenging photochemical process. The products of photocatalytic CO2 reduction are selective and depend on several conditions. To address the significant issues of the preliminary modification processes, it must tackle both the challenges of thermodynamic and kinetic factors. Being a stable linear molecule, CO2 has a high bond energy of 750 kJ mol–1. The thermodynamic aspect of photocatalytic CO2 reduction requires an efficient light-responsive semiconductor that can capture enough external energy to break the C=O bond, essential for producing the C–H bond and synthesizing hydrocarbons.152,153 The interactions of intermediates with the surface of the catalyst determine product selectivity in CO2 reduction. Therefore, to attain maximum output during photocatalytic CO2 reduction, the photocatalyst should first satisfy the fundamental criteria: interact with the substrate molecules to start the transition process on the catalyst’s surface (CO2 or H2O). Another governing factor is photon absorption for exciting electrons from the catalyst’s valence band (VB) to the conduction band (CB). After excitation, the higher energy electrons in the CB showed higher potentials for CO2 reduction. As a result, catalysts with suitable band gaps (1.4–3 eV) are necessary to attain the maximum CO2 reduction.154,155 Since the first report by Inoue and co-workers on the photocatalytic reduction of CO2 to formaldehyde and methanol,156 various photocatalytic materials have been developed for photocatalytic CO2 reduction, including metal oxides,157,158 sulfides,159 perovskite-like materials,160 double-layered hydroxides,161 and layered materials.162,163 Among them, the transition metal dichalcogenides (TMD) find their application for CO2 transformation through electrocatalytic, thermochemical, and photocatalytic ways. Even though photocatalytic H2 evolution using TMD is well-known, their use in photocatalytic CO2 reduction has been recently reported. Mainly, photocatalytic CO2 reduction using MoS2 is an emerging area that is considered a future photocatalyst for CO2 reduction and a favorable, cost-effective alternative for noble metal catalysts, exhibiting remarkable CO2 electrochemical reduction activity with a high current density and low overpotential. The 2D layered nanostructures enable the modification and tunability of other TMDs. The atomic thickness allows for excellent optical transparency and mechanical flexibility, possibly allowing more surface-active sites over MoS2. The small diffusion path length and presence of terminated edges of MoS2 act as an active center for the CO2 reduction reaction, facilitating multiple electron transfers.164166

Interestingly, MoS2-based catalysts exhibit superior electrochemical selectivity over alcohols than other catalysts. In particular, alcohol synthesis from syngas has also been evaluated.167,168 Asadi et al. in 2014 reported that a layer-stacked bulk MoS2 containing molybdenum (Mo)-terminated edges showed superior CO2 reduction activity. The metallic nature of MoS2, the large d-electron density, and the availability of molybdenum-terminated edges are the primary reasons for the catalytic activity, which facilitates multiple electron transfer processes in the CO2 reduction pathway.169 Recently, researchers have focused on photocatalytic CO2 reduction using MoS2-based photocatalysts. Xie et al., for example, showed that the catalytic process over MoS2 proceeds via a carboxylate pathway, viz., CO2 → COOH* → CO* → CO to produce CO as the final product.

Due to the low energy barrier of CO, the method suggests more excellent selectivity toward CO than other products like HCOOH and CH3OH. Theoretical calculations indicate that the center of the d band at Mo exposed atoms at the edges is much closer than the metallic catalysts (e.g., Pt); thus, close interactions occurred with the adsorbed species. Meanwhile, the partial density of states (PDOS) of the activated CO2* and the exposed Mo atoms shows the strong interactions of Mo and O compared to the Mo–C interactions, which initiates the cleavage of the C=O bond to activate CO2*. The exposed edges are responsible for adsorbing the CO2 molecule and activating the two neighboring Mo atoms. The first transfer of e or h+ across the MoS2 edges is thought to be a hydrogenation step that controls the reduction pathway. It forms the unstable intermediate HO-CO*, which quickly decomposes to OH* and CO*. The hydrogenation of COOH* controls the product selectivity for CO and other hydrocarbons.170 Due to the low band gap potential, the charge recombination occurring over the surface of MoS2 is becoming a significant challenge. Forming products like CH3OH and CH4 is also difficult because this involves 6 e and 8 e processes, respectively, compared to products like CO, which require only two e that can be directly used as fuels and transformed into other powers through energy-intensive processes.171 Therefore, effective doping and heterojunction formations are required to get a more significant number of electrons available for reducing CO2, thereby increasing the efficiency of the MoS2-based photocatalysts for CO2 reduction. The remaining sections will focus on the roles of MoS2 and the various transformations that occurred as a transpire photocatalyst in CO2 reduction reactions.

3.1. MoS2 Nanostructures for Photocatalytic CO2 Reduction

The MoS2 nanoflower-like morphology with dangling bonds at the stacked sheet edge makes it ideal for photocatalytic CO2 reduction. Meier et al. reported a controllable edge-rich nanoflower morphology with abundant edge sites through a chemical vapor deposition (CVD) method. Photocatalytic activity increases with increasing edge plane density of nanoflower due to unsaturated and dangling bonds. During CVD synthesis, the energy gap (Eg) was separately varied from 1.38 to 1.83 eV. Carbon monoxide (CO) was the most prominent product of CO2 photoreduction in the presence of H2O. The production rates of CO nearly doubled after a post-treatment reduction step. This reliable CVD technique contributes to developing efficient and stable 2D-MoS2 nanoflower materials for CO2 reduction.172 MoS2 nanosheets combined with metal oxides were identified as a cost-effective substitute for noble metal cocatalyst for photocatalytic reduction of CO2. Tu et al. reported the formation of a 2D-MoS2 nanosheet on TiO2 nanosheets. To reduce CO2 by UV–vis light irradiation, pure TiO2 nanosheets and MoS2/TiO2 hybrid nanosheets were placed in a CO2-saturated 1 M NaHCO3 aqueous solution. The CH3OH production of 0.5 wt % MoS2/TiO2 reached to 10.6 μmol g–1 h–1, which was 2.9 times that of pure TiO2 (3.7 μmol g–1 h–1). The fast electron transport from TiO2 to MoS2 is indicated by the PL decay lifetime of 0.51 ns. The weak van der Waals interactions in the S–Mo–S layers result in Mo-terminated edges containing high d-electron density, which contributes to excellent and selective photocatalytic CO2 reduction.173 Similarly, few-layered MoS2 with uniformly formed mesoporous TiO2 on the surface of 3D graphene showed higher photocatalytic CO2 reduction. The MoS2 layers in the 3D aerogel perform as a cocatalyst to minimize the backward charge transfer through graphene to TiO2. The few-layered MoS2 with unsaturated S atoms on exposed edges improved its activity and generated a high electron concentration in the composite between the MoS2 and graphene layers. The yield of CO formation was up to 92.33 μmol g–1 h–1.174

Jia et al. recently synthesized a MoS2/TiO2 heterojunction photocatalyst for CO2 reduction. The ultrathin MoS2 nanosheets were equally combined with the TiO2 nanoparticles. The conduction band (CB) edge potential of MoS2 (−0.93 V) was more negative than that of TiO2 (−0.55 V), so the photogenerated electrons on the surface of MoS2 sheets could easily migrate to TiO2. CO and CH4 were the major products in the experimental process, and the 10% MoS2/TiO2 composite gave the maximum yield. It is about 268.97 μmol g–1 and 49.93 μmol g–1, respectively.175 Xu et al. constructed a 1D/2D TiO2/MoS2 nanostructure composite. Here, the MoS2 sheet arrays with lateral dimensions of 80 nm with thicknesses of up to 2 nm were firmly connected to the surface of TiO2 nanofiber. The CO2 adsorption and surface area of the produced samples increase as the MoS2 loading increases. With increasing CO2 loading, the CH4 and CH3OH synthesis rates increase, reaching a maximum of 2.86 and 2.55 μmol g–1 h–1, respectively, with a quantum yield of 0.16% for the best TiO2/MoS2 sample (TM10). The DFT calculations further demonstrate that MoS2 has a higher work function than TiO2, resulting in electron transfer from MoS2 to TiO2 upon photoexcitation.176

A type II p-n heterojunction charge transfer model was formed by coupling the α-Fe2O3 and MoS2 composite. The loading of α-Fe2O3 promoted 2H to 1T MoS2 phase conversion and increased charge carrier migration. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) analysis revealed a deoxygenation pathway for CO2 conversion. Several intermediate species are formed upon the adsorption of CO2, which ultimately results from excellent CO2 reduction performance to CH4 and CH3OH.177 In another system, a few-layered MoS2-loaded flowerlike Bi2WO6 nanocomposite was reported by Dai et al.178 The as-prepared composites showed an excellent photoreduction of CO2 into hydrocarbons compared to pure Bi2WO6. The intimate interfacial contact between MoS2 and Bi2WO6 significantly influenced the transfer of photogenerated charge carriers. The composite also showed much stronger absorption within the visible region than pure Bi2WO6. The photocurrent response of the composites (Figure 11a) provided information about the higher separation efficiency of photogenerated charge carriers. The methanol and ethanol obtained from the optimized sample (0.4 wt % MoS2) after 4 h of visible light irradiation were 36.7 and 36.6 μmol g–1, respectively, 1.94 times higher than pure Bi2WO6.

Figure 11.

Figure 11

(a) Transient photocurrent response of Bi2WO6 and MoS2/Bi2WO6 nanocomposites under visible light irradiation. (b) Proposed CO2 photoreduction and charge transfer mechanism in the MoS2/Bi2WO6 nanocomposite. Reproduced with permission from ref (178). Copyright 2017 Elsevier.

A possible photocatalytic mechanism (Figure 11b) was also proposed. It indicated that the CO32–, HCO3, and H2CO3 formed in the aqueous solutions might serve as reactive carbon sources during the photoreduction process. The 2D-π conjugated structure could serve as an electron-accepting material and contribute to separating e-h+ pairs for CO2 reduction.178 A flowerlike MoS2/Ag/SnO2 ternary composite that serves as a bifunctional catalyst for photocatalytic CO2 reduction and degradation of pollutants was recently produced using a simple hydrothermal process. Because of the high specific surface area, several folded edges, and ultrathin nature, Ag nanoparticles and SnO2 NSs were distributed uniformly on the MoS2 surface, enhancing photoexcited charge transport. The modified MoS2/Ag/SnO2 nanocomposites exhibit outstanding visible-light photocatalytic activity for converting CO2 to methane (CH4). The quantum efficiency calculated at λ = 420 nm was found to be 2.38% which substantiates the high CO2 conversion efficiency of the catalyst. SnO2 accepts the photogenerated electron of MoS2 through Ag in the heterojunction, which directly reduces CO2 and produces CH4 as the final product. Besides CO2 reduction, the composite also shows high activities for pollutant degradation.179 Another binary composite using 2D MoS2 nanosheets with a metal sulfide has been reported. The p-n heterostructure Bi2S3-MoS2 nanocomposite revealed improved sunlight absorbance and photocatalytic reduction of CO2. The high catalytic performance was due to the increased S-defects on the MoS2 surface, hence increasing interfacial interaction between MoS2 and Bi2S3.180 As mentioned in the first section, graphitic carbon nitride (gC3N4) can be an efficient photocatalyst to construct the heterojunction with layered MoS2. Studies have been reported regarding the applications of MoS2/g-C3N4 nanocomposites for photocatalytic reactions to convert CO2 to hydrocarbons. Recently, Qin et al. mentioned the synthesis of MoS2/gC3N4 nanocomposites to convert CO2 into fuels by a simple hydrothermal followed by a calcination process. The synthesized 10% MoS2/g-C3N4 heterojunction exhibited the highest conversion efficiency in reducing CO2 to CO up to 58.59 μmol g–1 and showed significant stability after a seven-hour reaction. Z-Scheme charge transfer was constructed by correctly coupling the two semiconductor photocatalysts for CO2 reduction.181

Combining semiconductors with noble metal NPs (Ag, Au, or Pt) caused localized surface plasmon resonance (LSPR) to show great potential in photocatalytic CO2 reactions. Interestingly, studies have confirmed that the generation of LSPR can optimize the structure of MoS2. Sun et al. demonstrated the Au nanoparticle decorated ultrathin MoS2 nanosheet acts as an excellent heterogeneous catalyst for photocatalytic CO2 reduction. The DFT calculations also endorse the experimental data. DRIFTS spectrum proposed a reaction pathway for the formation of CH4, CO2 → COO– → −COOH → −HCO→ −HCOH → −CH2OH→ −CH3 → CH4 via the hydrogenation process. For the Au-MoS2 sample, the reaction barrier for forming a COOH intermediate decreases from 2.17 to 1.82 eV, suggesting faster CO2 reduction.182 Similarly, a DFT study on Ag-loaded 2H-MoS2 was stimulated by Ouyang et al. for the photocatalytic reduction of CO2. From the series of Ag loading on the 2H-MoS2 nanosheet, the 20 wt % Ag/MoS2 is most suitable for converting CO2 to CH4. The Gibbs free energy was reduced from 2.830 to 0.925 eV for the 20 wt % Ag/2H-MoS2 nanocomposite.183

For adjusting the band gap of MoS2, phase conversion via chemical exfoliation was used. In a recent study, Zheng et al. reported the synthesis of Ag/2H-MoS2 composites and their photocatalytic activity in reducing CO2 to methanol. A wet chemical process was invented to exfoliate MoS2 nanosheets and transform them from the 1T phase to the 2H phase. This composite obtained a larger specific surface area through Li-ion intercalation and exfoliation than the hydrothermally synthesized MoS2. The Ag nanoparticles can act as a pool for collecting the photogenerated electrons of MoS2 nanosheets and suppress the recombination of charge carriers. The maximum yield of methanol obtained after 10 h irradiation at 20 wt % Ag/2H-MoS2 was about 365.08 μmol g–1 h–1.184 A study by Lu et al. demonstrated the use of a noble metal-free catalyst (Mn0.2Cd0.8S) wrapped in a series of MoS2 ultrathin nanosheets for photocatalytic energy conversion. The optimal loading of MoS2 content was 3%, giving an H2 production rate of 335.02 μmol h–1 and a CH3OH production rate of 2.13 μmol h–1. The production of methanol increases with increasing the MoS2 content. The improved photocatalytic efficiency benefits from the intimate interfacial contact between Mn0.2Cd0.8S and MoS2, which effectively promotes the separation and transportation of photogenerated charge carriers.185

Recently, a composite comprising MoS2 and a halide perovskite material, CsPbBr3, formed a stable heterojunction. The heterojunction contacts between the two components increased the lifetime confirmed by the PL, surface, and transient photovoltage (SPV, TPV) and DFT studies. The temperature-programmed CO2 desorption (TPD) analyses showed that introducing MoS2 increased the ability of the CO2 uptake of the composite.186 CO2 photoreduction and simultaneous H2 production were achieved by a ternary nanocomposite of MoS2 and N-containing polymer polypyrrole (PPy) on reduced graphene oxide (rGO). The resulting rGO-MoS2/PPy nanocomposites reveal that rGO-MoS2 acts as a rigid template, ensuring consistent growth and heterogeneous nucleation of PPY, thereby increasing their photocatalytic reduction of CO2 into CH4, CO, and H2. Composites containing optimal amounts of PPy (rGO-MoS2/PPy-150) exhibited the highest photocatalytic activity to CO (3.95 μmol g–1 h–1), CH4 (1.50 μmol g–1 h–1), and H2 (4.19 μmol g–1 h–1), with a quantum yield of 0.30%.187 Similarly, a simple solvothermal method constructed a novel S–C–S heterojunction by Yin et al. The MoS2 was assembled into 3D-uniform nanoflower balls after a simultaneous coupling connection with rGO and SnS2. The valence band (VB) positions of pure MoS2 (1.61 eV) and SnS2 (1.50 eV) indicate that they can form an effective Z-type heterojunction so that the electrons can quickly be excited from the MoS2 CB to VB in SnS2 and thereby to rGO to reduce CO2 to CO and CH4.

The multilevel electron transport produces CO and CH4 up to 68.53 and 50.55 μmol g–1 h–1, respectively.188 A binary photocatalyst composed of the SiC@MoS2 nanoflower was reported by Wang et al. for simultaneous photocatalytic CO2 reduction and H2O oxidation under visible light. All the SiC@MoS2 samples show much higher CH4 evolution than the individual components. The optimized samples show the CH4 evolution up to 323 μL g–1 h–1 and O2 evolution of about 620 μL g–1 h–1 with high photocatalyst stability. The photocatalytic reduction of CO2 follows the hydrogenation pathway (Figure 12). One CO2 molecule is converted into one CH4 by four sequential hydrogenation half-cell reactions. This pathway takes advantage of the photoreduction potential of SiC and the powerful photoreduction ability of MoS2 nanosheets.189

Figure 12.

Figure 12

Reaction pathway for photocatalytic CO2 reduction by H2O on SiC@MoS2. Reproduced from ref (189). Copyright 2018 American Chemical Society.

Layered double hydroxides (LDH) are a new materials class that has recently attracted much interest. A study by Zhao et al. demonstrated that the heterojunction of MoS2 with Co-Al-LDH through an electrostatic interaction facilitates the charge transfer and syngas production by controlling the heterojunction concentration. Photogenerated electrons spontaneously transfer from LDH CB to MoS2, and the syngas ratio (H2:CO) was precisely turned from 1.3:1 to 15:1 by altering only the catalyst concentration in the photocatalytic CO2 reduction under visible light.190

A multidimensional ternary heterojunction of Bi2S3/TiO2/MoS2 was reported by Alkanad et al. for the photocatalytic CO2 reduction reaction where Bi2S3 and TiO2 were embedded and wrapped within the MoS2 nanosheet. The in situ irradiated photoelectron spectroscopy (ISI-XPS) reveals the switchable feature of the heterostructure. Electron flow follows the S-scheme approach in UV–vis-NIR (CH3OH and C2H5OH as the major products), whereas, in vis-NIR, it forms a type II heterojunction (CH4 and CO), thus exhibiting a selective photocatalytic CO2 reduction reaction. Moreover, the quantum efficiency of the system reaches 4.23% at 600 nm wavelength.191 Another higher CO2 reduction capacity was observed in a ZnO/MoS2 nanosheet composite where the hollow structure of MoS2 encapsulates the ZnO nanoparticles and enables multiple refractions of photons in the cell. MoS2 as a 2D sheet would enhance electron mobility, so the photogenerated electrons and holes are accumulated on the surface of MoS2. Therefore, CB and VB of MoS2 simultaneously reduce CO2 and oxidize H2O.192 Metallic or nonmetallic doping creates defects in the MoS2 lattice and facilitates charge transfer at the interface to facilitate CO2 reduction. Such a ternary system in which In2S3 was decorated with MoO3@MoS2 was fabricated for visible light photoreduction of CO2 to CH4 and CO. The presence of In2S4 induces S-rich sites and O vacancies in the composite for CO2 reduction, and a maximum yield of CH4 (29.4 μmol g–1 h–1) and CO (35.6 μmol g–1 h–1) was obtained.193 Similarly, duel In2S4-NiS decorated MoO3@MoS2 was reported by Zheng et al. through an in situ sulfurized approach of nonthermal plasma. Increasing the In2S4-NiS content into the MoO3@MoS2 system increased the CH4 yield of 49.11 μmol g–1 h–1.194 Metal–organic frameworks (MOFs) were an excellent platform for developing heterogeneous catalysts for CO2 reduction. By integrating MoS2 nanosheets into hierarchically porous defective UiO-66 (d-UiO-66), a composite of d-UiO-66/MoS2 was prepared for the selective photocatalytic reduction of CO2 to CH3COOH. The unsaturated Mo atoms on the MoS2 nanosheet would contact d-UiO-66 to form Mo–O–Zr bimetallic sites for selective CO2 reduction. By loading the composite on a nylon 66 microfiltration membrane, CH3COOH was formed with a rate of 39.0 μmol g–1 h–1 without any sacrificial reagent.195 Similarly, a new In2O3/MoS2 comodified ZnO nanorods also showed CO2 photoreduction to CO and CH4. Surface modification of the In2O3 and MoS2 species would create transfer channels for holes and electrons on the photocatalyst’s surface and improve the photocatalytic activity.196

3.2. MoS2-Based Photoelectrocatalytic CO2 Reduction

Cu2O is a promising low-cost photocatalyst for CO2 reduction, which showed higher selectivity for methanol formation. But their poor conductivity and charge recombination render their application. To overcome these drawbacks of Cu2O, heterojunctions with suitable components have been formed. Among these structures, a new MoS2-Cu2O heterostructure was fabricated by Singh et al. Here two types of MoS2, p-MoS2 and n-MoS2, were used to form heterostructures with CuO: Cu2O-c (cubic morphology) and Cu2O-co (cuboctahedron morphology). This new system exhibited both CO2 reduction and water oxidation capacity. Among these structures, the p-MoS2/Cu2O-c type system showed higher activity, following a Z-scheme charge transfer to yield CH3OH, and DFT calculations further confirm it.197 Recently, MoS2-rods/TiO2 nanotubes (NTs) heterojunction was introduced for exploiting photoenhanced electrocatalytic reduction (PEEC) of CO2 to methanol.

The resulting composite exhibited excellent optical performance but had an unmatched conduction band minimum (CBM) of −0.15 eV with low Faradaic efficiency. Therefore, developed a new PEEC to reduce CO2 to methanol on MoS2-rods/TiO2 NTs. Adding illumination to the electrocatalytic system improved the Faradaic efficiency to 111.58% more than the same electrocatalytic (EC) system. They introduced a unique mechanism for the PEEC reduction of CO2 to methanol. The reduction of CO2 to CH3OH consists of six electronic reactions, as shown in Figure 13a. On one side, when applying −1.3 V potential on the cathode, the anode potential is 1.82 V versus SCE and can oxidize H2O to O2 and protons. Then the protons will reach the cathode for CO2 EC reduction. Moreover, the catalyst has a valence band (VB) suitable for in situ PC oxidation of H2O to generate protons. So, it provides more protons for EC reduction of CO2 to methanol with high efficiency (Figure 13b).198Table 4 summarizes the recent progress in photocatalytic CO2 reduction using MoS2 nanocomposites.

Figure 13.

Figure 13

(a) Probable speculation reaction pathway for the reduction of CO2 to CH3OH. (b) Mechanism of the PEEC reducing CO2 process at MoS2-rods/TiO2NTs electrode. Reproduced with permission from ref (198). Copyright 2014 Elsevier.

Table 4. Summarized List of Phototocatalytic Carbon Dioxide Reduction Using MoS2 Compounds.

photocatalyst synthesis method light source amount of catalyst major products ref
MoS2 nanoflower CVD solar simulator 150 mg CO (172)
MoS2/TiO2 hydrothermal 300 W Xe lamp 0.1 g CH4 (10.6 μmol g–1 h–1) (173)
TiO2/MoS2/graphene hydrothermal 300 W Xe lamp CO (92.33 μmol g–1 h–1) (174)
MoS2/TiO2 hydrothermal 300 W Xe lamp 50 mg CO (268.97 μmol g–1) (175)
        CH4 (49.93 μmol g–1)  
TiO2/MoS2 hydrothermal 350 W Xe lamp 50 mg CH4 (2.86 μmol g–1 h–1) (176)
        CH3OH (2.55 μmol g–1 h–1)  
α-Fe2O3/MoS2 hydrothermal 300 W Xe lamp 50 mg CH4 (121 μmol g–1 h–1) (177)
        CH3OH (41 μmol g–1 h–1)  
MoS2/Bi2WO6 hydrothermal 300 W Xe lamp 50 mg CH3OH (36.7 μmol g–1) (178)
        C2H5OH (36.6 μmol g–1)
SnO2/Ag/MoS2 hydrothermal 300 W Xe lamp CH4 (20 μmol) (179)
        CO (9 μmol)  
Bi2S3/MoS2 hydrothermal 150 mW cm–2 Xe lamp 0.2 g CO (40 μmol g–1) (180)
        CH4 (2.5 μmol g–1)  
MoS2/g-C3N4 hydrothermal 300 W Xe lamp 50 mg CO (58.59 μmol g–1) (181)
Au/MoS2 hydrothermal and ultrasonic dispersion 300 W Xe lamp 0.025 g CH4 (19.38 μmol g–1 h–1) (182)
Ag/MoS2 hydrothermal, calcination and reduction 200 W Hg lamp 20 mg CH3OH (365.08 μmol g–1 h–1) (184)
MoS2/Mn0.2Cd0.8S hydrothermal 300 W Xe lamp 50 mg CH3OH (2.13 μmol h–1) (185)
polypyrrole/rGO/MoS2 hydrothermal and ultrasonication 300 W Xe lamp 50 mg CO (3.95 μmol g–1 h–1) (187)
        CH4 (1.50 μmol g–1 h–1)  
MoS2/SnS2/rGO solvothermal 8 W Hg lamp 20 mg CO (68.53 μmol g–1 h–1) (188)
        CH4 (50.55 μmol g–1 h–1)  
SiC/MoS2 self-assembly 300 W Xe lamp 10 mg CH4 (323 μL g–1 h–1) (189)
Bi2S3/TiO2/MoS2 microwave synthesis 300 W Xe lamp 50 mg CH3OH (119 μmol g–1) (191)
        C2H5OH (102.7 μmol g–1)  
MoS2@ZnO calcination 150 W Hg lamp 0.1 g CH3OH (170.5 μmol g–1) (192)
In2S3/MoO3@MoS2 in situ sulfurization 200 mW cm–2 Xe lamp 10 mg CH4 (29.4 μmol g–1 h–1) (193)
        CO (35.6 μmol g–1 h–1)  
In2S3/NiS/MoO3@MoS2 in situ sulfurization 300 W Xe lamp 10 mg CH4 (49.11 μmol g–1 h–1) (194)
        CO (6.19 μmol g–1 h–1)  
ZnO/In2O3/MoS2 hydrothermal and ultrasonication 300 W Hg lamp 100 mg CO (5.628 μmol g–1 h–1) (196)
        CH4 (3.048 μmol g–1 h–1)  

4. Conclusions and Perspectives

This review focused on 2D layered MoS2-based nanostructures for photocatalytic H2 evolution and CO2 reduction to hydrocarbon fuels, which have been discussed in the literature over the past several years.

The layered structure of MoS2 with more active edge sites mainly contributed to its photocatalytic activity. The first part of this review illustrates the H2 evolution reactions of binary and ternary composites of MoS2. Binary composites cover the composites of MoS2 with metal oxides, carbonaceous materials, metal sulfides, and metal–organic frameworks, and ternary heterosystems comprise the combinations of different binary structures with MoS2. Many studies have shown that combining MoS2 as a cocatalyst can overcome the barriers to H2 evolution in many systems. The recombination of charge carriers in several metal oxides is reduced by forming binary structures with MoS2, where the tuning of the band gap of metal oxide was achieved.60 The 2D/2D binary heterostructures formed with MoS2 and carbonaceous materials provide a high surface area that significantly benefits higher H2 evolution activity. Especially the H2 production was increased to a large extent by introducing thinner and fluffier flowerlike MoS2 structures in the framework of gC3N4 nanosheets.76,77 The significant drawbacks like photocorrosion and fast electron–hole recombination of metal sulfides can be effectively eliminated by forming binary structures with MoS2. Binary heterostructures formed with multimetal sulfides like ZnIn2S4, CuInS2, and MgIn2S4 with MoS2 are also widely used as a photocatalyst for H2 production. The flowerlike morphology of MoS2 provides an excellent platform for catalytic activity.104,105 The peculiar porous structure of MOFs exhibits many coordination centers for H2 evolution activity, and the association of MoS2 further increases its activity. Apart from these binary combinations, MoS2-based ternary composites have much interest due to their efficient multielectron transport mechanism for H2 production. In these systems, MoS2 acts as a channel to accept more photogenerated electrons for combining H+ to produce H2 at lower overpotentials.131,132 Ternary composites of MoS2 containing noble metals always show a high H2 evolution rate with excellent stability. Such plasmonic metal hybrid systems with MoS2 contributed much better photocatalytic activity than others. For example, an Au-deposited MoS2/ZnO system exhibited the highest H2 evolution rate with good stability.148

Though encouraging progress has been achieved toward photocatalytic H2 evolution using MoS2-based nanocomposites, some problems and challenges still exist. One of the crucial difficulties is that MoS2 alone has negligible photocatalytic activity, and its bulk form is an indirect-band gap semiconductor. The band structure of MoS2 varies with the number of layers. When it became a monolayer, it gave a high optical absorption coefficient and showed better HER performance. Therefore, more research is needed to explore single-layer MoS2 as a hydrogen evolution catalyst. The combinations of MoS2 with other transition metal dichalcogenides (TMDs) such as WS2, MoSe2, and WSe2 would give much better activity. For example, the WS2-MoS2 binary structure with CdS was an efficient noble-metal-free catalyst with high durability. So, the development of such type of hybrid metal sulfide combinations would be more acceptable.199

Furthermore, a few reports are available using MoS2 with inorganic compounds like Mxenes, giving the highest H2 evolution rate.126 Similar results were obtained for the 3D structure of graphene containing MoS2. In this context, efforts for efficient photocatalysts are also to be developed with different compositions for enhanced H2 evolution performance.79 In most cases, H2 evolution occurs in aqueous solutions using diverse sacrificial reagents like triethanolamine, lactic acid, Na2S-Na2SO3 pair, and methanol as electron donors. The use of such sacrificial reagents in MoS2-based nanocomposites for H2 evolution implied a higher cost for future applications on a large scale. So, it is essential to develop MoS2-based photocatalysts for H2 production without using additional sacrificial reagents.

Photocatalytic CO2 reduction to hydrocarbon fuels mainly involves multielectron transfer mechanisms and includes a variety of redox potentials. As a result, several products can be generated during the photocatalytic CO2 reduction process, and the product selectivity is related to the chemical pathways of the reaction. Thus, the role of MoS2 in photocatalytic CO2 reduction must be better understood to improve selectivity for a particular fuel. Layered MoS2 has been widely employed to improve photocatalytic H2-production performance. However, research on MoS2-based photocatalysts for CO2 reduction is still in the early stages. Efforts in developing layered MoS2-based photocatalysts capable of simultaneous H2 evolution and CO2 conversion are an area of current interest.185,187 Besides the photochemical conversion of CO2, MoS2-based systems for CO2 reduction using photoelectrochemical splitting are highly required due to their high Faradaic efficiency. Theoretical studies stimulate the electronic properties of MoS2-nanocomposites, which will help one to understand the efficiency of the catalyst in photoelectrochemical studies.197,198 In short, developing 2D layered MoS2-based nanocomposites is a substantial breakthrough in photocatalytic H2 evolution and CO2 reduction.

Acknowledgments

We are grateful for the financial support of the University Grants Commission (UGC), India. We also thank Gladiya Mani, Sreerenjini C. R., Sithara Gopinath, and Aswathy V. Kumar for fruitful discussions (Mahatma Gandhi University, Kottayam, Kerala).

Author Present Address

§ Assistant Professor, Department of Chemistry, Morning Star Home Science College, Angamali South, Kerala 683573, India

The authors declare no competing financial interest.

References

  1. Chen X.; Shen S.; Guo L.; Mao S. S. Semiconductor Based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. 10.1021/cr1001645. [DOI] [PubMed] [Google Scholar]
  2. Fan W.; Zhang Q.; Wang Y. Semiconductor-Based Nanocomposites for Photocatalytic H2 Production and CO2 Conversion. Phys. Chem. Chem. Phys. 2013, 15, 2632–2649. 10.1039/c2cp43524a. [DOI] [PubMed] [Google Scholar]
  3. Kamat P. V.; Bisquert J. Solar Fuels. Photocatalytic Hydrogen Generation. J. Phys. Chem. C 2013, 117, 14873–14875. 10.1021/jp406523w. [DOI] [Google Scholar]
  4. Xiang Q.; Cheng B.; Yu J. Graphene-Based Photocatalysts for Solar-Fuel Generation. Angew. Chem. 2015, 54, 11350–11366. 10.1002/anie.201411096. [DOI] [PubMed] [Google Scholar]
  5. Tahir M. Photocatalytic Carbon Dioxide Reduction to Fuels in Continuous Flow Monolith Photoreactor using Montmorillonite Dispersed Fe/TiO2Nanocatalyst. J. Clean. Prod. 2018, 170, 242–250. 10.1016/j.jclepro.2017.09.118. [DOI] [Google Scholar]
  6. Chen X.; Jin F. Photocatalytic Reduction of Carbon Dioxide by Titanium Based Semiconductors to Produce Fuels. Front. Energy Res. 2019, 13, 207–220. 10.1007/s11708-019-0628-9. [DOI] [Google Scholar]
  7. Kumar S.; Kumar A.; Kumar A.; Krishnan V. Nanoscale Zinc Oxide Based Heterojunctions as Visible Light Active Photocatalysts for Hydrogen Energy and Environmental Remediation. Catal. Rev. Sci. Eng. 2020, 62, 346–405. 10.1080/01614940.2019.1684649. [DOI] [Google Scholar]
  8. Yadav A. A.; Hunge Y. M.; Kang S. W. Spongy Ball-Like Copper Oxide Nanostructure Modified by Reduced Graphene Oxide for Enhanced Photocatalytic Hydrogen Production. Mater. Res. Bull. 2021, 133, 111026. 10.1016/j.materresbull.2020.111026. [DOI] [Google Scholar]
  9. Zhang Z.; Bian L.; Tian H.; Liu Y.; Bando Y.; Yamauchi Y.; Wang Z. L. Tailoring the Surface and Interface Structures of Copper-Based Catalysts for Electrochemical Reduction of CO2 to Ethylene and Ethanol. Small. 2022, 18, 2107450. 10.1002/smll.202107450. [DOI] [PubMed] [Google Scholar]
  10. Luo M.; Yang J.; Li X.; Eguchi M.; Yamauchi Y.; Wang Z. L. (2023). Insights into Alloy/Oxide or Hydroxide Interfaces in Ni-Mo-Based Electrocatalysts for Hydrogen Evolution Under Alkaline Conditions. Chem. Sci. 2023, 14, 3400–3414. 10.1039/D2SC06298D. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Guo C.; Tian K.; Wang L.; Liang F.; Wang F.; Chen D.; Ning J.; Zhong Y.; Hu Y. Approach of Fermi Level and Electron Tap-Level in Cadmium Sulfide Nanorods via Molybdenum Doping with Enhanced Carrier Separation for Boosted Photocatalytic Hydrogen Production. J. Colloid Interface Sci. 2021, 583, 661–671. 10.1016/j.jcis.2020.09.093. [DOI] [PubMed] [Google Scholar]
  12. Lyulyukin M. N.; Kurenkova A. Y.; Bukhtiyarov A. V.; Kozlova E. A. Carbon Dioxide Reduction Under Visible Light: A Comparison of Cadmium Sulfide and Titania Photocatalysts. Mendeleev Commun. 2020, 30, 192–194. 10.1016/j.mencom.2020.03.021. [DOI] [Google Scholar]
  13. Wang H.; Li S.; Wan Q.; Su X.; Song T.; Wang X.; Wang J. Highly Efficient Solution Exfoliation of Few-layered Molybdenum Disulfide Nanosheets for Photocatalytic Hydrogen Evolution. J. Colloid Interface Sci. 2020, 577, 38–47. 10.1016/j.jcis.2020.05.034. [DOI] [PubMed] [Google Scholar]
  14. Sasan K.; Lin Q.; Mao C.; Feng P. Open Framework Metal Chalcogenides as Efficient Photocatalysts for Reduction of CO2 into Renewable Hydrocarbon Fuel. Nanoscale. 2016, 8, 10913–10916. 10.1039/C6NR02525K. [DOI] [PubMed] [Google Scholar]
  15. Liu S.; Jiang T.; Fan M.; Tan G.; Cui S.; Shen X. Nanostructure Rod-like TiO2-Reduced Graphene Oxide Composite Aerogels for Highly Efficient Visible Light Photocatalytic CO2 Reduction. J. Alloys Compd. 2021, 861, 158598. 10.1016/j.jallcom.2021.158598. [DOI] [Google Scholar]
  16. Jiang L.; Li J.; Wang K.; Zhang G.; Li Y.; Wu X. Low Boiling Point Solvent Mediated Strategy to Synthesize Functionalized Monolayer Carbon Nitride for Superior Photocatalytic Hydrogen Evolution. Appl. Catal., B 2020, 260, 118181. 10.1016/j.apcatb.2019.118181. [DOI] [Google Scholar]
  17. Tian L.; Min S.; Wang F.; Zhang Z. Enhanced Photocatalytic Hydrogen Evolution on TiO2 Employing Vanadium Carbide as an Efficient and Stable Co-catalyst. Int. J. Hydrog. Energy. 2020, 45, 1878–1889. 10.1016/j.ijhydene.2019.11.094. [DOI] [Google Scholar]
  18. Wu Z.; Li C.; Li Z.; Feng K.; Cai M.; Zhang D.; Wang S.; Chu M.; Zhang C.; Shen J.; Huang Z.; Xiao Y.; Ozin G. A.; Zhang X.; He L. Niobium and Titanium Carbides (MXenes) as Superior Photothermal Supports for CO2 Photocatalysis. ACS Nano 2021, 15, 5696–5705. 10.1021/acsnano.1c00990. [DOI] [PubMed] [Google Scholar]
  19. Wang Z.-L.; Choi J.; Xu M.; Hao X.; Zhang H.; Jiang Z.; Zuo M.; Kim J.; Zhou W.; Meng X.; Yu Q.; Sun Z.; Wei S.; Ye J.; Wallace G. G.; Officer D. L.; Yamauchi Y. Optimizing Electron Densities of Ni-N-C Complexes by Hybrid Coordination for Efficient Electrocatalytic CO2 Reduction. ChemSusChem. 2020, 13, 929–937. 10.1002/cssc.201903427. [DOI] [PubMed] [Google Scholar]
  20. Duplock E. J.; Scheffler M.; Lindan P. J. Hallmark of Perfect Graphene. Phys. Rev. Lett. 2004, 92, 225502. 10.1103/PhysRevLett.92.225502. [DOI] [PubMed] [Google Scholar]
  21. Novoselov K. S.; Geim A. K.; Morozov S. V.; Jiang D.; Katsnelson M. I.; Grigorieva I. V.; Dubonos S. V.; Firsov A. A. Two-dimensional Gas of Massless Dirac Fermions in Graphene. Nature. 2005, 438, 197–200. 10.1038/nature04233. [DOI] [PubMed] [Google Scholar]
  22. Lou Q.; Lou G.; Peng R.; Liu Z.; Wang W.; Ji M.; Chen C.; Zhang X.; Liu C.; Ge Z. Synergistic Effect of Lewis Base Polymers and Graphene in Enhancing the Efficiency of Perovskite Solar Cells. ACS Appl. Energy Mater. 2021, 4, 3928–3936. 10.1021/acsaem.1c00299. [DOI] [Google Scholar]
  23. Reynard D.; Nagar B.; Girault H. Photonic Flash Synthesis of Mo2C/Graphene Electrocatalyst for the Hydrogen Evolution Reaction. ACS Catal. 2021, 11, 5865–5872. 10.1021/acscatal.1c00770. [DOI] [Google Scholar]
  24. Wang H.; Liu G.; Chen C.; Tu W.; Lu Y.; Wu S.; O’Hare D.; Xu R. Single-Ni Sites Embedded in Multilayer Nitrogen-Doped Graphene Derived from Amino-Functionalized MOF for Highly Selective CO2 Electroreduction. ACS Sustain. Chem. Eng. 2021, 9, 3792–3801. 10.1021/acssuschemeng.0c08749. [DOI] [Google Scholar]
  25. Liu L.; Dong R.; Ye D.; Lu Y.; Xia P.; Deng L.; Duan Y.; Cao K.; Chen S. Phosphomolybdic Acid-Modified Monolayer Graphene Anode for Efficient Organic and Perovskite Light-Emitting Diodes. ACS Appl. Mater. Interfaces. 2021, 13, 12268–12277. 10.1021/acsami.0c22456. [DOI] [PubMed] [Google Scholar]
  26. Tang Z.; Liu C.; Huang X.; Zeng S.; Liu L.; Li J.; Jiang Y.-G.; Zhang D. W.; Zhou P. A Steep-Slope MoS2/Graphene Dirac-Source Field-Effect Transistor with a Large Drive Current. Nano Lett. 2021, 21, 1758–1764. 10.1021/acs.nanolett.0c04657. [DOI] [PubMed] [Google Scholar]
  27. Wang Y.; Ho V. X.; Pradhan P.; Cooney M. P.; Vinh N. Q. Interfacial Photogating Effect for Hybrid Graphene-Based Photodetectors. ACS Appl. Nano Mater. 2021, 4, 8539–8545. 10.1021/acsanm.1c01931. [DOI] [Google Scholar]
  28. Xie J.; Zhang H.; Li S.; Wang R.; Sun X.; Zhou M.; Zhou J.; Lou X. W. D.; Xie Y. Defect-Rich MoS2 Ultra-Thin Nanosheets with Additional Active Edge Sites for Enhanced Electrocatalytic Hydrogen Evolution. Adv. Mater. 2013, 25, 5807–5813. 10.1002/adma.201302685. [DOI] [PubMed] [Google Scholar]
  29. Das S.; Chen H. Y.; Penumatcha A. V.; Appenzeller J. High-Performance Multilayer MoS2Transisters with Scandium Contacts. J. Nano Lett. 2013, 13, 100–105. 10.1021/nl303583v. [DOI] [PubMed] [Google Scholar]
  30. Benck J. D.; Hellstern T. R.; Kibsgaard J.; Chakthranont P.; Jaramillo T. F. Catalyzing the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide Nanomaterials. ACS Catalysis. 2014, 4, 3957–3971. 10.1021/cs500923c. [DOI] [Google Scholar]
  31. Xing F.; Wang C.; Liu S.; Jin S.; Jin H.; Li J. Interfacial Chemical Bond Engineering in a Direct Z-Scheme g-C3N4/MoS2 Heterojunction. ACS Appl. Mater. Interfaces. 2023, 15, 11731–11740. 10.1021/acsami.2c21046. [DOI] [PubMed] [Google Scholar]
  32. Chen L.; Arshad M.; Chuang Y.; Nguyen T. B.; Wu C. H.; Chen C. W.; Dong C. D. A Novel Nano-heterojunction MoS2/α-Fe2O3 Photocatalysts with High Photocatalytic and Photoelectrochemical Performance under Visible Light Irradiation. J. Alloys Compd. 2023, 947, 169577. 10.1016/j.jallcom.2023.169577. [DOI] [Google Scholar]
  33. Wu C.; Huang W.; Liu H.; Lv K.; Li Q. Insight into Synergistic Effect of Ti3C2 MXene and MoS2 on Anti-Photocorrosion and Photocatalytic of CdS for Hydrogen Production. Appl. Catal. B: Environ. 2023, 330, 122653. 10.1016/j.apcatb.2023.122653. [DOI] [Google Scholar]
  34. Wang Y.; Yang C.; Guo L.; Yang Z.; Jin B.; Du R.; Fu F.; Wang D. Plate-on-plate Structured MoS2/Cd0.6Zn0.4S Z-Scheme Heterostructure with Enhanced Photocatalytic Hydrogen Production Activity via Hole Sacrificial Agent Synchronously Strengthen Half-Reactions. J. Colloid Interface Sci. 2023, 630, 341–351. 10.1016/j.jcis.2022.10.053. [DOI] [PubMed] [Google Scholar]
  35. Yang X.; Lan X.; Zhang Y.; Li H.; Bai G. Rational Design of MoS2@COF Hybrid Composites Promoting CC Coupling for Photocatalytic CO2 Reduction to Ethane. Appl. Catal. B: Environ. 2023, 325, 122393. 10.1016/j.apcatb.2023.122393. [DOI] [Google Scholar]
  36. Nair Gopalakrishnan V.; Becerra J.; Mohan S.; Ahad J. M.; Béland F.; Do T. O. Cobalt-Doped MoS2-Integrated Hollow Structured Covalent Organic Framework Nanospheres for the Effective Photoreduction of CO2 under Visible Light. Energy amp; Fuels 2023, 37, 2329–2339. 10.1021/acs.energyfuels.2c03292. [DOI] [Google Scholar]
  37. Hisatomi T.; Kubota J.; Domen K. Recent Advances in Semiconductors for Photocatalytic and Photoelectrochemical Water Splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. 10.1039/C3CS60378D. [DOI] [PubMed] [Google Scholar]
  38. Tollefson J. Hydrogen Vehicles: Fuel of the Future?. Nature News. 2010, 464, 1262–1264. 10.1038/4641262a. [DOI] [PubMed] [Google Scholar]
  39. Yuan Y. J.; Lu H. W.; Yu Z. T.; Zou Z. G. Noble-Metal-Free Molybdenum Disulfide Cocatalyst for Photocatalytic Hydrogen Production. ChemSusChem. 2015, 8, 4113–4127. 10.1002/cssc.201501203. [DOI] [PubMed] [Google Scholar]
  40. Corredor J.; Rivero M. J.; Rangel C. M.; Gloaguen F.; Ortiz I. Comprehensive Review and Future Perspectives on the Photocatalytic Hydrogen Production. J. Chem. Technol. Biotechnol. 2019, 94, 3049–3063. 10.1002/jctb.6123. [DOI] [Google Scholar]
  41. Fujishima A.; Honda K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature. 1972, 238, 37–38. 10.1038/238037a0. [DOI] [PubMed] [Google Scholar]
  42. Sharma R.; Almasi M.; Nehra S. P.; Rao V. S.; Panchal P.; Paul D. R.; Jain I. P.; Sharma A. Photocatalytic Hydrogen Production using Graphitic Carbon Nitride (GCN): A Precise Review. Renew. Sust. Energy Rev. 2022, 168, 112776. 10.1016/j.rser.2022.112776. [DOI] [Google Scholar]
  43. Han B.; Hu Y. H. MoS2 as a Co-catalyst for Photocatalytic Hydrogen Production from Water. Energy Sci. Eng. 2016, 4, 285–304. 10.1002/ese3.128. [DOI] [Google Scholar]
  44. Ding Q.; Song B.; Xu P.; Jin S. Efficient Electrocatalytic and Photoelectrochemical Hydrogen Generation Using MoS2 and Related Compounds. Chem. 2016, 1, 699–726. 10.1016/j.chempr.2016.10.007. [DOI] [Google Scholar]
  45. Xu J.; Shao G.; Tang X.; Lv F.; Xiang H.; Jing C.; Zhou Z. Frenkel-Defected Monolayer MoS2 Catalysts for Efficient Hydrogen Evolution. Nat. Commun. 2022, 1, 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Eknapakul T.; King P. D. C.; Asakawa M.; Buaphet P.; He R.-H.; Mo S.-K.; Takagi H.; Shen K. M.; Baumberger F.; Sasagawa T.; Jungthawan S.; Meevasana W. Electronic Structure of a Quasi-Freestanding MoS2 Monolayer. Nano Lett. 2014, 14, 1312–1316. 10.1021/nl4042824. [DOI] [PubMed] [Google Scholar]
  47. Ding Q.; Song B.; Xu P.; Jin S. Efficient Electrocatalytic and Photoelectrochemical Hydrogen Generation Using MoS2 and Related Compounds. Chem. 2016, 1, 699–726. 10.1016/j.chempr.2016.10.007. [DOI] [Google Scholar]
  48. Xiang Q.; Yu J.; Jaroniec M. Synergetic Effect of MoS2 and Graphene as Co-catalysts for Enhanced Photocatalytic H2 Production Activity of TiO2 Nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575–6578. 10.1021/ja302846n. [DOI] [PubMed] [Google Scholar]
  49. Wang Y.; Wang Q.; Zhan X.; Wang F.; Safdar M.; He J. Visible Light-Driven Type II Heterostructures and their Enhanced Photocatalysis Properties: A Review. Nanoscale. 2013, 5, 8326–8339. 10.1039/c3nr01577g. [DOI] [PubMed] [Google Scholar]
  50. Kadantsev E. S.; Hawrylak P. Electronic Structure of a Single MoS2 Monolayer. Solid State Commun. 2012, 152, 909–913. 10.1016/j.ssc.2012.02.005. [DOI] [Google Scholar]
  51. Wang F.; Shifa T. A.; Zhan X.; Huang Y.; Liu K.; Cheng Z.; Jiang C.; He J. Recent Advances in Transition-metal Dichalcogenide Based Nanomaterials for Water Splitting. Nanoscale. 2015, 7, 19764–19788. 10.1039/C5NR06718A. [DOI] [PubMed] [Google Scholar]
  52. Kubacka A.; Fernandez-Garcia M.; Colon G. Advanced Nanoarchiectures for Solar Photocatalytic Applications. Chem. Rev. 2012, 112, 1555–1614. 10.1021/cr100454n. [DOI] [PubMed] [Google Scholar]
  53. Hinnemann B.; Moses P. G.; Bonde J.; Jørgensen K. P.; Nielsen J. H.; Horch S.; Chorkendorff I.; Nørskov J. K. Biomimetic Hydrogen Evolution: MoS2 Nanoparticles as Catalysts for Hydrogen Evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. 10.1021/ja0504690. [DOI] [PubMed] [Google Scholar]
  54. Bonde J.; Moses P. G.; Jaramillo T. F.; Nørskov J. K.; Chorkendorff I. Hydrogen Evolution on Nano-particulate Transition Metal Sulfides. Faraday Discuss. 2009, 140, 219–231. 10.1039/B803857K. [DOI] [PubMed] [Google Scholar]
  55. Tsai C.; Chan K.; Abild-Pedersen F.; Nørskov J. K. Active Edge-Sites in MoSe2 and WSe2 Catalysts for Hydrogen Evolution Reaction: A Density Functional Study. Phys. Chem. Chem. Phys. 2014, 16, 13156–13164. 10.1039/C4CP01237B. [DOI] [PubMed] [Google Scholar]
  56. Jaramillo T. F.; Jørgensen K. P.; Bonde J.; Nielsen J. H.; Horch S.; Chorkendorff I. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science. 2007, 317, 100–102. 10.1126/science.1141483. [DOI] [PubMed] [Google Scholar]
  57. Khalid N. R.; Israr Z.; Tahir M. B.; Iqbal T. Highly Efficient Bi2O3/MoS2 on Heterojunction Photocatalysts for Hydrogen Evolution from Water Splitting. Int. J. Hydrog. Energy. 2020, 45, 8479–8489. 10.1016/j.ijhydene.2020.01.031. [DOI] [Google Scholar]
  58. Pramoda K.; Servottam S.; Kaur M.; Rao C. N. R. Layered Nanocomposites of Polymer-Functionalized Reduced Graphene Oxide and Borocarbonitride with MoS2 and MoSe2 and Their Hydrogen Evolution Reaction Activity. ACS Applied Nano Materials. 2020, 3, 1792–1799. 10.1021/acsanm.9b02482. [DOI] [Google Scholar]
  59. Wang W.; Zhu S.; Cao Y.; Tao Y.; Li X.; Pan D.; Phillips D. L.; Zhang D.; Chen M.; Li G.; Li H. Edge Enriched Ultrathin MoS2 Embedded Yolk-Shell TiO2 with Boosted Charge Transfer for Superior Photocatalytic H2 Evolution. Adv. Funct. Mater. 2019, 29, 1901958. 10.1002/adfm.201901958. [DOI] [Google Scholar]
  60. Zhou X.; Licklederer M.; Schmuki P. Thin MoS2 on TiO2 Nanotube Layers: An Efficient Co-catalyst/Harvesting System for Photocatalytic H2 Evolution. Electrochem. Commun. 2016, 73, 33–37. 10.1016/j.elecom.2016.10.008. [DOI] [Google Scholar]
  61. Agrawal A. V.; Kumar R.; Venkatesan S.; Zakhidov A.; Yang G.; Bao J.; Kumar M.; Kumar M. Photoactivated Mixed In-Plane and Edge-Enriched p-Type MoS2 Flake-Based NO2 Sensor Working at Room Temperature. ACS Sens. 2018, 3, 998–1004. 10.1021/acssensors.8b00146. [DOI] [PubMed] [Google Scholar]
  62. Zhu Y.; Ling Q.; Liu Y.; Wang H.; Zhu Y. Photocatalytic H2 Evolution on MoS2-TiO2 Catalysts Synthesized via Mechanochemistry. Phys. Chem. Chem. Phys. 2015, 17, 933–940. 10.1039/C4CP04628E. [DOI] [PubMed] [Google Scholar]
  63. Kumar S.; Kumar A.; Navakoteswara Rao V.; Kumar A.; Shankar M. V.; Krishnan V. Defect-Rich MoS2 Ultrathin Nanosheets-Coated Nitrogen-Doped ZnO Nanorod Heterostructures: An Insight into in-situ-Generated ZnS for Enhanced Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2019, 2, 5622–5634. 10.1021/acsaem.9b00790. [DOI] [Google Scholar]
  64. Yuan Y.-J.; Wang F.; Hu B.; Lu H. W.; Yu Z. T.; Zou Z. G. Significant Enhancement in Photocatalytic Hydrogen Evolution from Water Using MoS2-Nanosheet-Coated ZnO Heterostructure Photocatalyst. Dalton Trans. 2015, 44, 10997–11003. 10.1039/C5DT00906E. [DOI] [PubMed] [Google Scholar]
  65. Yuan Y.-J.; Tu J.-R.; Ye Z.-J.; Lu H.-W.; Ji Z.-G.; Hu B.; Li Y.-H.; Cao D.-P.; Yu Z.-T.; Zou Z.-G. Visible-Light-Driven Hydrogen Production from in a Noble-Metal-Free System Catalyzed by Zinc Porphyrin Sensitized MoS2/ZnO. Dyes Pigm. 2015, 123, 285–292. 10.1016/j.dyepig.2015.08.014. [DOI] [Google Scholar]
  66. Shu X.; Zhao C.; Zhang X.; Liu X.; Xing Z.; Fang D.; Wang J.; Song Y. Highly Efficient Near-Infrared Light Photocatalytic Hydrogen Evolution Over MoS2 Supported Ta2O5 Combined with an Up-Conversion Luminescence Agent β-Tm3+, Yb3+: NaYF4/MoS2-Ta2O5 Nanocomposite. New J. Chem. 2018, 42, 6134–6143. 10.1039/C7NJ04993E. [DOI] [Google Scholar]
  67. Goud B. S.; Koyyada G.; Jung J. H.; Reddy G. R.; Shim J.; Nam N. D.; Vattikuti S. P. Surface Oxygen Vacancy Facilitated Z-Scheme MoS2/Bi2O3 Heterojunction for Enhanced Visible-Light Driven Photocatalysis-Pollutant Degradation and Hydrogen Production. Int. J. Hydrog. Energy. 2020, 45, 18961–18975. 10.1016/j.ijhydene.2020.05.073. [DOI] [Google Scholar]
  68. Ye L.; Xu H.; Zhang D.; Chen S. Synthesis of Bilayer MoS2 Nanosheets by a Facile Hydrothermal Method, and their Methyl Orange Adsorption Capacity. Mater. Res. Bull. 2014, 55, 221–228. 10.1016/j.materresbull.2014.04.025. [DOI] [Google Scholar]
  69. Opoku F.; Govender K. K.; van Sittert C. G. C. E.; Govender P. P. Role of MoS2 and WS2 Monolayers on Photocatalytic Hydrogen Production and the Pollutant Degradation of BiVO4: A First-principles Study. New J. Chem. 2017, 41, 11701–11713. 10.1039/C7NJ02340E. [DOI] [Google Scholar]
  70. Swain G.; Sultana S.; Naik B.; Parida K. Coupling of Crumpled-Type Novel MoS2 with CeO2 Nanoparticles: A Noble-Metal-Free p-n Heterojunction Composite for Visible Light Photocatalytic H2 Production. ACS Omega. 2017, 2, 3745–3753. 10.1021/acsomega.7b00492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Deng D.; Novoselov K. S.; Fu Q.; Zheng N.; Tian Z.; Bao X. Catalysis with Two-Dimensional Materials and their Heterostructures. Nat. Nanotechnol. 2016, 11, 218–230. 10.1038/nnano.2015.340. [DOI] [PubMed] [Google Scholar]
  72. Yin M.; Sun J.; Li Y.; Ye Y.; Liang K.; Fan Y.; Li Z. Efficient photocatalytic Hydrogen Evolution Over MoS2/Activated Carbon Composite Sensitized by Erythrosin B Under LED Light Irradiation. Catal. Commun. 2020, 142, 106029. 10.1016/j.catcom.2020.106029. [DOI] [Google Scholar]
  73. Koutsouroubi E. D.; Vamvasakis I.; Papadas I. T.; Drivas C.; Choulis S. A.; Kennou S.; Armatas G. S. Interface Engineering of MoS2-Modified Graphitic Carbon Nitride Nano-Photocatalysts for an Efficient Hydrogen Evolution Reaction. ChemPlusChem. 2020, 85, 1379–1388. 10.1002/cplu.202000096. [DOI] [PubMed] [Google Scholar]
  74. Li N.; Zhou J.; Sheng Z.; Xiao W. Molten Salt-Mediated formation of g-C3N4-MoS2 for Visible-Light-Driven Photocatalytic Hydrogen Evolution. Appl. Surf. Sci. 2018, 430, 218–224. 10.1016/j.apsusc.2017.08.086. [DOI] [Google Scholar]
  75. Zheng D.; Zhang G.; Hou Y.; Wang X. Layering MoS2 on Soft Hollow g-C3N4 Nanostructures for Photocatalytic Hydrogen Evolution. APPL CATAL A-GEN. 2016, 521, 2–8. 10.1016/j.apcata.2015.10.037. [DOI] [Google Scholar]
  76. Ge L.; Han C.; Xiao X.; Guo L. Synthesis and Characterization of Composite Visible Light Active Photocatalysts MoS2-g-C3N4 with Enhanced Hydrogen Evolution Activity. Int. J. Hydrog. Energy. 2013, 38, 6960–6969. 10.1016/j.ijhydene.2013.04.006. [DOI] [Google Scholar]
  77. Liu Y.; Xu X.; Zhang J.; Zhang H.; Tian W.; Li X.; Tade M. O.; Sun H.; Wang S. Flower-Like MoS2 on Graphitic Carbon Nitride for Enhanced Photocatalytic and Electrochemical Hydrogen Evolutions. Appl. Catal., B 2018, 239, 334–344. 10.1016/j.apcatb.2018.08.028. [DOI] [Google Scholar]
  78. Yuan Y.-J.; Shen Z.; Wu S.; Su Y.; Pei L.; Ji Z.; Ding M.; Bai W.; Chen Y.; Yu Z.-T.; Zou Z. Liquid Exfoliation of g-C3N4 Nanosheets to Construct 2D-2D MoS2/g-C3N4 Photocatalysts for Enhanced Photocatalytic Hydrogen Production Activity. Appl. Catal., B 2019, 246, 120–128. 10.1016/j.apcatb.2019.01.043. [DOI] [Google Scholar]
  79. Li X.; Guo S.; Li W.; Ren X.; Su J.; Song Q.; Sobrido A. J.; Wei B. Edge-Rich MoS2 Grown on Edge-Oriented Three-Dimensional Graphene Glass for High-Performance Hydrogen Evolution. Nano Energy. 2019, 57, 388–397. 10.1016/j.nanoen.2018.12.044. [DOI] [Google Scholar]
  80. Masimukku S.; Tsai D. L.; Lin Y. T.; Chang I. L.; Wu J. J. Low-Frequency Vibration Induced Piezoelectric Boost to Photocatalytic Hydrogen Evolution through 2D-2D-Stacked MoS2-Carbon Nitride. Appl. Surf. Sci. 2023, 614, 156147. 10.1016/j.apsusc.2022.156147. [DOI] [Google Scholar]
  81. Li Y.; Chen G.; Zhou C.; Sun J. A Simple-Template-Free Synthesis of Nanoporous ZnS-In2S3-Ag2S Solid Solutions for Highly Efficient Photocatalytic H2 Evolution Under Visible Light. ChemComm. 2009, 15, 2020–2022. 10.1039/b819300b. [DOI] [PubMed] [Google Scholar]
  82. Kadam S. R.; Panmand R. P.; Sonawane R. S.; Gosavi S. W.; Kale B. B. A Stable Bi2S3-Quantum Dot-Glass Nanosystem: Size Tunable Photocatalytic Hydrogen Production Under Solar Light. RSC Adv. 2015, 5, 58485–58490. 10.1039/C5RA10244H. [DOI] [Google Scholar]
  83. Wang Q.; An N.; Bai Y.; Hang H.; Li J.; Lu X.; Liu Y.; Wang F.; Li Z.; Lei Z. High Photocatalytic Hydrogen Production from Methanol Aqueous Solution Using the Photocatalysts CuS/TiO2. Int. J. Hydrog. Energy. 2013, 38, 10739–10745. 10.1016/j.ijhydene.2013.02.131. [DOI] [Google Scholar]
  84. Li L.-L.; Yin X.-L.; Pang D.-H.; Du X.-X.; Xue H.; Zhou H.-W.; Yao Q.-X.; Wang H.-W.; Qian J.-C.; Yang J.; Li D.-C.; Dou J.-M. One-Pot Synthesis of MoS2/CdS Nanosphere Heterostructures for Efficient H2 Evolution Under Visible Light Irradiation. Int. J. Hydrog. Energy. 2019, 44, 31930–31939. 10.1016/j.ijhydene.2019.10.056. [DOI] [Google Scholar]
  85. Lv Y.; Chen P.; Foo J. J.; Zhang J.; Qian W.; Chen C.; Ong W. J. Dimensionality-Dependent MoS2 toward Efficient Photocatalytic Hydrogen Evolution: From Synthesis to Modifications in Doping, Surface, and Heterojunction Engineering. Mater. Today Nano. 2022, 18, 100191. 10.1016/j.mtnano.2022.100191. [DOI] [Google Scholar]
  86. Xu X.; Luo F.; Zhou G.; Hu J.; Zeng H.; Zhou Y. Self-Assembly Optimization of Cadmium/Molybdenum Sulfide by Cation Coordination Competition Toward Extraordinarily Efficient Photocatalytic Hydrogen Evolution. J. Mater. Chem. A 2018, 6, 18396–18402. 10.1039/C8TA07911K. [DOI] [Google Scholar]
  87. Ma S.; Xie J.; Wen J.; He K.; Li X.; Liu W.; Zhang X. Constructing 2D Layered Hybrid CdS Nanosheets/MoS2 Heterojunctions for Enhanced Visible-Light Photocatalytic H2 Generation. Appl. Surf. Sci. 2017, 391, 580–591. 10.1016/j.apsusc.2016.07.067. [DOI] [Google Scholar]
  88. Li C.; Wang H.; Ming J.; Liu M.; Fang P. Hydrogen Generation by Photocatalytic Reforming of Glucose with Heterostructured CdS/MoS2 Composites Under Visible Light Irradiation. Int. J. Hydrog. Energy. 2017, 42, 16968–16978. 10.1016/j.ijhydene.2017.05.137. [DOI] [Google Scholar]
  89. Kumar D. P.; Seo S.; Rangappa A. P.; Kim S.; Joshi Reddy K. A.; Gopannagari M.; Bhavani P.; Reddy D. A.; Kim T. K. Ultrathin layered Zn-doped MoS2 nanosheets deposited onto CdS nanorods for spectacular photocatalytic hydrogen evolution. J. Alloys Compd. 2022, 905, 164193. 10.1016/j.jallcom.2022.164193. [DOI] [Google Scholar]
  90. Guo S.; Yang L.; Zhang Y.; Huang Z.; Ren X.; Sha W. E. I.; Li X. Enhanced Hydrogen Evolution via Interlaced Ni3S2/MoS2 Heterojunction Photocatalysts with Efficient Interfacial Contact and Broadband Absorption. J. Alloys Compd. 2018, 749, 473–480. 10.1016/j.jallcom.2018.03.329. [DOI] [Google Scholar]
  91. Zhao H.; Fu H.; Yang X.; Xiong S.; Han D.; An X. MoS2/CdS Rod-like Nanocomposites as High-Performance Visible Light Photocatalyst for Water Splitting Photocatalytic Hydrogen Production. Int. J. Hydrog. Energy. 2022, 47, 8247–8260. 10.1016/j.ijhydene.2021.12.171. [DOI] [Google Scholar]
  92. Zhang K.; Mou Z.; Cao S.; Meng T.; Li J.; Ling G.; Zhang X.; Zhou Z.; Meng W. MoS2 Grown in-situ on CdS Nanosheets for Boosted Photocatalytic Hydrogen Evolution under Visible Light. Int. J. Hydrog. Energy. 2022, 47, 2967–2975. 10.1016/j.ijhydene.2021.10.256. [DOI] [Google Scholar]
  93. Huang L.; Han B.; Huang X.; Liang S.; Deng Z.; Chen W.; Peng M.; Deng H. Ultrathin 2D/2D ZnIn2S4/MoS2 Hybrids for Boosted Photocatalytic Hydrogen Evolution Under Visible Light. J. Alloys Compd. 2019, 798, 553–559. 10.1016/j.jallcom.2019.05.162. [DOI] [Google Scholar]
  94. Li W.; Lin Z.; Yang G. A 2D Self-Assembled MoS2/ZnIn2S4 Heterostructure for Efficient Photocatalytic Hydrogen Evolution. Nanoscale. 2017, 9, 18290–18298. 10.1039/C7NR06755K. [DOI] [PubMed] [Google Scholar]
  95. Wang Y.; Yang C.; Guo L.; Yang Z.; Jin B.; Du R.; Fu F.; Wang D. Plate-on-Plate Structured MoS2/Cd0.6Zn0.4S Z-Scheme Heterostructure with Enhanced Photocatalytic Hydrogen Production Activity via Hole Sacrificial Agent Synchronously Strengthen Half-reactions. J. Colloid Interface Sci. 2023, 630, 341–351. 10.1016/j.jcis.2022.10.053. [DOI] [PubMed] [Google Scholar]
  96. Fang H.; Cai J.; Li H.; Wang J.; Li Y.; Zhou W.; Mao K.; Xu Q. Fabrication of Ultrathin Two-Dimensional/Two-Dimensional MoS2/ZnIn2S4 Hybrid Nanosheets for Highly Efficient Visible-Light-Driven Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2022, 5, 8232–8240. 10.1021/acsaem.2c00767. [DOI] [Google Scholar]
  97. Chand M.; Rawat A. S.; Khanuja M.; Rawat S. Hydrogen production activity of MoS2-ZnIn2S4 Nanocomposite under Visible Light Irradiation. Mater. Today: Proc. 2022, 66, 1951–1954. 10.1016/j.matpr.2022.05.429. [DOI] [Google Scholar]
  98. Zhang Z.; Huang L.; Zhang J.; Wang F.; Xie Y.; Shang X.; Gu Y.; Zhao H.; Wang X. In Situ Constructing Interfacial Contact MoS2/ZnIn2S4 Heterostructure for Enhancing Solar Photocatalytic Hydrogen Evolution. Appl. Catal., B 2018, 233, 112–119. 10.1016/j.apcatb.2018.04.006. [DOI] [Google Scholar]
  99. Huang T.; Chen W.; Liu T. Y.; Hao Q. L.; Liu X. H. ZnIn2S4 Hybrid with MoS2: A Non-noble Metal Photocatalyst with Efficient Photocatalytic Activity for Hydrogen Evolution. Powder Technol. 2017, 315, 157–162. 10.1016/j.powtec.2017.03.054. [DOI] [Google Scholar]
  100. Tian G.; Chen Y.; Ren Z.; Tian C.; Pan K.; Zhou W.; Fu H. Enhanced Photocatalytic Hydrogen Evolution over Hierarchical Composites of ZnIn2S4 Nanosheets Grown on MoS2 Slices. Chem.—Asian J. 2014, 9, 1291–1297. 10.1002/asia.201301646. [DOI] [PubMed] [Google Scholar]
  101. Wei L.; Chen Y.; Lin Y.; Wu H.; Yuan R.; Li Z. MoS2 as Non-Noble-Metal Co-catalyst for Photocatalytic Hydrogen Evolution over Hexagonal ZnIn2S4 Under Visible Light Irradiations. Appl. Catal., B 2014, 144, 521–527. 10.1016/j.apcatb.2013.07.064. [DOI] [Google Scholar]
  102. Swain G.; Parida K. Design of a Novel p-MoS2@n-ZnIn2S4 Heterojunction Based Semiconducting Photocatalyst Towards Photocatalytic HER. Mater. Today: Proc. 2021, 35, 268–274. 10.1016/j.matpr.2020.05.756. [DOI] [Google Scholar]
  103. Pudkon W.; Bahruji H.; Miedziak P. J.; Davies T. E.; Morgan D. J.; Pattisson S.; Kaowphong S.; Hutchings G. J. Enhanced Visible-Light-Driven Photocatalytic H2 Production and Cr (VI) Reduction of a ZnI2S4/MoS2 Heterojunction Synthesized by the Bio-molecule Assisted Microwave Heating Method. Catal. Sci. Technol. 2020, 10, 2838–2854. 10.1039/D0CY00234H. [DOI] [Google Scholar]
  104. Swain G.; Sultana S.; Parida K. Constructing a Novel Surfactant Free MoS2 Nanosheet Modified MgIn2S4 Marigold Microflower: An Efficient Visible-Light-Driven 2D-2D p-n Heterojunction Photocatalyst Toward HER and pH Regulated NRR. ACS Sustain. Chem. Eng. 2020, 8, 4848–4862. 10.1021/acssuschemeng.9b07821. [DOI] [Google Scholar]
  105. Swain G.; Sultana S.; Moma J.; Parida K. Fabrication of Hierarchical Two-Dimensional MoS2 Nanoflowers Decorated upon Cubic CaIn2S4 Microflowers: Facile Approach to Construct Novel Metal-Free p-n Heterojunction Semiconductors with Superior Charge Separation Efficiency. Inorg. Chem. 2018, 57, 10059–10071. 10.1021/acs.inorgchem.8b01221. [DOI] [PubMed] [Google Scholar]
  106. Fan K.; Jin Z.; Wang G.; Yang H.; Liu D.; Hu H.; Lu G.; Bi Y. Distinctive Organized Molecular Assemble of MoS2, MOF, and Co3O4, for Efficient Dye-Sensitized Photocatalytic H2 Evolution. Catal. Sci. Technol. 2018, 8, 2352–2363. 10.1039/C8CY00380G. [DOI] [Google Scholar]
  107. Jin Z.; Wang Z.; Yuan H.; Han F. Inseting MOF into flaky CdS Photocatalyst Forming Special Structure and Active Sites for Efficient Hydrogen Production. Int. J. Hydrog. Energy. 2019, 44, 19640–19649. 10.1016/j.ijhydene.2019.05.233. [DOI] [Google Scholar]
  108. Lin L.; Huang S.; Zhu Y.; Du B.; Zhang Z.; Chen C.; Wang X.; Zhang N. Construction of CdS/MoS2 Heterojunction from Core-Shell MoS2@Cd MOF for Efficient Photocatalytic Hydrogen Evolution. Dalton Trans. 2019, 48, 2715–2721. 10.1039/C8DT04745F. [DOI] [PubMed] [Google Scholar]
  109. Qiu J.; Zheng W.; Yuan R.; Yue C.; Li D.; Liu F.; Zhu J. A Novel 3D Nanofibrous Aerogel-Based MoS2@Co3S4 Heterojunction Photocatalyst for Water Remediation and Hydrogen Evolution Under Stimulated Solar Irradiation. Appl. Catal., B 2020, 264, 118514. 10.1016/j.apcatb.2019.118514. [DOI] [Google Scholar]
  110. Ma B.; Chen T.-T.; Li Q.-Y.; Qin H.-N.; Dong X.-Y.; Zang S.-Q. Bimetal-Organic-Framework-Derived Nanohybrids Cu0.9Co2.1S4@MoS2 for High-Performance Visible-Light-Catalytic Hydrogen Evolution Reaction. ACS Appl. Energy Mater. 2019, 2, 1134–1148. 10.1021/acsaem.8b01691. [DOI] [Google Scholar]
  111. Qiao Z.; Wang W.; Liu N.; Huang H.-T.; Karuppasamy L.; Yang H.-J.; Liu C.-H.; Wu J. J. Synthesis of MOF/MoS2 Composite Photocatalysts with Enhanced Photocatalytic Performance for Hydrogen Evolution from Water Splitting. Int. J. Hydrog. Energy. 2022, 47, 40755–40767. 10.1016/j.ijhydene.2021.10.021. [DOI] [Google Scholar]
  112. Tahir M. B. Construction of MoS2/CND-WO3 Ternary Composite for Photocatalytic Hydrogen Evolution. J. Inorg. Organomet. Polym. Mater. 2018, 28, 2160–2168. 10.1007/s10904-018-0867-y. [DOI] [Google Scholar]
  113. Kumar D. P.; Hong S.; Reddy D. A.; Kim T. K. Ultrathin MoS2 Layers Anchored Exfoliated Reduced Graphene Oxide Nanosheet Hybrid as a Highly Efficient Co-catalyst for CdS Nanorods Towards Enhanced Photocatalytic Hydrogen Production. Appl. Catal., B 2017, 212, 7–14. 10.1016/j.apcatb.2017.04.065. [DOI] [Google Scholar]
  114. Zhu C.; Wang Y.; Jiang Z.; Xu F.; Xian Q.; Sun C.; Tong Q.; Zou W.; Duan X.; Wang S. CeO2 Nanocrystal Modified Layered MoS2/g-C3N4 as 0D/2D Ternary Composite for Visible Light Photocatalytic Hydrogen Evolution: Interfacial Consecutive Multy-step Electron Transfer and Enhanced H2O Reactant Adsorption. Appl. Catal., B 2019, 259, 118072. 10.1016/j.apcatb.2019.118072. [DOI] [Google Scholar]
  115. Xu X.; Si Z.; Liu L.; Wang Z.; Chen Z.; Ran R.; He Y.; Weng D. Co-MoS2/rGO/C3N4 Ternary Heterojunctions Catalysts with High Photocatalytic Activity and Stability for Hydrogen Evolution Under Visible Light Irradiation. Appl. Surf. Sci. 2018, 435, 1296–1306. 10.1016/j.apsusc.2017.12.001. [DOI] [Google Scholar]
  116. Kumar S.; Reddy N. L.; Kushwaha H. S.; Kumar A.; Shankar M. V.; Bhattacharyya K.; Halder A.; Krishnan V. Efficient Electron Transfer Across a ZnO-MoS2-Reduced Graphene Oxide Heterojunction for Enhanced-Sunlight Driven Photocatalytic Hydrogen Evolution. ChemSusChem. 2017, 10, 3588–3603. 10.1002/cssc.201701024. [DOI] [PubMed] [Google Scholar]
  117. Reddy D. A.; Park H.; Ma R.; Kumar D. P.; Lim M.; Kim T. K. Heterostructured WS2-MoS2 Ultrathin nanosheets Integrated on CdS Nanorods to Promote Charge Separation and Migration and Improve Solar-Driven Photocatalytic Hydrogen Evolution. ChemSusChem. 2017, 10, 1563–1570. 10.1002/cssc.201601799. [DOI] [PubMed] [Google Scholar]
  118. Choi J.; Amaranatha Reddy D.; Han N. S.; Jeong S.; Hong S.; Praveen Kumar D.; Song J. K.; Kim T. K. Modulation of Charge Carrier Pathways in CdS Nanospheres by Integrating MoS2 and NiP for Improved Migration and Separation Toward Enhanced Photocatalytic Hydrogen Evolution. Catal. Sci. Technol. 2017, 7, 641–649. 10.1039/C6CY02145J. [DOI] [Google Scholar]
  119. Yu X.; Du R.; Li B.; Zhang Y.; Liu H.; Qu J.; An X. Biomolecule-Assisted Self-Assembly of CdS/MoS2/Graphene Hollow Spheres as High-Efficiency Photocatalysts for Hydrogen Evolution without Noble Metals. Applied Catalysis B: Environmental. 2016, 182, 504–512. 10.1016/j.apcatb.2015.09.003. [DOI] [Google Scholar]
  120. Sasikala R.; Gaikwad A. P.; Jayakumar O. D.; Girija K. G.; Rao R.; Tyagi A. K.; Bharadwaj S. R. Nanohybrid MoS2-PANI-CdS Photocatalyst for Hydrogen Evolution from Water. Colloids Surf, A PhysicochemEng. Asp. 2015, 481, 485–492. 10.1016/j.colsurfa.2015.06.027. [DOI] [Google Scholar]
  121. Sun J.; Maimaiti H.; Zhai P.; Zhang H.; Feng L.; Bao J.; Zhao X. Preparation of a Coal-Based MoS2/SiO2/GO Composite Catalyst and Its Performance in the Photocatalytic Degradation of Wastewater and Hydrogen Production. Langmuir. 2022, 38, 3305–3315. 10.1021/acs.langmuir.2c00163. [DOI] [PubMed] [Google Scholar]
  122. Dong J.; Fang W.; Yuan H.; Xia W.; Zeng X.; Shangguan W. Few-Layered MoS2/ZnCdS/ZnS Heterostructures with an Enhanced Photocatalytic Hydrogen Evolution. ACS Applied Energy Materials. 2022, 5, 4893–4902. 10.1021/acsaem.2c00301. [DOI] [Google Scholar]
  123. Wang J.; Zhu H.; Tang S.; Li M.; Zhang Y.; Xing W.; Xue Q.; Yu L. Sandwich Structure MoO2/MoS2/TiO2 Photocatalyst for Superb Hydrogen Evolution. J. Alloys Compd. 2020, 842, 155869. 10.1016/j.jallcom.2020.155869. [DOI] [Google Scholar]
  124. Feng H. Q.; Xi Y.; Xie H. Q.; Li Y. K.; Huang Q. Z. An Efficient Ternary Mn0.2Cd0.8S/MoS2/Co3O4 Heterojunction for Visible-Light-Driven Photocatalytic Hydrogen Evolution. Int. J. Hydrog. Energy. 2020, 45, 10764–10774. 10.1016/j.ijhydene.2020.02.030. [DOI] [Google Scholar]
  125. Liu Y.; Xu C.; Xie Y.; Yang L.; Ling Y.; Chen L. Au-Cu Nanoalloy/TiO2/MoS2 Ternary Hybrid with Enhanced Photocatalytic Hydrogen Production. J. Alloys Compd. 2020, 820, 153440. 10.1016/j.jallcom.2019.153440. [DOI] [Google Scholar]
  126. Chen R.; Wang P.; Chen J.; Wang C.; Ao Y. Synergistic Effect of MoS2 and MXene on the Enhanced H2 Evolution Performance of CdS Under Visible Light Irradiation. Appl. Surf. Sci. 2019, 473, 11–19. 10.1016/j.apsusc.2018.12.071. [DOI] [Google Scholar]
  127. Wang B.; Deng Z.; Fu X.; Li Z. MoS2/CQDs Obtained by Photoreduction for Assembly of a Ternary MoS2/CQDs/ZnIn2S4 Nanocomposite for Efficient Photocatalytic Hydrogen Evolution Under Visible Light. J. Mater. Chem. A 2018, 6, 19735–19742. 10.1039/C8TA07797E. [DOI] [Google Scholar]
  128. Guan Z.; Wang P.; Li Q.; Li G.; Yang J. Constructing a ZnIn2S4 Nanoparticle/MoS2-RGO Nanosheet 0D/2D Heterojunction for Significantly Enhanced Visible Light Photocatalytic H2 Production. Dalton Trans. 2018, 47, 6800–6807. 10.1039/C8DT00946E. [DOI] [PubMed] [Google Scholar]
  129. Du J.; Wang H.; Yang M.; Zhang F.; Wu H.; Cheng X.; Yuan S.; Zhang B.; Li K.; Wang Y.; Lee H. Highly Efficient Hydrogen Evolution Catalysis Based on MoS2/CdS/TiO2 Porous Composites. Int. J. Hydrog. Energy. 2018, 43, 9307–9315. 10.1016/j.ijhydene.2018.03.208. [DOI] [Google Scholar]
  130. Yuan Y.-J.; Fang G.; Chen D.; Huang Y.; Yang L.-X.; Cao D.-P.; Wang J.; Yu Z.-T.; Zou Z.-G. High Light Harvesting Efficiency CuInS2 quantum dots/TiO2/MoS2 Photocatalysts for Enhanced Visible Light Photocatalytic Hydrogen Production. Dalton Trans. 2018, 47, 5652–5659. 10.1039/C8DT00356D. [DOI] [PubMed] [Google Scholar]
  131. Zhao H.; Sun S.; Wu Y.; Jiang P.; Dong Y.; Xu Z. J. Ternary Graphitic Carbon Nitride/Red Phosphorous/Molybdenum Disulfide Heterostructure: An Efficient and Low-Cost Photocatalyst for Visible-Light -Driven H2 Evolution from Water. Carbon. 2017, 119, 56–61. 10.1016/j.carbon.2017.03.100. [DOI] [Google Scholar]
  132. Qin N.; Xiong J.; Liang R.; Liu Y.; Zhang S.; Li Y.; Li Z.; Wu L. Highly Efficient Photocatalytic H2 Evolution over MoS2/CdS-TiO2 Nanofibers Prepared by an Electrospinning Mediated Photoreduction Method. Appl. Catal., B 2017, 202, 374–380. 10.1016/j.apcatb.2016.09.040. [DOI] [Google Scholar]
  133. Jo W.-K.; Sagaya Selvam N. C. Fabrication of Photostable Ternary CdS/MoS2/MWCNTs Hybrid Photocatalysts with Enhanced H2 Generation Activity. Appl. Catal. A Gen. 2016, 525, 9–22. 10.1016/j.apcata.2016.06.036. [DOI] [Google Scholar]
  134. Jiang W.; Liu Y.; Zong R.; Li Z.; Yao W.; Zhu Y. Photocatalytic Hydrogen Generation on Bifunctional Ternary Heterostructured In2S3/MoS2/CdS Composites with High Activity and Stability Under Visible Light Irradiation. J. Mater. Chem. A 2015, 3, 18406–18412. 10.1039/C5TA04258E. [DOI] [Google Scholar]
  135. Du R.; Zhang Y.; Li B.; Yu X.; Liu H.; An X.; Qu J. Biomolecule-Assisted Synthesis of Defect-Mediated Cd1-xZnxS/MoS2/Graphene Hollow Spheres for Highly Efficient Hydrogen Evolution. Phys. Chem. Chem. Phys. 2016, 18, 16208–16215. 10.1039/C6CP01322H. [DOI] [PubMed] [Google Scholar]
  136. Yan W.; Xu Y.; Hao S.; He Z.; Wang L.; Wei Q.; Xu J.; Tang H. Promoting Charge Separation in Hollow-Structured C/MoS2@ZnIn2S4/Co3O4 Photocatalysts via Double Heterojunctions for Enhanced Photocatalytic Hydrogen Evolution. Inorg. Chem. 2022, 61, 4725–4734. 10.1021/acs.inorgchem.2c00045. [DOI] [PubMed] [Google Scholar]
  137. Xu Y.; Yan A.; Jiang L.; Huang F.; Hu D.; Duan G.; Zheng F. MoS2/HCSs/ZnIn2S4 Nanocomposites with Enhanced Charge Transport and Photocatalytic Hydrogen Evolution Performance. J. Alloys Compd. 2022, 895, 162504. 10.1016/j.jallcom.2021.162504. [DOI] [Google Scholar]
  138. Ran Q.; Yu Z.; Jiang R.; Qian L.; Hou Y.; Yang F.; Li F.; Li M.; Sun Q.; Zhang H. Path Electron Transfer Created in S-Doped NH2-UiO-66 Bridged ZnIn2S4/MoS2 Nanosheet Heterostructure for Boosting Photocatalytic Hydrogen Evolution. Catal. Sci. Technol. 2020, 10, 2531–2539. 10.1039/D0CY00127A. [DOI] [Google Scholar]
  139. Lin H.; Sun B.; Wang H.; Ruan Q.; Geng Y.; Li Y.; Wu J.; Wang W.; Liu J.; Wang X. Unique 1D Cd1-xZnxS@O-MoS2/NiOx Nanohybrids: Highly Efficient Visible-Light-Driven Photocatalytic Hydrogen Evolution via Integrated Structural Regulation. Small. 2019, 15, 1804115. 10.1002/smll.201804115. [DOI] [PubMed] [Google Scholar]
  140. Liu X.; Wang B.; Heng Q.; Chen W.; Li X.; Mao L.; Shangguan W. Promoted Charge Separation on 3D Interconnected Ti3C2/MoS2/CdS Composite for Enhanced Photocatalytic H2 Production. Int. J. Hydrog. Energy. 2022, 47, 8284–8293. 10.1016/j.ijhydene.2021.12.180. [DOI] [Google Scholar]
  141. Huang X.; El-Sayed M. A. Gold Nanoparticles: Optical Properties and Implementations in Cancer Diagnosis and Photothermal Therapy. J. Adv. Res. 2010, 1, 13–28. 10.1016/j.jare.2010.02.002. [DOI] [Google Scholar]
  142. Pany S.; Naik B.; Martha S.; Parida K. Plasmon Induced Nano Au Particle Decorated over S, N-Modified TiO2 Exceptional Photocatalytic Hydrogen Evolution Under Visible Light. ACS Appl. Mater. Interfaces. 2014, 6, 839–846. 10.1021/am403865r. [DOI] [PubMed] [Google Scholar]
  143. Zhang J.; Liu M.; Wang Y.; Shi F. Au/MoS2/Ti3C2 Composite Catalyst for Efficient Photocatalytic Hydrogen Evolution. CrystEngComm. 2020, 22, 3683–3691. 10.1039/D0CE00345J. [DOI] [Google Scholar]
  144. Li S.; Pu T.; Wang J.; Fang X.; Liu Y.; Kang S.; Cui L. Efficient Visible-Light-Driven Hydrogen Evolution over Ternary MoS2/Pt-TiO2 Photocatalyst with Low Overpotential. Int. J. Hydrog. Energy. 2018, 43, 16534–16542. 10.1016/j.ijhydene.2018.07.050. [DOI] [Google Scholar]
  145. Wi D. H.; Park S. Y.; Lee S.; Sung J.; Hong J. W.; Han S. W. Metal-Semiconductor Ternary Hybrids for Efficient Visible-Light Photocatalytic Hydrogen Evolution. J. Mater. Chem. A 2018, 6, 13225–13235. 10.1039/C8TA03462A. [DOI] [Google Scholar]
  146. Huang T.; Hua Y.; Yao J.; Luo Y.; Liu X. The Synergetic Effect of Graphene and MoS2 on AgInZnS for Visible-Light-Driven Photocatalytic H2 Evolution. Mater. Chem. Phys. 2018, 212, 506–512. 10.1016/j.matchemphys.2018.03.068. [DOI] [Google Scholar]
  147. Chava R. K.; Do J. Y.; Kang M. Smart Hybridization of Au Coupled CdS Nanorods with Few Layered MoS2 Nanosheets for High-Performance Photocatalytic Hydrogen Evolution Reaction. ACS Sustain. Chem. Eng. 2018, 6, 6445–6457. 10.1021/acssuschemeng.8b00249. [DOI] [Google Scholar]
  148. Guo S.; Li X.; Zhu J.; Tong T.; Wei B. Au NPs@MoS2 Sub-Micrometer Sphere-ZnO Nanorod Hybrid Structures for Efficient Photocatalytic Hydrogen Evolution with Excellent Stability. Small. 2016, 12, 5692–5701. 10.1002/smll.201602122. [DOI] [PubMed] [Google Scholar]
  149. Cowan A. J.; Durrant J. R. Long-Lived Charge Separated States in Nanostructured Semiconductor Photoelectrodes of the Production of Solar Fuels. Chem. Soc. Rev. 2013, 42, 2281–2293. 10.1039/C2CS35305A. [DOI] [PubMed] [Google Scholar]
  150. Izumi Y. Recent Advances in the Photocatalytic Conversion of Carbon Dioxide to Fuels with Water and/or Hydrogen Using Solar Energy and Beyond. Coord. Chem. Rev. 2013, 257, 171–186. 10.1016/j.ccr.2012.04.018. [DOI] [Google Scholar]
  151. Yu J.; Low J.; Xiao W.; Zhou P.; Jaroniec M. Enhanced Photocatalytic CO2-Reduction Activity of Anatase TiO2 by Co-exposed {001} and {101} Facets. J. Am. Chem. Soc. 2014, 136, 8839–8842. 10.1021/ja5044787. [DOI] [PubMed] [Google Scholar]
  152. Hoffmann M. R.; Moss J. A.; Baum M. M. Artificial Photosynthesis: Semiconductor Photocatalytic Fixation of CO2 to Afford Higher Organic Compounds. Dalton Trans. 2011, 40, 5151–5158. 10.1039/c0dt01777a. [DOI] [PubMed] [Google Scholar]
  153. Sun H.; Wang S. Research Advances in the Synthesis of Nanocarbon-Based Photocatalysts and their Applications for Photocatalytic Conversion of Carbon Dioxide to Hydrocarbon Fuels. Energy Fuels. 2014, 28, 22–36. 10.1021/ef401426x. [DOI] [Google Scholar]
  154. Li X.; Yu J.; Jaroniec M.; Chen X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. 10.1021/acs.chemrev.8b00400. [DOI] [PubMed] [Google Scholar]
  155. Xiong Z.; Lei Z.; Li Y.; Dong L.; Zhao Y.; Zhang J. A Review on Modification of Facet-Engineered TiO2 for Photocatalytic CO2 Reduction. J. Photochem. Photobiol. C 2018, 36, 24–47. 10.1016/j.jphotochemrev.2018.07.002. [DOI] [Google Scholar]
  156. Inoue T.; Fujishima A.; Konishi S.; Honda K. Photocatalytic Reduction of Carbon Dioxide in Aqueous Suspensions of Semiconductor Powders. Nature. 1979, 277, 637–638. 10.1038/277637a0. [DOI] [Google Scholar]
  157. Singh S.; Modak A.; Pant K. K. CO2 Reduction to Methanol Using a Conjugated Organic-Inorganic Hybrid TiO2-C3N4 Nano-assembly. Transactions of the Indian National Academy of Engineering. 2021, 6, 395–404. 10.1007/s41403-021-00201-6. [DOI] [Google Scholar]
  158. Aguirre M. E.; Zhou R.; Eugene A. J.; Guzman M. I.; Grela M. A. Cu2O/TiO2 Heterostructures for CO2 Reduction Through a Direct Z-scheme: Protecting Cu2O from Photocorrosion. Appl. Catal. B: Environ. 2017, 217, 485–493. 10.1016/j.apcatb.2017.05.058. [DOI] [Google Scholar]
  159. Wei Z. H.; Wang Y. F.; Li Y. Y.; Zhang L.; Yao H. C.; Li Z. J. Enhanced Photocatalytic CO2 Reduction Activity of Z-scheme CdS/BiVO4 Nanocomposite with Thinner BiVO4 Nanosheets. J. CO2 Util. 2018, 28, 15–25. 10.1016/j.jcou.2018.09.008. [DOI] [Google Scholar]
  160. Wang S.; Teramura K.; Hisatomi T.; Domen K.; Asakura H.; Hosokawa S.; Tanaka T. Optimized Synthesis of Ag-Modified Al-Doped SrTiO3 Photocatalyst for the Conversion of CO2 Using H2O as an Electron Donor. ChemistrySelect. 2020, 5, 8779–8786. 10.1002/slct.202001693. [DOI] [Google Scholar]
  161. Miao Y. F.; Guo R. T.; Gu J. W.; Liu Y. Z.; Wu G. I.; Duan C. P.; Zhang X. D.; Pan W. G. Fabrication of In2S3/NiAl-LDH Heterojunction Photocatalyst with Enhanced Separation of Charge Carriers for Efficient CO2 Photocatalytic Reduction. Appl. Surf. Sci. 2020, 527, 146792. 10.1016/j.apsusc.2020.146792. [DOI] [Google Scholar]
  162. Fu J.; Zhu B.; Jiang C.; Cheng B.; You W.; Yu J. Hierarchical Porous O-doped g-C3N4 with Enhanced Photocatalytic CO2 Reduction Activity. Small. 2017, 13, 1603938. 10.1002/smll.201603938. [DOI] [PubMed] [Google Scholar]
  163. Gusain R.; Kumar P.; Sharma O. P.; Jain S. L.; Khatri O. P. Reduced Graphene Oxide-CuO Nanocomposites for Photocatalytic Conversion of CO2 into Methanol Under Visible Light Irradiation. Appl. Catal. B: Environ. 2016, 181, 352–362. 10.1016/j.apcatb.2015.08.012. [DOI] [Google Scholar]
  164. Li Z.; Meng X.; Zhang Z. Recent Development on MoS2-Based Photocatalysis: A review. J. Photochem. Photobiol. C: Photochem. Rev. 2018, 35, 39–55. 10.1016/j.jphotochemrev.2017.12.002. [DOI] [Google Scholar]
  165. Lv R.; Terrones H.; Elias A. L.; Perea-Lopez N.; Gutierrez H. R.; Cruz-Silva E.; Rajukumar L. P.; Dresselhaus M. S.; Terrones M. Two-Dimensional Transition Metal Dichalcogenides: Clusters, Ribbons, Sheets and More. Nano Today. 2015, 10, 559–592. 10.1016/j.nantod.2015.07.004. [DOI] [Google Scholar]
  166. Zhou W.; Zou X.; Najmaei S.; Liu Z.; Shi Y.; Kong J.; Lou J.; Ajayan P. M.; Yakobson B. I.; Idrobo J.-C. Intrinsic Structural Defects in Monolayer Molybdenum Disulfide. Nano letters. 2013, 13, 2615–2622. 10.1021/nl4007479. [DOI] [PubMed] [Google Scholar]
  167. Liu P.; Choi Y.; Yang Y.; White M. G. Methanol Synthesis from H2 and CO2 on a Mo6S8 Cluster: A Density Functional Study. J. Phys. Chem. A 2010, 114, 3888–3895. 10.1021/jp906780a. [DOI] [PubMed] [Google Scholar]
  168. Santos V. P.; van der Linden B.; Chojecki A.; Budroni G.; Corthals S.; Shibata H.; Meima G. R.; Kapteijn F.; Makkee M.; Gascon J. Mechanistic Insight into the Synthesis of Higher Alcohols from Syngas: The Role of K Promotion on MoS2 Catalysts. ACS Catal. 2013, 3, 1634–1637. 10.1021/cs4003518. [DOI] [Google Scholar]
  169. Asadi M.; Kumar B.; Behranginia A.; Rosen B. A.; Baskin A.; Repnin N.; Pisasale D.; Phillips P.; Zhu W.; Haasch R.; Klie R. F.; Kral P.; Abiade J.; Salehi-Khojin A. Robust Carbon Dioxide Reduction on Molybdenum Disulfide Edges. Nat. Commun. 2014, 5, 1–8. 10.1038/ncomms5470. [DOI] [PubMed] [Google Scholar]
  170. Xie Y.; Li X.; Wang Y.; Li B.; Yang L.; Zhao N.; Liu M.; Wang X.; Yu Y.; Liu J.-M. Reaction Mechanisms for Reduction of CO2 to CO on Monolayer MoS2. Appl. Surf. Sci. 2020, 499, 143964. 10.1016/j.apsusc.2019.143964. [DOI] [Google Scholar]
  171. Wu J.; Huang Y.; Ye W.; Li Y. CO2 Reduction: from the Electrochemical to Photochemical Approach. Adv. Sci. 2017, 4, 1700194. 10.1002/advs.201700194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Meier A. J.; Garg A.; Sutter B.; Kuhn J. N.; Bhethanabotla V. R. MoS2 Nanoflowers as a Gateway for Solar-Driven CO2 Photoreduction. ACS Sustain. Chem. Eng. 2019, 7, 265–275. 10.1021/acssuschemeng.8b03168. [DOI] [Google Scholar]
  173. Tu W.; Li Y.; Kuai L.; Zhou Y.; Xu Q.; Li H.; Wang X.; Xiao M.; Zou Z. Construction of Unique Two-Dimensional MoS2-TiO2 Hybrid Nanojunctions: MoS2 as a Promising Cost-effective Co-catalyst Toward Improved Photocatalytic Reduction of CO2 to Methanol. Nanoscale 2017, 9, 9065–9070. 10.1039/C7NR03238B. [DOI] [PubMed] [Google Scholar]
  174. Jung H.; Cho K. M.; Kim K. H.; Yoo H. W.; Al-Saggaf A.; Gereige I.; Jung H. T. Highly Efficient Stable CO2 Reduction Photocatalyst with a Hierarchical Structure of Mesoporous TiO2 on 3D Graphene with Few-Layered MoS2. ACS Sustain. Chem. Eng. 2018, 6, 5718–5724. 10.1021/acssuschemeng.8b00002. [DOI] [Google Scholar]
  175. Jia P.-y.; Guo R.-t.; Pan W.-g.; Huang C.-y.; Tang J.-y.; Liu X.-y.; Qin H.; Xu Q.-y. The MoS2/TiO2 Heterojunction Composites with Enhanced Activity for CO2 Photocatalytic Reduction Under Visible Light Irradiation. Colloids Surf, A Physicochem Eng. Asp. 2019, 570, 306–316. 10.1016/j.colsurfa.2019.03.045. [DOI] [Google Scholar]
  176. Xu F.; Zhu B.; Cheng B.; Yu J.; Xu J. 1D/2D TiO2/MoS2 Hybrid Nanostructures for Enhanced Photocatalytic CO2 Reduction. Adv. Opt. Mater. 2018, 6, 1800911. 10.1002/adom.201800911. [DOI] [Google Scholar]
  177. Zhao Y.; Cai W.; Shi Y.; Tang J.; Gong Y.; Chen M.; Zhong Q. Construction of Nano-Fe2O3-Decorated Flower-Like MoS2 with Fe-S Bonds for Efficient Photoreduction of CO2 under Visible-Light Irradiation. ACS Sustain. Chem. Eng. 2020, 8, 12603–12611. 10.1021/acssuschemeng.0c04042. [DOI] [Google Scholar]
  178. Dai W.; Yu J.; Deng Y.; Hu X.; Wang T.; Luo X. Facile Synthesis of MoS2/Bi2WO6 Nanocomposites for Enhanced CO2 Photoreduction Activity Under Visible Light Irradiation. Appl. Surf. Sci. 2017, 403, 230–239. 10.1016/j.apsusc.2017.01.171. [DOI] [Google Scholar]
  179. Khan B.; Raziq F.; Bilal Faheem M.; Umar Farooq M.; Hussain S.; Ali F.; Ullah A.; Mavlonov A.; Zhao Y.; Liu Z.; Tian H.; Shen H.; Zu X.; Li S.; Xiao H.; Xiang X.; Qiao L. Electronic and Nanostructure Engineering of Bifunctional MoS2 Towards Exceptional Visible-light Photocatalytic CO2 Reduction and Pollutant Degradation. J. Hazard. Mater. 2020, 381, 120972. 10.1016/j.jhazmat.2019.120972. [DOI] [PubMed] [Google Scholar]
  180. Kim R.; Kim J.; Do J. Y.; Seo M. W.; Kang M. Carbon Dioxide Photoreduction on the Bi2S3/MoS2 Catalysts. Catalysts. 2019, 9, 998. 10.3390/catal9120998. [DOI] [Google Scholar]
  181. Qin H.; Guo R.-T.; Liu X.-Y.; Pan W.-G.; Wang Z.-Y.; Shi X.; Tang J.-Y.; Huang C.-Y. Z-Scheme MoS2/g-C3N4 Heterojunction for Efficient Visible Light Photocatalytic CO2 Reduction. Dalton Trans. 2018, 47, 15155–15163. 10.1039/C8DT02901F. [DOI] [PubMed] [Google Scholar]
  182. Sun S.; An Q.; Watanabe M.; Cheng J.; Ho Kim H.; Akbay T.; Takagaki A.; Ishihara T. Highly Correlation of CO2 Reduction Selectivity and Surface Electron Accumulation: A Case Study of Au-MoS2 and Ag-MoS2 Catalyst. Appl. Catal., B 2020, 271, 118931. 10.1016/j.apcatb.2020.118931. [DOI] [Google Scholar]
  183. Ouyang T.; Fan W.; Guo J.; Zheng Y.; Yin X.; Shen Y. DFT Study on Ag Loaded 2H-MoS2 for Understanding the Mechanism of Improved Photocatalytic Reduction of CO2. Phys. Chem. Chem. Phys. 2020, 22, 10305–10313. 10.1039/D0CP01485K. [DOI] [PubMed] [Google Scholar]
  184. Zheng Y.; Yin X.; Jiang Y.; Bai J.; Tang Y.; Shen Y.; Zhang M. Nano Ag-Decorated MoS2 Nanosheets from 1T to 2H Phase Conversion for Photocatalytically Reducing CO2 to Methanol. Energy Technol. 2019, 7, 1900582. 10.1002/ente.201900582. [DOI] [Google Scholar]
  185. Lu J.; Zhang Z.; Cheng L.; Liu H. MoS2-Wrapped Mn0.2Cd0.8S Nanospheres Towards Efficient Photocatalytic H2 Generation and CO2 Reduction. New J. Chem. 2020, 44, 13728–13737. 10.1039/D0NJ02174A. [DOI] [Google Scholar]
  186. Wang X.; He J.; Mao L.; Cai X.; Sun C.; Zhu M. CsPbBr3 Perovskite Nanocrystals Anchoring on Monolayer MoS2 Nanosheets for Efficient Photocatalytic CO2 Reduction. Chem. Eng. J. 2021, 416, 128077. 10.1016/j.cej.2020.128077. [DOI] [Google Scholar]
  187. Kumar N.; Kumar S.; Gusain R.; Manyala N.; Eslava S.; Ray S. S. Polypyrrole-Promoted rGO-MoS2 Nanocomposites for Enhanced Photocatalytic Conversion of CO2 and H2O to CO, CH4, and H2 Products. ACS Appl. Energy Mater. 2020, 3, 9897–9909. 10.1021/acsaem.0c01602. [DOI] [Google Scholar]
  188. Yin S.; Li J.; Sun L.; Li X.; Shen D.; Song X.; Huo P.; Wang H.; Yan Y. Construction of Heterogenous S-C-S MoS2/SnS2/r-GO Heterojunction for Efficient CO2 Photoreduction. Inorg. Chem. 2019, 58, 15590–15601. 10.1021/acs.inorgchem.9b02676. [DOI] [PubMed] [Google Scholar]
  189. Wang Y.; Zhang Z.; Zhang L.; Luo Z.; Shen J.; Lin H.; Long J.; Wu J. C. S.; Fu X.; Wang X.; Li C. Visible-Light Driven Overall Conversion of CO2 and H2O to CH4 and O2 on 3D SiC@2D MoS2 Heterostructure. J. Am. Chem. Soc. 2018, 140, 14595–14598. 10.1021/jacs.8b09344. [DOI] [PubMed] [Google Scholar]
  190. Qiu C.; Hao X.; Tan L.; Wang X.; Cao W.; Liu J.; Zhao Y.; Song Y. F. 500 nm Induced Tunable Syngas Synthesis from CO2 Photoreduction by Controlling Heterojunction Concentration. ChemComm. 2020, 56, 5354–5357. 10.1039/D0CC00971G. [DOI] [PubMed] [Google Scholar]
  191. Alkanad K.; Hezam A.; Drmosh Q. A.; Ganganakatte Chandrashekar S. S.; AlObaid A. A.; Warad I.; Bajiri M. A.; Neratur Krishnappagowda L. Construction of Bi2S3/TiO2/MoS2 S-Scheme Heterostructure with a Switchable Charge Migration Pathway for Selective CO2 Reduction. Sol. RRL. 2021, 5, 2100501. 10.1002/solr.202100501. [DOI] [Google Scholar]
  192. Geioushy R. A.; Hegazy I. M.; El-Sheikh S. M.; Fouad O. A. Construction of 2D MoS2@ZnO Heterojunction as Superior Photocatalyst for Highly Efficient and Selective CO2 Conversion into Liquid Fuel. J. Environ. Chem. Eng. 2022, 10, 107337. 10.1016/j.jece.2022.107337. [DOI] [Google Scholar]
  193. Wang F.; Liu D.; Wen J.; Zheng X. In-situ Sulfurized In2S3/MoO3@MoS2 Heterojunction for Visible Light-induced CO2 Photoreduction. J. Environ. Chem. Eng. 2021, 9, 106042. 10.1016/j.jece.2021.106042. [DOI] [Google Scholar]
  194. Zheng X.; Wang H.; Wen J.; Peng H. In2S3-NiS Co-decorated MoO3@MoS2 Composites for Enhancing the Solar-light Induced CO2 Photoreduction Activity. Int. J. Hydrog. Energy. 2021, 46, 36848–36858. 10.1016/j.ijhydene.2021.08.161. [DOI] [Google Scholar]
  195. Yu F.; Jing X.; Wang Y.; Sun M.; Duan C. Hierarchically Porous Metal-Organic Framework/MoS2 Interface for Selective Photocatalytic Conversion of CO2 with H2O into CH3COOH. Angew. Chem. 2021, 133, 25053–25057. 10.1002/ange.202108892. [DOI] [PubMed] [Google Scholar]
  196. Li H.; Sun T.; Zhang L.; Cao Y. Improved Photocatalytic Activity of ZnO via the Modification of In2O3 and MoS2 Surface Species for the Photoreduction of CO2. Appl. Surf. Sci. 2021, 566, 150649. 10.1016/j.apsusc.2021.150649. [DOI] [Google Scholar]
  197. Singh S.; Punia R.; Pant K. K.; Biswas P. Effect of Work-function and Morphology of Heterostructure Components on CO2 Reduction Photo-catalytic Activity of MoS2-Cu2O Heterostructure. Chem. Eng. J. 2022, 433, 132709. 10.1016/j.cej.2021.132709. [DOI] [Google Scholar]
  198. Li P.; Hu H.; Xu J.; Jing H.; Peng H.; Lu J.; Wu C.; Ai S. New Insights into the Photo-Enhanced Electrocatalytic Reduction of Carbon Dioxide on MoS2-Rods/TiO2 NTs with Unmatched Energy Bands. Appl. Catal., B 2014, 147, 912–919. 10.1016/j.apcatb.2013.10.010. [DOI] [Google Scholar]
  199. Kim M. S.; Don W. J.; Hong S. I.; Ri M. I.; Yang S. I. Photocatalytic Property of Two Dimensional Heterostructure MoS2/WS2 for Hydrogen Evolution via Water Splitting; a First Principles Calculation. Int. J. Hydrog. Energy. 2023, 48, 9371–9376. 10.1016/j.ijhydene.2022.12.027. [DOI] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES