Next Article in Journal
N-Dealkylation of Amines
Next Article in Special Issue
MXenes as Emerging Materials: Synthesis, Properties, and Applications
Previous Article in Journal
A Comprehensive Review of Rosmarinic Acid: From Phytochemistry to Pharmacology and Its New Insight
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review

School of Environmental & Chemical Engineering, Jiangsu University of Science and Technology, Zhenjiang 212100, China
*
Authors to whom correspondence should be addressed.
Submission received: 4 April 2022 / Revised: 9 May 2022 / Accepted: 17 May 2022 / Published: 20 May 2022
(This article belongs to the Special Issue Recent Progress in the Synthesis and Applications of Photocatalysts)

Abstract

:
Molybdenum disulfide (MoS2), with a two-dimensional (2D) structure, has attracted huge research interest due to its unique electrical, optical, and physicochemical properties. MoS2 has been used as a co-catalyst for the synthesis of novel heterojunction composites with enhanced photocatalytic hydrogen production under solar light irradiation. In this review, we briefly highlight the atomic-scale structure of MoS2 nanosheets. The top-down and bottom-up synthetic methods of MoS2 nanosheets are described. Additionally, we discuss the formation of MoS2 heterostructures with titanium dioxide (TiO2), graphitic carbon nitride (g-C3N4), and other semiconductors and co-catalysts for enhanced photocatalytic hydrogen generation. This review addresses the challenges and future perspectives for enhancing solar hydrogen production performance in heterojunction materials using MoS2 as a co-catalyst.

1. Introduction

Hydrogen is a clean, renewable energy source and alternative to fossil fuels [1] that can be stored at high mass-specific energy density, and its only product on combustion is water [2]. At present, about 96% of hydrogen is industrially produced from coal gasification and steam methane reformation processes [1,2]. However, these processes of hydrogen production also generate secondary pollutants or greenhouse gases, such as CO2 and N2O, that affect the environment [2]. Methane pyrolysis produces hydrogen and solid carbon as a byproduct [3]. This process generates CO2-free hydrogen and has an advantage over conventional steam methane reformation and coal gasification processes. However, methane pyrolysis is a temporary solution and not a sustainable process due to the depletion of natural gas reserves [3].
To overcome energy challenges and environmental problems, hydrogen production from electrochemical water splitting using highly active catalysts is a promising strategy [4,5]. Less than 4% of hydrogen is produced through electrocatalysis at the industrial level [2]. The electrocatalysis of water for hydrogen production is a high-cost technique, which has hindered its large-scale industrialization. As an alternative, photocatalytic hydrogen evolution reaction (HER) from water splitting over a particular semiconductor material has been the most interesting way to address these issues. Generally, the photocatalytic efficiency depends upon three processes, including light absorption in the solar spectrum, charge separation, and surface active sites for catalytic activity [1,2,6].
A photocatalyst that can absorb sunlight across the whole solar spectrum is considered to be an ideal candidate for photocatalysis [6,7]. In 1972, Fujishima et al. reported photo-induced water splitting on TiO2 electrodes [8]. Since then, much research has been focused on TiO2 and other related semiconducting materials such as metal oxides, metal sulfides, conjugated polymers, nanosheets, graphitic carbon nitride, metal organic frameworks, and covalent organic frameworks, etc., as photocatalysts for hydrogen production [9,10,11,12,13,14,15,16,17]. However, the available photocatalysts for hydrogen production are still limited due to low visible light absorption and high electron–hole recombination rates.
Molybdenum disulfide (MoS2), with a 2D nanostructure, has attracted huge attention due to its outstanding optical and electronic properties and promising applications [18,19,20,21,22]. MoS2 nanomaterials as co-catalysts are promising photocatalysts for HER [18,23]. It is reported that the exposed edges of layers of MoS2 contain active sites for catalytic activity while its basal planes are mostly inactive [19,24]. In addition to the active photocatalytic sites, the band gap of MoS2 nanosheets is an important parameter for photocatalytic HER. The band gap of MoS2 increases from bulk (1.2 eV) to single layer (1.9 eV) due to quantum confinement [19,24]. As a result, the location of the conduction band (CB) of MoS2 moves towards a more negative potential than the proton reduction potential (H+/H2), which consequently enhances the reduction in adsorbed H+ and photocatalytic hydrogen evolution.
It is widely reported that loading a co-catalyst over semiconductors is a promising approach with superior photocatalytic performance due to the photoelectron separation and charge transfer [18,19]. MoS2-decorated semiconductor materials constitute a promising approach that has shown superior hydrogen production due to their heterojunctions with controllable nanoscale architectures, design for enhanced performance in terms of light absorption, charge separation, and surface catalytic reactions [15,19,23,24].
In this review, we briefly introduce the basic aspects and synthetic methods of MoS2 nanosheets. Different types of MoS2-based heterojunction composites are also discussed. The role of MoS2 nanomaterials as co-catalysts in heterojunction composites for enhanced HER performance is addressed. Additionally, some important issues are highlighted and useful opinions are discussed to further improve photocatalytic hydrogen production using MoS2 as a co-catalyst.

2. Atomic-Scale Structure of MoS2

A single layer of MoS2 has a sandwich structure of S-Mo-S, where the Mo atoms are covalently bonded with the S atoms (Figure 1). MoS2 has several polymorphs, including 1T1, 1T2, 1H, 2T, 2H, 2R1, 2R2, 3Ha, 3Hb, 3R, and 4T [25,26,27,28,29]. Among them the 1T MoS2, 2H MoS2, and 3R-MoS2 polymorphs of MoS2 have been most investigated for different applications [25,27,28,29]. A single-layer 1T MoS2 sheet is metallic and has good electrical properties [30,31], while single-layer 2H MoS2 and 3R-MoS2 sheets behave as a semiconductor with a direct band gap [28,32].
Generally, MoS2 sheets are stacked together by weak van der Waals forces and form few-layer MoS2 nanosheets. As the band gap of MoS2 nanosheets increases from bulk (1.2 eV) to single layer (1.9 eV) [33], it absorbs the visible region of the solar spectrum. Thus, MoS2 can play an important role as a co-catalyst during photocatalysis [29]. MoS2-based semiconductor composites act as co-catalysts that can significantly enhance the efficiency of photocatalytic hydrogen production [7,34,35,36,37].

3. Photochemical Hydrogen Evolution Reaction

As mentioned above, Fujishima and Honda reported on photo-induced water splitting on TiO2 electrodes. Hydrogen can also be directly produced from photochemical water splitting. Usually, a photoelectrolytic cell is designed to carry out the photochemical water splitting process. A typical photoelectrolytic cell for water splitting is shown in Figure 2a [38]. Using light sources, the photocatalytic water splitting takes place in several steps: the absorption of light by catalyst on electrode; the generation of charges followed by the excitation of electrons in the valence band; the separation of charge as well as the transport of charge carriers; and the oxidation of water and generation of hydrogen during water splitting, which occur at separate electrodes. The pure, overall water splitting process comprises two half-reactions to generate hydrogen and oxygen molecules, as shown in Figure 2b [39]. Water oxidation occurs at the anode to produce oxygen, whereas H+ ions are reduced on the cathode into hydrogen gas. For more details of photocatalytic water splitting, see the review of Jeong et al. [39].

4. Synthesis of MoS2

Nanostructured MoS2 can be fabricated via both top-down and bottom-up approaches. In the case of the top-down method, the commercially available bulk crystal of MoS2 is physically downsized into MoS2 nanomaterials (Figure 3) [29,40,41], while in the bottom-up approach, MoS2 nanomaterials are synthesized via chemical reaction with small molecules using chemical vapor deposition (CVD) and hydrothermal or solvothermal methods, etc. [42,43,44]. Single layers, multilayers, nanoparticles, and quantum dots of MoS2 have also been reported [45,46,47,48]. Continued efforts have been reported for the fabrication of MoS2 nanomaterials via the top-down and bottom-up strategies [16,17,18,19,28,29,30,31,40,41,42,43,44].

4.1. Top-Down Approach

Exfoliation of MoS2

Due to the layered structure and van der Waals interactions, MoS2 nanosheets can be easily prepared through the exfoliation method. Mechanical, chemical, electrochemical, and liquid-phase exfoliation processes have been reported for the synthesis of MoS2 nanosheets [39,40,41,42,43,44,45,46,47,48,49,50,51,52]. For example, in the mechanical exfoliation technique, the suitable MoS2 flakes are peeled off from the bulk crystal of MoS2 using adhesive tape and shifted onto a specific substrate [46,53]. When the scotch tape is detached, some parts of MoS2 remain on the substrate. As result, single- or few-layer MoS2 nanosheets with random shapes and sizes are obtained. The 2D materials prepared by the exfoliation method have good quality and allow to study the pristine properties of materials and device performance. However, during this process, the thickness and size of the MoS2 are difficult to control, and the resulting materials are inappropriate for large-scale production and scaled-up applications [53,54]. Li et al., mechanically exfoliated single- and multilayer MoS2 nanosheets from SiO2/Si with the adhesive tape method [41]. The flakes of MoS2 were mechanically stripped on Si/SiO2 substrate. The obtained single-layer and multilayer MoS2 materials were characterized using a bright-field optical microscope and an atomic force microscope (AFM). From the AFM measurements, the height of a single MoS2 sheet was found to be 0.8 nm, while the thickness of two, three, and four layers of MoS2 nanosheets was 1.5, 2.1, and 2.9 nm, respectively (Figure 4). The MoS2 nanosheet monolayers showed an enhanced optical performance, especially single-layer MoS2 nanosheets. It was observed that the van der Waals interactions between MoS2 to SiO2 were much weaker. For this purpose, gold can be used as a substrate to exfoliate the MoS2 nanosheets due to its strong affinity for sulfur. It can exfoliate the MoS2 monolayer from the bulk because of the strong van der Waals interactions between Au and MoS2 layers [55,56,57]. Huang et al. prepared large-area MoS2 nanosheets using a Au-assisted exfoliation strategy [50]. In a typical synthesis, a Au thin layer was deposited on a Ti or Cr adhesion-covered substrate. To develop good contact between a MoS2 bulk crystal on tape and a Au-covered substrate, it should be passed under high pressure. The monolayer sheets with a large area were collected from the surface of the Au after peeling off the tape.
In the top-down approaches, single- and multilayer MoS2 nanosheets are prepared, which have been used to study some fundamental properties of MoS2 nanosheets.

4.2. Bottom-Down Approach

4.2.1. Chemical Vapor Deposition

The CVD technique has a long history and is commonly used for the synthesis of high-quality semiconductor materials. In a typical CVD process of MoS2 nanosheets, the Mo sources are solid precursors of Mo or MoO3 powder, and the S sources are H2S gas or solid S powder [58,59,60,61]. The solid MoO3 and vaporized S react with each other in a low-pressure chamber, forming nuclei for the growth of MoS2 [58]. Then, MoS2 slowly grows and enlarges its size on the substrates under carrier gas flow. The temperatures at which MoS2 grows during the CVD process are usually between 700 and 1000 °C, with a metal catalyst such as Au [61]. Plasma-enhanced CVD requires a low temperature (150–300 °C) for the growth of MoS2 nanosheets, and MoS2 can even be directly deposited on the plastic substrate [62]. Recently, metal organic CVD has been reported for the synthesis of MoS2 nanosheets [63,64], where organometallic precursors were used as starting materials.

4.2.2. Physical Vapor Deposition

Advanced technology such as molecular beam epitaxy (MBE) can be used to prepare single-crystal semiconductor thin films. However, its applications are limited to the synthesis of 2D materials [65]. Ordinary physical vapor deposition is rarely reported for 2D materials. A MoS2–Ti composite was prepared by direct current magnetron sputtering, using Ti and MoS2 materials [66]. In this process, the MoS2 was amorphous.

4.2.3. Solution-Based Process

Solution-based processes are commonly used to synthesize MoS2 nanosheets. Hydrothermal and solvothermal methods are the most interesting for the preparation of MoS2 nanosheets [67,68]. In these methods, the Mo source is commonly a molybdate, such as Na2MoO4 or (NH4)6Mo7O24, and the S source is thiourea and thioacetamide and L-cysteine [69,70,71,72,73]. The molybdate reacts with the S or S compound in a stainless steel autoclave. The physicochemical reaction takes place at high temperatures (160–200 °C) and pressure for at least a few hours. In the solvothermal method, organic solvents such as 1-methyl-2-pyrrolidinone, N,N-dimethylformamide, and polyethylene glycol-600 are used to proceed with the reaction, while in the hydrothermal method, water is used as a solvent. The MoS2 powders obtained from these methods have different sizes and shapes. The sizes and shapes of the products can be adjusted by altering the experimental conditions. To improve the crystalline quality of MoS2, the products are usually post-annealed at high temperature.
The MoS2 nanomaterials prepared through different bottom-up approaches have various sizes, shapes, morphologies, and thicknesses and can be used for many applications.

5. Application of MoS2 as a Co-Catalyst in Photocatalysis for Hydrogen Production

5.1. MoS2/Titanium Dioxide Composites

The semiconducting material titanium dioxide (TiO2) has been employed for hydrogen production due to its good UV light response, non-toxic nature, low cost, chemical stability, and good availability [1,12,24]. However, the photocatalytic energy conversion efficiency of TiO2 for hydrogen production is low due to its wide band gap structure (Eg ≈ 3.2 eV), photogenerated charge recombination, and some reverse reactions [1,2,12,24]. Many strategies have been attempted to improve the catalytic activity of TiO2 nanomaterials, including micro/nanostructure constructing, crystal facet, crystal phase, surface, and tailoring the band gap [9,74,75], but the photocatalytic activity of TiO2 still cannot reach the expected efficiency.
MoS2 is considered a potential co-catalyst for TiO2 materials to boost the efficiency of photocatalytic hydrogen production. Zhu and his coworkers fabricated MoS2/TiO2 photocatalysts with various compositions through a facial mechanochemistry method [76]. The photocatalytic activity of the prepared composite was studied for hydrogen generation under UV irradiation. The 4% MoS2 loaded on TiO2 (4%-MoS2/TiO2) showed maximum hydrogen production at a rate of 150.7 μmol h−1, which is about 48.6 times higher than that of pure TiO2 at ~3.1 μmol·h−1. The improved photocatalytic activity of MoS2/TiO2 composites is mainly due to electron transfer from TiO2 to MoS2 nanosheets and the active sites that produce hydrogen. Meanwhile, the recombination rate of electron–hole pairs is also reduced. Furthermore, the relatively good conductivity of MoS2 nanosheets also assisted the photo-induced charge separation, leading to an enhanced photocatalytic performance. Ma et al. reported flower-like MoS2/TiO2 nanohybrid composite photocatalysts obtained from a metal organic framework-derived precursor via facial hydrothermal methods [77]. The flower-like morphology of the MoS2/TiO2 composites was confirmed from SEM images, as shown in Figure 5. In order to investigate the photocatalytic activity, the experiments were conducted under visible light conditions with fluorescein as a photosensitizer. An outstanding improvement in the photocatalytic activity was achieved for the optimized sample (14.6 wt% MoS2 loaded on TiO2) with a hydrogen evolution rate of 10046 μmol·h−1·g−1. They concluded that this high performance of the MoS2/TiO2 composites is associated with the formation of active centers as well as the uniform distribution of MoS2 and TiO2 phases, inducing electrons’ motion to reduce protons. In the proposed photocatalytic activity mechanism, excited electrons from fluorescein transfer to the CB of TiO2. These electrons further move to the surface of MoS2 and combine with protons to produce hydrogen. Liu and his coworker prepared MoS2 nanosheets rooted in TiO2 nanofibers (TiO2@MoS2) using a hydrothermal strategy [19]. They reported single- to few-layer MoS2 nanosheets and TiO2 nanofibers’ porous structure. The MoS2 nanosheets grew vertically on the porous structure of TiO2, and deep rooting MoS2 nanosheets into TiO2 nanofibers put them in close contact for the electron transfer process and structural stability. The hydrogen production rates of the TiO2@MoS2 sample were 1.68 under UV–vis light and 0.49 mmol·h−1·g−1 under visible light.
TiO2 nanomaterials combined with a MoS2 co-catalyst can enhance hydrogen production rates up to several times.

5.2. MoS2/Graphitic Carbon Nitride Composites

Graphitic carbon nitride (g-C3N4) is considered one of the promising candidates for photocatalysis due to its high chemical stability, environmentally friendly nature, and suitable energy bands that can efficiently absorb solar spectrum irradiation [78,79,80,81]. However, g-C3N4 suffers from a small specific surface area, high exciton binding energy, stacking back into a bulk, and low efficiency under visible light [82,83,84]. Recently, much interest has been devoted to g-C3N4-based composites for solar hydrogen production under visible light. To enhance the efficiency of its photocatalytic activity, various non-precious co-catalysts such as Co2P, Mo2C, and MoS2 have been incorporated with C3N4 [85,86,87]. Among them, MoS2 as a co-catalyst in MoS2/C3N4 composites shows promising efficiency for photogenerated hydrogen production [87,88].
The design of and nano-interface coupling between MoS2 and C3N4 can significantly enhance the photocatalytic HER performance. The appropriate MoS2/C3N4 composites with an optimal ratio are believed to enhance solar absorption, increase the interfaces, and decrease the electron transfer distance of the photo-excited electrons between C3N4 and MoS2 co-catalysts. Yuan’s group reported MoS2/g-C3N4 composites with various contents of MoS2 developed using the solvent thermal method. The composite catalysts were evaluated for photocatalytic H2 generation [87]. They found that MoS2/g-C3N4 composites containing 0.75% MoS2 nanosheets performed better and had a reaction rate of 1155 μmol·h−1·g−1 under visible light irradiation. The apparent quantum yield was about 6.8% under a monochromatic light of 420 nm. Furthermore, they explained that the large surface area of g-C3N4 nanosheets and the nano-interface coupling between MoS2 nanosheets and g-C3N4 were mainly responsible for the outstanding photocatalytic hydrogen production of the MoS2/g-C3N4 composite. Recently, Li et al., reported the in situ synthesis of a g-CN/MoS2 composite [89]. The composite exhibited enhanced photocatalytic hydrogen production compared to pristine g-CN under visible light irradiation. The rod-like MoS2 plays an important role as co-catalyst in the g-CN/MoS2 composite in the enhancement of the hydrogen production rate. Zhang et al. reported sulfur-doped C3N4 with covalently crosslinked MoS2 nanosheets (MoS2/SC3N4) for improved photocatalytic hydrogen production [88]. The ultrathin array-like nanosheet structure of the MoS2/SC3N3 composites was observed by SEM characterizations (see Figure 6). MoS2/SC3N3 composites were studied for photocatalytic HER under visible light conditions. MoS2/SC3N3 with 2.5% MoS2 nanosheets showed the optimal hydrogen production rate of 702.53 μmol·h−1·g−1. The array-like porous morphology had a rich exposed surface, covalent bonding structure, and enhanced visible light absorption by the cyano group of MoS2/SC3N3 composites. This facilitates the photogenerated electrons’ transfer from the CB of SC3N3 to MoS2 via a heterojunction interface that consequently enhances the photocatalytic hydrogen evolution. Zhang et al. reported a MoS2/Fe2O3/g-C3N4 ternary composite photocatalyst under hydrothermal conditions for hydrogen production [90]. The obtained ternary composite showed a hydrogen production rate about five times higher compared to g-C3N4. In addition, 1T MoS2/C3N4 composites also show enhanced photocatalytic hydrogen production [91,92,93,94]. Li et al. loaded metallic 1T-phase MoS2 quantum dots onto CdS nanorods (1T-MoS2-CdS) using a one-step hydrothermal method at different temperatures [91]. The 1T-MoS2-CdS composite prepared at 180 ºC showed remarkable photocatalytic hydrogen production (131.7 mmol·h−1·g−1) under visible light (λ > 420 nm). This rate of hydrogen evolution reaction was over 65 times greater than that of pure CdS (mmol·h−1·g−1) and two times that of Pt-loaded CdS.
Besides 1T-phase MoS2, amorphous MoSx nanomaterials are efficient electrocatalysts as well as co-catalysts for hydrogen production [95,96,97]. They provide more unsaturated active S atoms, which can rapidly capture protons from the solution to convert them into hydrogen molecules. Yu et al. reported amorphous MoSx/g-C3N4 (a-MoSx/g-C3N4) composites developed using an adsorption in situ transformation method [95]. The a-MoSx/g-C3N4 composites were compared with crystalline MoSx/g-C3N4 and g-C3N4 catalysts, and all of the a-MoSx/g-C3N4 catalysts displayed better photocatalytic performances than the crystalline MoSx/g-C3N4 and C3N4 catalysts. Among the a-MoSx/g-C3N4 composites, the a-MoSx/g-C3N4 catalyst with 3 wt% Mo showed the best photocatalytic performance and a hydrogen production rate of 273.1 μmol·h−1·g−1.
Similar to TiO2/MoS2 photocatalysts, MoS2/g-C3N4 heterojunction composites can improve hydrogen production.

5.3. MoS2 Coupling with Other Semiconductor Materials

As discussed earlier, MoS2 as a co-catalyst for other semiconductor compounds can efficiently enhance the photocatalytic activity of hydrogen generation. The interfacial coupling of MoS2 with semiconductor compounds has been designed in many strategies. An appropriate ratio, increased interface area, and decreased migration distance of the photogenerated electrons between the MoS2 and the semiconductor compounds can effectively improve photocatalytic hydrogen production. Zhang et al. reported MoS2/CdS composites with willow branch-shaped morphology developed using a one-pot hydrothermal method [98]. The MoS2/CdS composite with 5 wt% MoS2 as a co-catalyst displayed an enhanced photocatalytic performance and produced 250.8 μmol·h−1 hydrogen evolution with an apparent quantum efficiency of 3.66% at 420 nm. Preparation of the willow branch-shaped nano-heterojunction morphology enhances the visible light absorption and also promotes the separation of photogenerated electron–hole pairs.
Ma et al. reported a layered CdS/MoS2 heterostructure photocatalyst developed using ultrasonicated MoS2 and CdS nanosheets, produced from hydro- and solvothermal methods, respectively [99]. When MoS2 co-catalysts were loaded onto CdS nanosheets, the photocatalytic performance of the CdS/MoS2 heterostructure was twice that of the pure CdS photocatalyst. The designing of a layered CdS/MoS2 heterostructure could efficiently enhance the photogenerated charge separation and electron transfer, which improves the surface hydrogen evolution kinetics. Patriarchea and coworkers synthesized CdS nanoparticles using polymer-templated oxidative aggregation, and subsequently, MoS2 nanosheets were deposited on it via the wet chemical method [100]. The obtained optimized MoS2/CdS catalyst showed a good hydrogen production rate of about 0.4 mmol h−1 under visible light compared to the CdS catalyst. The enhanced hydrogen generation was due to the presence of the MoS2 co-catalyst.
Samaniego-Benitez and coworkers prepared ZnS/MoS2 heterostructure materials using a one-pot solvothermal method [101]. The hydrogen production yield of the ZnS/MoS2 sample with 10% Mo reached 2600 μmol·h−1 under UV light for 4 h. They concluded that the enhanced photocatalytic activity was due to the synergistic effect between ZnS and MoS2 and sulfur vacancies created in the ZnS structure during the synthesis process. In the proposed mechanism, a photoexcited electron moves from the CB of ZnS to the CB of MoS2, where it interacts with the proton and produces hydrogen.
Recently, Guan et al., used MoS2 as a co-catalyst for methylammonium lead iodide to split hydrogen iodide for photocatalytic HER [102]. The methylammonium lead iodide microcrystals and MoS2 nanoflowers (MAPbI3/MoS2) formed a heterostructure. The MoS2 nanoflowers have plenty of active catalyst sites for hydrogen evolution. The hydrogen evolution rate of MAPbI3/MoS2 reached ~30,000 μmol·h−1·g−1 and a solar-derived hydrogen iodide splitting efficiency of 7.35% was achieved under visible light irradiation. This hydrogen evolution rate is more than 1000 times higher compared to that of pristine MAPbI3. The MoS2 can induce charge separation and provide abundant active sites for photocatalytic hydrogen evolution.
For these examples, we can conclude that MoS2 is an efficient co-catalyst for CdS, ZnS, and MAPbI3 etc, catalysts to produce hydrogen.

5.4. MoS2 and Other Co-Catalyst Heterojunction Composites

The heterojunction of a MoS2 co-catalyst with other co-catalysts is an attractive strategy because it can improve the photogenerated electron transfer from a semiconductor to a MoS2 co-catalyst during photocatalysis, which enhances the activity via the catalytic sites on MoS2 co-catalysts [103,104,105]. The heterojunctions between MoS2 and highly conductive co-catalysts decrease the resistance effect and increase the electron transfer process during photocatalysis [106].
For improved photocatalytic H2 evolution, a widely studied example of anchoring a MoS2 co-catalyst on graphene has been reported [107,108]. Xiang et al., synthesized a TiO2/MoS2/graphene hybrid photocatalyst for hydrogen production [18]. The hybrid photocatalyst showed significant enhancement of photocatalytic H2 generation under UV illumination, with an apparent quantum efficiency of 9.7% at 365 nm. The improved activity is described in terms of synergetic effects between MoS2 and the conductive graphene co-catalysts and TiO2 leading to outstanding photocatalytic hydrogen evolution activity. These authors have proposed a mechanism for the significant boost of photocatalytic H2 generation. They reported that this enhancement is due to the transfer of photogenerated electrons from the CB of TiO2 nanoparticles to the CB of MoS2 nanosheets through highly conductive graphene sheets (Figure 7), where H+ ions are adsorbed at an active site of MoS2. Apart from graphene, other highly conductive materials such as metal sulfides and phosphides can also be used as interfacial electron transfer sources to enhance photocatalytic hydrogen evolution. Lu and coworkers synthesized g-C3N4, Ni2P, and MoS2 heterojunctions by hydrothermal and ultrasonic methods [109]. The hydrogen production rate of the g-C3N4-1%Ni2P-1.5%MoS2 composite was about 532.41 μmol·h−1·g−1 under visible light, which is 5.15- and 2.47-fold higher than those of g-C3N4-1%Ni2P and g-C3N4-1.5%MoS2, respectively. The Ni2P co-catalyst could be acting as an interface electron bridge between g-C3N4 and MoS2 nanosheets. It provides interfacial electron transfer channels in g-C3N4/MoS2 heterostructure composites and prevents the rapid recombination process of photogenerated charge carriers.
Finally, we summarize some heterojunction composites with semiconducting and MoS2 materials in which the MoS2 nanomaterial acts as a co-catalyst for enhanced photocatalytic hydrogen production. Table 1 and Figure 8 show different strategies used for various types of catalysts combined with a MoS2 co-catalyst to form heterojunction composites for enhanced photocatalytic hydrogen production.
Table 1. Summary of MoS2 usage as a co-catalyst for various materials to form heterostructures for photocatalytic hydrogen generation.
Table 1. Summary of MoS2 usage as a co-catalyst for various materials to form heterostructures for photocatalytic hydrogen generation.
CatalystSynthesis MethodLight SourcePhotocatalytic ActivityNo. of CyclesTotal Times
of Cycles (h)
Ref.
MoS2 nanoparticles/TiO2 nanoparticlesMechanochemistry300 W Xe lamp (λ = 250–380 nm)150.7 μmol·h−1·g−1318[76]
TiO2 nanofibers @MoS2 nanosheetsHydrothermal300 W xenon lamp
λ > 320 nm
or λ > 420 nm
1.68 mmol·h−1·g−1
0.49 mmol·h−1·g−1
630[19]
Flower-like MoS2@TiO2 nanohybridsMetal organic framework-derived300 W Xe lamp (λ ≥ 420 nm)10046 µmol ·h−1·g−1310[77]
MoS2 nanosheets/TiO2 nanotubesHydrothermal process300 W Xe-lamp (λ ≥ 420 nm)143.32 μmol·h−1·g−1414[110]
MoS2 nanosheets/g-C3N4 nanosheetsSolvothermal method300 W Xe-lamp (λ > 420 nm1155 μmol·h−1·g−1312[87]
S-doped C3N4 nanosheets/MoS2 nanosheetsOne-step solid-state strategyVisible LED lamp702.53 μmol·h−1·g−1316[88]
Amorphous MoSx nanoparticles/g-C3N4 nanosheetsAdsorption in situ transformation methodLow-power LEDs (3W, 420 nm)273.1 μmol·h−1·g−1412[95]
g-C3N4/NCDS/MoS2Thermal polymerization and solvothermal approach300 W Xe lamp (λ ≥ 420 nm)212.41 μmol·h−1·g−1416[111]
ZnS/MoS2 particlesOne-pot solvothermalHg pen-lamp (254 nm), (4.4 mW/cm2)606 μmol·h−1·g−1--[101]
MoS2 clusters/CdS nanorodSolvothermal method300 W Xe lamp (λ ≥ 420 nm)12.38 mmol·h−1·g−14 [112]
MoS2/ZnIn2S4 microspheresImpregnation method300 W Xe-lamp (λ > 420 nm)3.06 mmol·h−1·g−1315[113]
MoS2 nanosheets/ZnIn2S4 microspheresIn situ photo-assisted deposition300 W Xe-lamp (λ > 420 nm)8.047 mmol·h−1·g−1--[114]
MoS2 nanoflake-Mn0.2Cd0.8S nanorod/MnS nanoparticleOne-pot solvothermal300 W Xe lamp (λ ≥ 420 nm)995 μmol·h−1520[115]

6. Conclusions and Outlook

In summary, we highlighted the significance of MoS2 as a co-catalyst to improve hydrogen evolution. A comprehensive analysis of the literature led us to conclude that MoS2 is a good co-catalyst for other semiconducting materials such as TiO2, C3N4, CdS, ZnS, etc., which form heterostructure nanocomposites and consequently boost the photocatalytic hydrogen generation ability. However, there are still some critical issues that must be resolved, such as the downsizing of MoS2 nanosheets for appropriate band gap alignment and the high density of catalytic active sites. These issues can be solved by reducing the size of MoS2 to quantum dots or the molecular level, which will certainly enhance the catalytic active sites. The photoexcited electron transfers between photocatalysts and the MoS2 co-catalyst play an important role during photocatalytic hydrogen generation. The electron transfer mechanism at the interface of a semiconductor photocatalyst and a MoS2 co-catalyst is yet to be fully investigated and completely understood. It is important to conduct theoretical studies such as density functional theory (DFT) simulations and apply in situ testing methods to understand electron transfer paths. Although MoS2 nanosheets as a co-catalyst are a promising candidate for photocatalytic hydrogen production, all the challenges require further efforts and study.

Author Contributions

Conceptualization and validation, S.A.S., I.K. and A.Y.; writing—original draft preparation, S.A.S. and I.K.; writing—review and editing, S.A.S. and A.Y.; supervision, A.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

No supporting data is reported.

Acknowledgments

We are thankful to the National Natural Science Foundation of China (under research grant no. 22150410332) and the start-up foundation for the introduction of talent at Jiangsu University of Science and Technology, China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kampouri, S.; Stylianou, K.C. Dual-functional photocatalysis for simultaneous hydrogen production and oxidation of organic substances. ACS Catal. 2019, 9, 4247–4270. [Google Scholar] [CrossRef]
  2. Wang, Q.; Domen, K. Particulate photocatalysts for light-driven water splitting: Mechanisms, challenges, and design strategies. Chem. Rev. 2020, 120, 919–985. [Google Scholar] [CrossRef] [PubMed]
  3. Sanchez-Bastardo, N.; Robert, S.; Ruland, H. Methane pyrolysis for zero-Emission hydrogen production: A potential bridge technology from fossil fuels to a renewable and sustainable hydrogen economy. Ind. Eng. Chem. Res. 2021, 60, 11855–11881. [Google Scholar] [CrossRef]
  4. Shah, S.A.; Shen, X.; Yuan, A.; Ji, Z.; Yue, X.; Zhu, G.; Zhou, H.; Xu, K.; Zhu, J.; Chen, Y. One step in-situ synthesis of Ni3S2/Fe2O3/N-doped carbon composites on Ni foam as an efficient electrocatalyst for overall water splitting. Appl. Surf. Sci. 2020, 527, 146918. [Google Scholar] [CrossRef]
  5. Ola, O.; Ullah, H.; Chen, Y.; Thummavichai, K.; Wang, N.; Zhu, Y. DFT and experimental studies of iron oxide-based nanocomposites for efficient electrocatalysis. J. Mater. Chem. C 2021, 9, 6409–6417. [Google Scholar] [CrossRef]
  6. Wang, Z.; Li, C.; Domen, K. Recent developments in heterogeneous photocatalysts for solar-driven overall water splitting. Chem. Soc. Rev. 2019, 48, 2109–2125. [Google Scholar] [CrossRef]
  7. Zhao, H.; Jiang, Z.; Xiao, K.; Sun, H.; Chan, H.S.; Tsang, T.H.; Yang, S.; Wong, P.K. Photo-assisted separation of noble-metal-free oxidation and reduction cocatalysts for graphitic carbon nitride nanosheets with efficient photocatalytic hydrogen evolution. Appl. Catal. B-Environ. 2021, 280, 119456. [Google Scholar] [CrossRef]
  8. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  9. Wen, J.; Li, X.; Liu, W.; Fang, Y.; Xie, J.; Xu, Y. Strategies for engineering metal-organic frameworks as efficient photocatalysts. Chin. J. Catal. 2015, 36, 2049–2070. [Google Scholar] [CrossRef]
  10. Kong, D.; Fan, H.; Yin, D.; Zhang, D.; Pu, X.; Yao, S.; Su, C. AgFeO2 Nanoparticle/ZnIn2S4 Microsphere p–n heterojunctions with hierarchical nanostructures for efficient visible-light-driven H2 evolution. ACS Sustain. Chem. Eng. 2021, 9, 2673–2683. [Google Scholar] [CrossRef]
  11. Mahler, B.; Hoepfner, V.; Liao, K.; Ozin, G.A. Colloidal synthesis of 1T-WS2 and 2H-WS2 nanosheets: Applications for photocatalytic hydrogen evolution. J. Am. Chem. Soc. 2014, 136, 14121–14127. [Google Scholar] [CrossRef] [PubMed]
  12. Li, Y.; Ding, L.; Yin, S.; Liang, Z.; Xue, Y.; Wang, X.; Cui, H.; Tian, J. Photocatalytic H2 Evolution on TiO2 Assembled with Ti3C2 MXene and Metallic 1T-WS2 as Co-catalysts. Nano-Micro Lett. 2020, 12, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Dai, C.; Liu, B. Conjugated polymers for visible-light-driven photocatalysis. Energy Environ. Sci. 2020, 13, 24. [Google Scholar] [CrossRef]
  14. Wan, Y.; Deng, J.; Gu, C.; Ma, Y. Highly efficient photocatalytic hydrogen evolution based on conjugated molecular micro/nano-crystalline sheets. J. Mater. Chem. A 2021, 9, 2120. [Google Scholar] [CrossRef]
  15. Luo, L.; Gong, Z.; Ma, J.; Wang, K.; Zhu, H.; Li, K.; Xiong, L.; Guo, X.; Tang, J. Ultrathin sulfur-doped holey carbon nitride nanosheets with superior photocatalytic hydrogen production from water. Appl. Catal. B-Environ. 2021, 284, 119742. [Google Scholar] [CrossRef]
  16. Sun, K.; Liu, M.; Pei, J.; Li, D.; Ding, C.; Wu, K.; Jiang, H. Incorporating transition-metal phosphides into metal-organic frameworks for enhanced photocatalysis. Angew. Chem. Int. Ed. 2020, 59, 22749. [Google Scholar] [CrossRef]
  17. Chen, W.; Wang, L.; Mo, D.; He, F.; Wen, Z.; Wu, X.; Xu, H.; Chen, L. Modulating benzothiadiazole-based covalent organic frameworks via halogenation for enhanced photocatalytic water splitting. Angew. Chem. Int. Ed. 2020, 59, 16902. [Google Scholar] [CrossRef]
  18. Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic effect of MoS2 and graphene as cocatalysts for enhanced photocatalytic H2 production activity of TiO2 nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575–6578. [Google Scholar] [CrossRef]
  19. Liu, C.; Wang, L.; Tang, Y.; Luo, S.; Liu, Y.; Zhang, S.; Zeng, Y.; Xu, Y. Vertical single or few-layer MoS2 nanosheets rooting into TiO2 nanofibers for highly efficient photocatalytic hydrogen evolution. Appl. Catal. B-Environ. 2015, 164, 1–9. [Google Scholar] [CrossRef]
  20. Shah, S.A.; Xu, L.; Sayyar, R.; Bian, T.; Liu, Z.; Yuan, A.; Shen, X.; Khan, I.; Tahir, A.A.; Ullah, H. Growth of MoS2 nanosheets on M@N-doped carbon particles (M = Co, Fe or CoFe Alloy) as an efficient electrocatalyst toward hydrogen evolution reaction. Chem. Eng. J. 2022, 428, 132126. [Google Scholar] [CrossRef]
  21. Wilcoxon, J.P.; Newcomer, P.P.; Samara, G.A. Synthesis and optical properties of MoS2 and isomorphous nanoclusters in the quantum confinement regime. J. Appl. Phys. 1997, 81, 7934. [Google Scholar] [CrossRef]
  22. Thurston, T.R.; Wilcoxon, J.P. Photooxidation of organic chemicals catalyzed by nanoscale MoS2. J. Phys. Chem. B 1999, 103, 11–17. [Google Scholar] [CrossRef]
  23. Liu, H.; Li, Y.; Xiang, M.; Zeng, H.; Shao, X. Single-layered MoS2 directly grown on rutile TiO2 (110) for enhanced interfacial charge transfer. ACS Nano 2019, 13, 6083–6089. [Google Scholar] [CrossRef] [PubMed]
  24. Li, Z.; Meng, X.; Zhang, Z. Recent development on MoS2-based photocatalysis: A review. J. Photochem. Photobiol. C 2018, 35, 39–55. [Google Scholar] [CrossRef]
  25. Khalila, R.M.A.; Hussaina, F.; Rana, A.M.; Imran, M.; Murtaza, G. Comparative study of polytype 2H-MoS2 and 3R-MoS2 systems by employing DFT. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 106, 338–345. [Google Scholar] [CrossRef]
  26. Samanian, M.; Ghatee, M.H. Study of the molybdenum dichalcogenide crystals: Recent developments and novelty of the P-MoS2 structure. J. Mol. Model. 2021, 27, 268. [Google Scholar] [CrossRef]
  27. Eidsavg, H.; Rasukkannu, M.; Velauthapillai, D.; Vajeeston, P. In-depth first-principle study on novel MoS2 polymorphs. RSC Adv. 2021, 11, 3759–3769. [Google Scholar] [CrossRef]
  28. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef]
  29. Liang, Z.; Shen, R.; Ng, Y.H.; Zhang, P.; Xiang, Q.; Li, X. A review on 2D MoS2 cocatalysts in photocatalytic H2 production. J. Mater. Sci. Technol. 2020, 56, 89–121. [Google Scholar] [CrossRef]
  30. Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T phase MoS2 nanosheets as supercapacitor electrode materials. Nat. Nanotechnol. 2015, 10, 313–318. [Google Scholar] [CrossRef]
  31. Cheng, P.; Sun, K.; Hu, Y.H. Memristive Behavior and Ideal Memristor of 1T Phase MoS2 Nanosheets. Nano Lett. 2016, 16, 572–576. [Google Scholar] [CrossRef] [PubMed]
  32. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kuc, A.; Zibouche, N.; Heine, T. Influence of quantum confinement on the electronic structure of the transition metal sulfide TS2. Phys. Rev. B 2011, 83, 245213. [Google Scholar] [CrossRef] [Green Version]
  34. Su, L.; Luo, L.L.; Wang, J.W.; Song, T.; Tu, W.X.; Wang, Z.J. Lamellar flower-like porous MoS2 as an efficient cocatalyst to boost photocatalytic hydrogen evolution of CdS. Catal. Sci. Technol. 2021, 11, 1292–1297. [Google Scholar] [CrossRef]
  35. Zhang, S.; Duan, S.; Chen, G.; Meng, S.; Zheng, X.; Fan, Y.; Fu, X.; Chen, S. MoS2/Zn3In2S6 composite photocatalysts for enhancement of visible light-driven hydrogen production from formic acid. Chin. J. Catal. 2021, 42, 193–204. [Google Scholar] [CrossRef]
  36. Sun, J.; Yin, M.; Li, Y.; Liang, K.; Fan, Y.; Li, Z. Efficient photocatalytic hydrogen production of ternary composite constituted by cubic CdS, MoS2 and activated carbon. J. Alloys Compd. 2021, 874, 59930. [Google Scholar] [CrossRef]
  37. Sun, J.; Duan, L.; Wu, Q.; Yao, W. Synthesis of MoS2 quantum dots cocatalysts and their efficient photocatalytic performance for hydrogen evolution. Chem. Eng. J. 2018, 332, 449–455. [Google Scholar] [CrossRef]
  38. Lee, G.J.; Chen, H.C.; Wu, J.J. (In, Cu) Co-doped ZnS nanoparticles for photoelectrochemical hydrogen production. Int. J. Hydrogen Energy 2019, 44, 110–117. [Google Scholar] [CrossRef]
  39. Jeong, S.Y.; Song, J.; Lee, S. Photoelectrochemical device designs toward practical solar water splitting: A review on the recent progress of BiVO4 and BiFeO3 photoanodes. Appl. Sci. 2018, 8, 1388. [Google Scholar] [CrossRef] [Green Version]
  40. Zhou, K.G.; Mao, N.N.; Wang, H.X.; Peng, Y.; Zhang, H.L. A mixed-solvent strategy for efficient exfoliation of inorganic graphene analogues. Angew. Chem. Int. Ed. 2011, 50, 10839–10842. [Google Scholar] [CrossRef]
  41. Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.; Tok, A.I.; Zhang, Q.; Zhang, H. Fabrication of single- and multilayer MoS2 film-based field-effect transistors for sensing NO at room temperature. Small 2012, 8, 63–67. [Google Scholar] [CrossRef] [PubMed]
  42. Hu, D.; Xiang, J.; Zhou, Q.W.; Su, S.Q.; Zhang, Z.B.; Wang, X.; Jin, M.L.; Nian, L.; Nozel, R.; Zhou, G.F.; et al. One-step chemical vapor deposition of MoS2 nanosheets on SiNWs as photocathodes for efficient and stable solar-driven hydrogen production. Nanoscale 2018, 10, 3518–3525. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, T.; Chen, W.; Liu, T.Y.; Hao, Q.L.; Liu, X.H. Hybrid of AgInZnS and MoS2 as efficient visible-light driven photocatalyst for hydrogen production. Int. J. Hydrogen Energy 2017, 42, 12254–12261. [Google Scholar] [CrossRef]
  44. Wei, J.; Huang, H.; Luo, Q.; Liu, N.; Wang, X.Y.; Wang, X.F.; Zhong, M.; Huang, X.L. Synthesis of few layers amorphous 1T/2H MoS2 by a one-step ethanol/water solvothermal method and its hydrodesulfurization performance. Catal. Lett. 2022, 152, 263–275. [Google Scholar] [CrossRef]
  45. Yang, Y.; Liu, Z.; Shu, K.; Li, L.; Li, J. Improved performances of CVD-grown MoS2 based phototransistors enabled by encapsulation. Adv. Mater. Interfaces 2021, 8, 2100164. [Google Scholar] [CrossRef]
  46. Li, H.; Wu, J.; Yin, Z.; Zhang, H. Preparation and applications of mechanically exfoliated single-layer and multilayer MoS2 and WSe2 nanosheets. Acc. Chem. Res. 2014, 47, 1067–1075. [Google Scholar] [CrossRef]
  47. Jeon, J.; Lee, J.; Yoo, G.; Park, J.H.; Yeom, G.Y.; Jang, Y.H.; Lee, S. Size-tunable synthesis of monolayer MoS2 nanoparticles and their applications in non-volatile memory devices. Nanoscale 2016, 8, 16995–17003. [Google Scholar] [CrossRef]
  48. Zhang, S.; Liu, X.; Liu, C.; Luo, S.; Wang, L.; Cai, T.; Zeng, Y.; Yuan, J.; Dong, W.; Pei, Y.; et al. MoS2 quantum dots growth induced by S vacancy in ZnIn2S4 monolayer: Atomic-level heterostructure for photocatalytic hydrogen production. ACS Nano 2018, 12, 751–758. [Google Scholar] [CrossRef]
  49. Huang, Y.; Pan, Y.H.; Yang, R.; Bao, L.H.; Meng, L.; Luo, H.L.; Cai, Y.Q.; Liu, G.D.; Zhao, W.J.; Zhou, Z.; et al. Universal mechanical exfoliation of large-area 2D crystals. Nat. Commun. 2020, 11, 2453. [Google Scholar] [CrossRef]
  50. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef]
  51. Ambrosi, A.; Pumera, M. Electrochemical exfoliation of MoS2 crystal for hydrogen electrogeneration. Chem. Eur. J. 2018, 24, 18551–18555. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Y.; He, X.; Hanlon, D.; Harvey, A.; Coleman, J.N.; Li, Y. Liquid phase exfoliated MoS2 nanosheets percolated with carbon nanotubes for high volumetric/areal capacity sodium-ion batteries. ACS Nano 2016, 10, 8821–8828. [Google Scholar] [CrossRef] [PubMed]
  53. Ottaviano, L.; Palleschi, S.; Perrozzi, F.; DOlimpio, G.; Priante, F.; Donarelli, M.; Benassi, P.; Nardone, M.; Gonchigsuren, M.; Gombosuren, M.; et al. Mechanical exfoliation and layer number identification of MoS2 revisited. 2D Mater. 2017, 4, 045013. [Google Scholar] [CrossRef]
  54. Zhang, H. Ultrathin two-dimensional nanomaterials. ACS Nano 2015, 9, 9451–9469. [Google Scholar] [CrossRef] [PubMed]
  55. Magda, G.; Peto, J.; Dobrik, G.; Hwang, C.; Biro, L.P.; Tapaszto, L. Exfoliation of large-area transition metal chalcogenide single layers. Sci. Rep. 2015, 5, 14714. [Google Scholar] [CrossRef] [Green Version]
  56. Desai, S.B.; Madhvapathy, S.R.; Amani, M.; Kiriya, D.; Hettick, M.; Tosun, M.; Zhou, Y.; Dubey, M.; Ager, J.W.; Chrzan, D.; et al. Gold-mediated exfoliation of ultralarge optoelectronically-perfect monolayers. Adv. Mater. 2016, 28, 4053–4058. [Google Scholar] [CrossRef]
  57. Velicky, M.; Donnelly, G.E.; Hendren, W.R.; McFarland, S.; Scullion, D.; DeBenedetti, W.J.I.; Correa, G.C.; Han, Y.; Wain, A.J.; Hines, M.A.; et al. Mechanism of Gold-assisted exfoliation of centimeter-sized transition-metal dichalcogenide monolayers. ACS Nano 2018, 12, 10463–10472. [Google Scholar] [CrossRef] [Green Version]
  58. Lin, Y.C.; Zhang, W.; Huang, J.K.; Liu, K.K.; Lee, Y.H.; Liang, C.T.; Chu, C.W.; Li, L.J. Wafer-scale MoS2 thin layers prepared by MoO3 sulfurization. Nanoscale 2012, 4, 6637–6641. [Google Scholar] [CrossRef]
  59. Cai, J.; Jian, J.; Chen, X.; Lei, M.; Wang, W. Regular hexagonal MoS2 microflakes grown from MoO3 precursor. Appl. Phys. A 2007, 89, 783–788. [Google Scholar] [CrossRef]
  60. Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P.; Lou, J. Large-area vapor-phase growth and characterization of MoS2 atomic layers on a SiO2 substrate. Small 2014, 8, 966–971. [Google Scholar] [CrossRef] [Green Version]
  61. Song, I.; Park, C.; Hong, M.; Baik, J.; Shin, H.-J.; Choi, H. Patternable large-scale molybdenium disulfide atomic layers grown by gold-assisted chemical vapor deposition. Angew. Chem. Int. Ed. 2014, 53, 1266–1269. [Google Scholar] [CrossRef] [PubMed]
  62. Ahn, C.; Lee, J.; Kim, H.-U.; Bark, H.; Jeon, M.; Ryu, G.; Lee, Z.; Yeom, G.; Kim, K.; Jung, J.; et al. Low-temperature synthesis of large-scale molybdenum disulfide thin films directly on a plastic substrate using plasma-enhanced chemical vapor deposition. Adv. Mater. 2015, 27, 5223–5229. [Google Scholar] [CrossRef] [PubMed]
  63. Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P.; Mak, K.; Kim, C.-J.; Muller, D.; Park, J. High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity. Nature 2015, 520, 656–660. [Google Scholar] [CrossRef] [PubMed]
  64. Kumar, V.; Dhar, S.; Choudhury, T.; Shivashankar, S.; Raghavan, S. A predictive approach to CVD of crystalline layers of TMDs: The case of MoS2. Nanoscale 2015, 7, 7802–7810. [Google Scholar] [CrossRef]
  65. Vishwanath, S.; Liu, X.; Rouvimov, S.; Mende, P.; Azcatl, A.; McDonnell, S.; Wallace, R.; Feenstra, R.; Furdyna, J.; Jena, D.; et al. Comprehensive structural and optical characterization of MBE grown MoSe2 on graphite, CaF2 and graphene. 2D Mater. 2015, 2, 024007. [Google Scholar] [CrossRef]
  66. Qin, X.; Ke, P.; Wang, A.; Kim, K. Microstructure, mechanical and tribological behaviors of MoS2-Ti composite coatings deposited by a hybrid HIPIMS method. Surf. Coat. Technol. 2013, 228, 275–281. [Google Scholar] [CrossRef]
  67. Feng, X.; Tang, Q.; Zhou, J.; Fang, J.; Ding, P.; Sun, L.; Shi, L. Novel mixed–solvothermal synthesis of MoS2 nanosheets with controllable morphologies. Cryst. Res. Technol. 2013, 48, 363–368. [Google Scholar] [CrossRef]
  68. Liu, Y.; Xu, X.; Li, H.; Si, Z.; Wu, X.; Ra, R.; Duan, W. A facile one step synthesis of MoS2/g-C3N4 photocatalyst with enhanced visible light photocatalytic hydrogen production. Catal. Lett. 2022, 152, 972–979. [Google Scholar] [CrossRef]
  69. Deng, Z.; Hu, Y.; Ren, D.; Lin, S.; Jiang, H.; Li, C. Reciprocal hybridization of MoO2 nanoparticles and few-layer MoS2 for stable lithium-ion batteries. Chem. Commun. 2015, 51, 13838–13841. [Google Scholar] [CrossRef]
  70. Jagminas, A.; Naujokaitis, A.; Gaigalas, P.; Ramanavicius, S.; Kurtinaitiene, M.; Trusovas, R. Substrate impact on the structure and electrocatalyst properties of molybdenum disulfide for HER from water. Metals 2020, 10, 1251. [Google Scholar] [CrossRef]
  71. Guo, X.; Wang, Z.; Zhu, W.; Yang, H. The novel and facile preparation of multilayer MoS2 crystals by a chelation-assisted sol-gel method and their electrochemical performance. RSC Adv. 2017, 7, 9009–9014. [Google Scholar] [CrossRef] [Green Version]
  72. Shah, S.A.; Shen, X.; Xie, M.; Zhu, G.; Ji, Z.; Zhou, H.; Xu, K.; Yue, X.; Yuan, A.; Zhu, J.; et al. Nickel@nitrogen-doped carbon@MoS2 nanosheets: An efficient electrocatalyst for hydrogen evolution reaction. Small 2019, 15, 1804545. [Google Scholar] [CrossRef] [PubMed]
  73. Shah, S.A.; Zhu, G.; Shen, X.; Kong, L.; Ji, Z.; Xu, K.; Zhou, H.; Zhu, J.; Song, P.; Song, C.; et al. Controllable sandwiching of reduced graphene oxide in hierarchical defect-rich MoS2 ultrathin nanosheets with expanded interlayer spacing for electrocatalytic hydrogen evolution reaction. Adv. Mater. Interfaces 2018, 5, 1801093. [Google Scholar] [CrossRef]
  74. Zhu, J.; Zhu, S.; Kong, X.; Liang, Y.; Li, Z.; Wu, S.; Luo, S.; Chang, C.; Cui, Z. Rutile-coated B-phase TiO2 heterojunction nanobelts for photocatalytic H2 evolution. ACS Appl. Nano Mater. 2020, 3, 10349–10359. [Google Scholar] [CrossRef]
  75. Yang, W.; Li, M.; Pan, K.; Guo, L.; Wu, J.; Li, Z.; Yang, F.; Lin, K.; Zhou, W. Surface engineering of mesoporous anatase titanium dioxide nanotubes for rapid spatial charge separation on horizontal-vertical dimensions and efficient solar-driven photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2021, 586, 75–83. [Google Scholar] [CrossRef]
  76. 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. [Google Scholar] [CrossRef]
  77. Ma, B.; Guan, P.Y.; Li, Q.Y.; Zhang, M.; Zang, S.Q. MOF derived flower-like MoS2@TiO2 nanohybrids with enhanced activity for hydrogen evolution. ACS Appl. Mater. Interfaces 2016, 8, 26794–26800. [Google Scholar] [CrossRef]
  78. Zhou, Z.; Zhang, Y.; Shen, Y.; Liu, S.; Zhang, Y. Molecular engineering of polymeric carbon nitride: Advancing applications from photocatalysis to biosensing and more. Chem. Soc. Rev. 2018, 47, 2298–2321. [Google Scholar] [CrossRef]
  79. Khan, I.; Luo, M.; Guo, L.; Khan, S.; Shah, S.A.; Khan, I.; Khan, A.; Wang, C.; Ai, B.; Zaman, S. Synthesis of phosphate-bridged g-C3N4/LaFeO3 nanosheets Z-scheme nanocomposites as efficient visible photocatalysts for CO2 reduction and malachite green degradation. Appl. Catal. A Gen. 2022, 629, 118418. [Google Scholar] [CrossRef]
  80. Zhao, D.; Dong, C.; Wang, B.; Chen, C.; Huang, Y.; Diao, Z.; Li, S.; Guo, L.; Shen, S. Synergy of dopants and defects in graphitic carbon nitride with exceptionally modulated band structures for efficient photocatalytic oxygen evolution. Adv. Mater. 2019, 31, 1903545. [Google Scholar] [CrossRef]
  81. Han, X.; Xu, D.; An, L.; Hou, C.; Li, Y.; Zhang, Q.; Wang, H. Ni-Mo nanoparticles as co-catalyst for drastically enhanced photocatalytic hydrogen production activity over g-C3N4. Appl. Catal. B-Environ. 2019, 243, 136–144. [Google Scholar] [CrossRef]
  82. Wu, X.; Gao, D.; Yu, H.; Yu, J. High-yield lactic acid-mediated route for a g-C3N4 nanosheet photocatalyst with enhanced H2-evolution performance. Nanoscale 2019, 11, 9608–9616. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, G.; Li, G.; Heil, T.; Zafeiratos, S.; Lai, F.; Savateev, A.; Antonietti, M.; Wang, X. Tailoring the grain boundary chemistry of polymeric carbon nitride for enhanced solar hydrogen production and CO2 reduction. Angew. Chem. Int. Ed. 2019, 58, 3433–3437. [Google Scholar] [CrossRef] [PubMed]
  84. Khan, I.; Luo, M.; Khan, S.; Asghar, H.; Saeed, M.; Khan, S.; Khan, A.; Humayun, M.; Guo, L.; Shi, B. Green synthesis of SrO bridged LaFeO3/g-C3N4 nanocomposites for CO2 conversion and bisphenol A degradation with new insights into mechanism. Environ. Res. 2022, 7, 112650. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, J.; Su, F.; Wen, P.; Wang, Y.; Zhao, B.; Jiang, L.; Sun, Y. In situ growth of Co2P nanocrystal on g-C3N4 for efficient and stable photocatalytic hydrogen evolution. Energy Fuels 2021, 35, 1859–1865. [Google Scholar] [CrossRef]
  86. Zhang, J.; Wu, M.; He, B.B.; Wang, R.; Wang, H.W.; Gong, Y.S. Facile synthesis of rod-like g-C3N4 by decorating Mo2C co-catalyst for enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2019, 470, 565–572. [Google Scholar] [CrossRef]
  87. Yuan, Y.J.; Shen, Z.; Wu, S.; Su, Y.; Pei, L.; Ji, Z.; Ding, M.; Bai, W.; Chen, Y.; Yu, Z.T.; et al. Liquid exfoliation of g-C3N4 nanosheets to construct 2D-2D MoS2/g-C3N4 photocatalyst for enhanced photocatalytic H2 production activity. Appl. Catal. B-Environ. 2019, 246, 120–128. [Google Scholar] [CrossRef]
  88. Zhang, S.; Dong, H.; An, C.; Li, Z.; Xu, D.; Xu, K.; Wu, Z.; Shen, J.; Chen, X.; Zhang, S. One-pot synthesis of array-like sulfur-doped carbon nitride with covalently crosslinked ultrathin MoS2 cocatalyst for drastically enhanced photocatalytic hydrogen evolution. J. Mater. Sci. Technol. 2021, 75, 59–67. [Google Scholar] [CrossRef]
  89. Li, J.; Wu, W.; Li, Y.; Zhang, H.; Xu, X.; Jiang, Y.; Lin, K. In situ synthesized rodlike MoS2 as a cocatalyst for enhanced photocatalytic hydrogen evolution by graphitic carbon nitride without a noble metal. ACS Appl. Energy Mater. 2021, 4, 11836–11843. [Google Scholar] [CrossRef]
  90. Zhang, Y.; Wan, J.; Zhang, C.; Cao, X. MoS2 and Fe2O3 co-modify g-C3N4 to improve the performance of photocatalytic hydrogen production. Sci. Rep. 2022, 12, 3261. [Google Scholar] [CrossRef]
  91. Li, X.; Lv, X.; Li, N.; Wu, J.; Zheng, Y.Z.; Tao, X. One-step hydrothermal synthesis of high-percentage 1T-phase MoS2 quantum dots for remarkably enhanced visible-light-driven photocatalytic H2 evolution. Appl. Catal. B-Environ. 2019, 243, 76–85. [Google Scholar] [CrossRef]
  92. Liang, Z.; Sun, B.; Xu, X.; Cui, H.; Tian, J. Metallic 1T-phase MoS2 quantum dots/g-C3N4 heterojunctions for enhanced photocatalytic hydrogen evolution. Nanoscale 2019, 11, 12266–12274. [Google Scholar] [CrossRef] [PubMed]
  93. Liang, Z.; Guo, Y.; Xue, Y.; Cui, H.; Tian, J. 1T-phase MoS2 quantum dots as a superior co-catalyst to Pt decorated on carbon nitride nanorods for photocatalytic hydrogen evolution from water. Mater. Chem. Front. 2019, 3, 2032–2040. [Google Scholar] [CrossRef]
  94. Shi, J.; Zou, Y.; Ma, D.; Fan, Z.; Cheng, L.; Sun, D.; Wang, Z.; Niu, C.; Wang, L. Stable 1T-phase MoS2 as an effective electron mediator promoting photocatalytic hydrogen production. Nanoscale 2018, 10, 9292–9303. [Google Scholar] [CrossRef]
  95. Yu, H.; Xiao, P.; Wang, P.; Yu, J. Amorphous molybdenum sulfide as highly efficient electron-cocatalyst for enhanced photocatalytic H2 evolution. Appl. Catal. B-Environ. 2016, 193, 217–225. [Google Scholar] [CrossRef]
  96. Peng, X.; Huang, C.; Zhang, B.; Liu, Y. Atomic-scale intercalation of amorphous MoS2 nanoparticles into N-doped carbon as a highly efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 27193–27201. [Google Scholar] [CrossRef]
  97. Lee, S.C.; Benck, J.D.; Tsai, C.; Park, J.; Koh, A.L.; Abild-Pedersen, F.; Jaramillo, T.F.; Sinclair, R. Chemical and phase evolution of amorphous molybdenum sulfide catalysts for electrochemical hydrogen production. ACS Nano 2016, 10, 624–632. [Google Scholar] [CrossRef]
  98. Zhang, Z.W.; Li, Q.H.; Qiao, X.Q.; Hou, D.; Li, D.S. One-pot hydrothermal synthesis of willow branch-shaped MoS2/CdS heterojunctions for photocatalytic H2 production under visible light irradiation. J. Catal. 2019, 40, 371–379. [Google Scholar] [CrossRef]
  99. Ma, S.; Xie, J.; Wen, J.Q.; He, K.L.; Li, X.; Liu, W.; Zhang, X.C. Constructing 2D layered hybrid CdS nanosheets/MoS2 heterojunctions for enhanced visible-light photocatalytic H2 generation. Appl. Surf. Sci. 2017, 391, 580–591. [Google Scholar] [CrossRef]
  100. Patriarchea, C.; Vamvasakis, I.; Koutsouroubi, E.D.; Armatas, G.S. Enhancing interfacial charge transfer in mesoporous MoS2/CdS nanojunction architectures for highly efficient visible-light photocatalytic water splitting. Inorg. Chem. Front. 2022, 9, 625–636. [Google Scholar] [CrossRef]
  101. Samaniego-Beniteza, J.E.; Lartundo-Rojas, L.; García-Garcíac, A.; Calderón, H.A.; Mantilla, A. One-step synthesis and photocatalytic behavior for H2 production from water of ZnS/MoS2 composite material. Catal. Today 2021, 360, 99–105. [Google Scholar] [CrossRef]
  102. Guan, W.; Li, Y.; Zhong, Q.; Liu, H.; Chen, J.; Hu, H.; Lv, K.; Gong, J.; Xu, Y.; Kang, Z.; et al. Fabricating MAPbI3/MoS2 Composites for improved photocatalytic performance. Nano Lett. 2021, 21, 597–604. [Google Scholar] [CrossRef] [PubMed]
  103. Meng, A.Y.; Zhang, J.; Xu, D.F.; Cheng, B.; Yu, J.G. Enhanced photocatalytic H2-production activity of anatase TiO2 nanosheet by selectively depositing dual-cocatalysts on {101} and {001} facets. Appl. Catal. B-Environ. 2016, 198, 286–294. [Google Scholar] [CrossRef]
  104. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual Cocatalysts in TiO2 Photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef]
  105. Zhao, S.; Xu, J.; Li, Z.T.; Liu, Z.Y.; Li, Y.R. Molybdenum disulfide coated nickel-cobalt sulfide with nickel phosphide loading to build hollow core-shell structure for highly efficient photocatalytic hydrogen evolution. J. Colloid. Interf. Sci. 2019, 555, 689–701. [Google Scholar] [CrossRef]
  106. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for selective photoreduction of CO2 into solar fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef]
  107. Lee, G.J.; Hou, Y.H.; Chen, C.Y.; Tsay, C.Y.; Chang, Y.C.; Chen, J.H.; Horng, T.L.; Anandan, S.; Wu, J.J. Enhanced performance for photocatalytic hydrogen evolution using MoS2/graphene hybrids. Int. J. Hydrogen Energy 2021, 46, 5938–5948. [Google Scholar] [CrossRef]
  108. Wang, X.; Long, R. Rapid charge separation boosts solar hydrogen generation at the graphene–MoS2 Junction: Time-domain Ab initio analysis. J. Phys. Chem. Lett. 2021, 12, 2763–2769. [Google Scholar] [CrossRef]
  109. Lu, X.Y.; Xie, J.; Chen, X.B.; Li, X. Engineering MPx (M = Fe, Co or Ni) interface electron transfer channels for boosting photocatalytic H2 evolution over g-C3N4/MoS2 layered heterojunctions. Appl. Catal. B-Environ. 2019, 252, 250–259. [Google Scholar] [CrossRef]
  110. Cao, D.; Wang, Q.; Zhu, S.; Zhang, X.; Li, Y.; Cui, Y.; Xue, Z.; Gao, S. Hydrothermal construction of flower-like MoS2 on TiO2 NTs for highly efficient environmental remediation and photocatalytic hydrogen evolution. Sep. Purif. Technol. 2021, 265, 118463. [Google Scholar] [CrossRef]
  111. Jiao, Y.; Huang, Q.; Wang, J.; He, Z.; Li, Z. A novel MoS2 quantum dots (QDs) decorated Z-scheme g-C3N4 nanosheet/N-doped carbon dots heterostructure photocatalyst for photocatalytic hydrogen evolution. Appl. Catal. B-Environ. 2019, 247, 124–132. [Google Scholar] [CrossRef]
  112. Feng, C.; Chen, Z.; Hou, J.; Li, J.; Li, X.; Xu, L.; Sun, M.; Zeng, R. Effectively enhanced photocatalytic hydrogen production performance of one-pot synthesized MoS2 clusters/CdS nanorod heterojunction material under visible light. Chem. Eng. J. 2018, 345, 404–413. [Google Scholar] [CrossRef]
  113. 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-Environ. 2014, 144, 521–527. [Google Scholar] [CrossRef]
  114. Chen, G.; Ding, N.; Li, F.; Fan, Y.; Luo, Y.; Li, D.; Meng, Q. Enhancement of photocatalytic H2 evolution on ZnIn2S4 loaded with in-situ photo-deposited MoS2 under visible light irradiation. Appl. Catal. B-Environ. 2019, 241, 130–140. [Google Scholar] [CrossRef]
  115. Wang, J.; Luo, J.; Liu, D.; Chen, S.; Peng, T. One-pot solvothermal synthesis of MoS2-modified Mn0.2Cd0.8S/MnS heterojunction photocatalysts for highly efficient visible-light-driven H2 production. Appl. Catal. B-Environ. 2014, 160–161, 614–620. [Google Scholar] [CrossRef]
Figure 1. (a) Side and (b) top views of MoS2 single layer.
Figure 1. (a) Side and (b) top views of MoS2 single layer.
Molecules 27 03289 g001
Figure 2. (a) Schematic device illustration of photoelectrochemical water splitting. Reprinted with permission from Ref. [38] (Copyright 2019 Elsevier). (b) Schematic representation of the photoelectrochemical water splitting process in a common PEC water splitting system consisting of a photoanode and a metal counterpart. Reprinted from Ref. [39].
Figure 2. (a) Schematic device illustration of photoelectrochemical water splitting. Reprinted with permission from Ref. [38] (Copyright 2019 Elsevier). (b) Schematic representation of the photoelectrochemical water splitting process in a common PEC water splitting system consisting of a photoanode and a metal counterpart. Reprinted from Ref. [39].
Molecules 27 03289 g002
Figure 3. Various synthetic methods for MoS2 preparation.
Figure 3. Various synthetic methods for MoS2 preparation.
Molecules 27 03289 g003
Figure 4. Mechanically exfoliated single- and multilayer MoS2 nanosheets on Si/SiO2. (AD) Optical microscope and (EH) AFM images of MoS2 nanosheets. The single MoS2 sheet thickness is 0.8 nm (E), while the thickness of two (F), three (G), and four (H) layers of MoS2 nanosheets is 1.5, 2.1, and 2.9 nm, respectively. Reprinted with permission from Ref. [41]. Copyright 2012 Wiley-VCH Verlag GmbH & Co.
Figure 4. Mechanically exfoliated single- and multilayer MoS2 nanosheets on Si/SiO2. (AD) Optical microscope and (EH) AFM images of MoS2 nanosheets. The single MoS2 sheet thickness is 0.8 nm (E), while the thickness of two (F), three (G), and four (H) layers of MoS2 nanosheets is 1.5, 2.1, and 2.9 nm, respectively. Reprinted with permission from Ref. [41]. Copyright 2012 Wiley-VCH Verlag GmbH & Co.
Molecules 27 03289 g004
Figure 5. (a,b) SEM images of MoS2@TiO2 composites. Reprinted with permission from Ref. [77]. Copyright 2016 American Chemical Society.
Figure 5. (a,b) SEM images of MoS2@TiO2 composites. Reprinted with permission from Ref. [77]. Copyright 2016 American Chemical Society.
Molecules 27 03289 g005
Figure 6. SEM images of (a) SC3N2, (b) MoS2 and ultrathin array-like nanosheet, (c) MoS2/SC3N4-0.5%, (d) MoS2/SC3N4-1.5%, (e), MoS2/SC3N4-2.5%, and (f) MoS2/SC3N4-5.0%. Reprinted with permission from Ref. [88]. Copyright 2021 Elsevier.
Figure 6. SEM images of (a) SC3N2, (b) MoS2 and ultrathin array-like nanosheet, (c) MoS2/SC3N4-0.5%, (d) MoS2/SC3N4-1.5%, (e), MoS2/SC3N4-2.5%, and (f) MoS2/SC3N4-5.0%. Reprinted with permission from Ref. [88]. Copyright 2021 Elsevier.
Molecules 27 03289 g006
Figure 7. Schematic illustration of the charge transfer and proposed mechanism of electron transfer in TiO2/MoS2/graphene composites. Reprinted with permission from Ref. [18]. Copyright 2012 American Chemical Society.
Figure 7. Schematic illustration of the charge transfer and proposed mechanism of electron transfer in TiO2/MoS2/graphene composites. Reprinted with permission from Ref. [18]. Copyright 2012 American Chemical Society.
Molecules 27 03289 g007
Figure 8. Photocatalytic hydrogen production of heterojunction materials using MoS2 as a co-catalyst. (i) MoS2 nanosheets/TiO2 nanotubes [110]. (ii) g-C3N4/NCDS/MoS2 [111]. (iii) Amorphous MoSx nanoparticles/g-C3N4 nanosheets [95]. (iv) TiO2 nanofibers/@MoS2 nanosheets [19]. (v) ZnS/MoS2 particles [101]. (vi) S-doped C3N4 nanosheets/MoS2 nanosheets [88]. (vii) MoS2 nanosheets/g-C3N4 nanosheets [87]. (viii) MoS2/ZnIn2S4 microspheres [113]. (ix) MoS2 nanosheets/ZnIn2S4 microspheres [114]. (x) Flower-like MoS2@TiO2 nanohybrids [77]. (xi) MoS2 clusters/CdS nanorod [112].
Figure 8. Photocatalytic hydrogen production of heterojunction materials using MoS2 as a co-catalyst. (i) MoS2 nanosheets/TiO2 nanotubes [110]. (ii) g-C3N4/NCDS/MoS2 [111]. (iii) Amorphous MoSx nanoparticles/g-C3N4 nanosheets [95]. (iv) TiO2 nanofibers/@MoS2 nanosheets [19]. (v) ZnS/MoS2 particles [101]. (vi) S-doped C3N4 nanosheets/MoS2 nanosheets [88]. (vii) MoS2 nanosheets/g-C3N4 nanosheets [87]. (viii) MoS2/ZnIn2S4 microspheres [113]. (ix) MoS2 nanosheets/ZnIn2S4 microspheres [114]. (x) Flower-like MoS2@TiO2 nanohybrids [77]. (xi) MoS2 clusters/CdS nanorod [112].
Molecules 27 03289 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shah, S.A.; Khan, I.; Yuan, A. MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review. Molecules 2022, 27, 3289. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103289

AMA Style

Shah SA, Khan I, Yuan A. MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review. Molecules. 2022; 27(10):3289. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103289

Chicago/Turabian Style

Shah, Sayyar Ali, Iltaf Khan, and Aihua Yuan. 2022. "MoS2 as a Co-Catalyst for Photocatalytic Hydrogen Production: A Mini Review" Molecules 27, no. 10: 3289. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27103289

Article Metrics

Back to TopTop