Next Article in Journal
Advancing the Industrial Sectors Participation in Demand Response within National Electricity Grids
Previous Article in Journal
Effect of Burner Wall Material on Microjet Hydrogen Diffusion Flames near Extinction: A Numerical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress and Approaches on Transition Metal Chalcogenides for Hydrogen Production

by
Sivagowri Shanmugaratnam
1,2,
Elilan Yogenthiran
3,
Ranjit Koodali
4,*,
Punniamoorthy Ravirajan
1,
Dhayalan Velauthapillai
2 and
Yohi Shivatharsiny
3,*
1
Clean Energy Research Laboratory (CERL), Department of Physics, University of Jaffna, Jaffna 40000, Sri Lanka
2
Faculty of Engineering and Science, Western Norway University of Applied Sciences, P.O. Box 7030, 5020 Bergen, Norway
3
Department of Chemistry, University of Jaffna, Jaffna 40000, Sri Lanka
4
Department of Chemistry, Western Kentucky University, 1906 College Heights Blvd., Bowling Green, KY 42101, USA
*
Authors to whom correspondence should be addressed.
Submission received: 1 November 2021 / Revised: 30 November 2021 / Accepted: 3 December 2021 / Published: 8 December 2021

Abstract

:
Development of efficient and affordable photocatalysts is of great significance for energy production and environmental sustainability. Transition metal chalcogenides (TMCs) with particle sizes in the 1–100 nm have been used for various applications such as photocatalysis, photovoltaic, and energy storage due to their quantum confinement effect, optoelectronic behavior, and their stability. In particular, TMCs and their heterostructures have great potential as an emerging inexpensive and sustainable alternative to metal-based catalysts for hydrogen evolution. Herein, the methods used for the fabrication of TMCs, characterization techniques employed, and the different methods of solar hydrogen production by using different TMCs as photocatalyst are reviewed. This review provides a summary of TMC photocatalysts for hydrogen production.

Graphical Abstract

1. Introduction

Renewably produced hydrogen gas is used as an alternative clean fuel for transportation. It is regarded as a clean and efficient carrier of energy and is considered as an ideal alternative to fossil fuels to meet the growing energy demand and address environmental issues [1]. When compared with other combustible fuels (i.e., deposited fossil fuels such as coal, oil, and natural gas [2]) hydrogen possesses a high energy density (140 MJkg−1) and the end product of this combustion reaction is only water [3,4]. Therefore, the growing energy demand requires developments in the storage of hydrogen, their transport, and production techniques. Although great progress has been made in H2 storage [5,6,7] and transport, the utilization of H2 energy is stymied by the lack of inexpensive and efficient production methods [8]. Hydrogen can be produced in different ways (Figure 1) including steam-methane reforming of natural gas, coal gasification, and biomass conversion [9], which are expensive and produce greenhouse gases such as carbon monoxide and carbon dioxide that are not favorable to the environment [10,11]. An attractive alternative technique, solar assisted water splitting, is considered as one of the most suitable and greener ways to produce hydrogen energy [12] and the technology of water splitting becomes mature day-by-day [13].
The simple method for the separation of water is electrolysis and various clean energy resources such as solar, geothermal, biomass, wind, ocean, or other renewables can be used for this process. Among them, using solar energy, water can be split by two methods: photoelectrochemical and photocatalytic water splitting.
Various types of photocatalytic materials have been utilized for photocatalytic applications. TiO2 [14,15,16,17], ZnO [18], mixed oxide materials [19], perovskite [20,21], dye [22], and metal-doped oxide materials [23,24] are some of them. Among these, transition metal chalcogenides (TMCs) have gained worldwide attention due to their electrocatalytic properties that include indirect band gaps, optoelectronic behavior, and their stability [25]. In addition, nanodots (quantum dots)/nano structures of metal chalcogenides show enhanced catalytic activity at the edges. Furthermore, by limiting their particle size in the 1–10 nm range, due to the quantum confinement effect, it is possible to utilize under solar simulated irradiation [26,27]. Generation of exciton pairs and the effective charge separation in these semiconductor photocatalysts lead to an effective photolysis process. The efficiency of charge transport mainly depends on the crystallinity, particle, and crystallite size of the materials among many other factors. Composites of TMCs with co-catalysts such as graphene [28], graphene oxides [29], carbon nitride [30], metal oxides [19,31,32,33], CdLa2S4 [34], and metals [35] have been found to increase the conductivity of electrons, provide active sites, and effectively separate the electron and hole pairs generated by bandgap irradiation of the semiconductor photocatalysts. This review mainly focuses on the synthesis and characterization of various types of TMCs and their application in hydrogen energy production through water splitting. This review may assist researchers by providing insights on the fabrication of highly efficient photocatalysts with low cost and high efficiency.

2. Transition Metal Chalcogenides (TMCs)

TMCs have gained worldwide attention in recent decades. These catalysts have been used in lithium-ion batteries, solar cells, and hydrogen evolution reactions. TMCs can be excited by solar simulated irradiation and due to their indirect band gaps, optoelectronic behavior, small bandgap, and their stability are favorable candidates as photocatalysts. Nowadays, many researchers have focused on the production and photocatalytic activity of TMCs. These TMCs have been synthesized by different techniques that include single-step hydrothermal synthesis [13,36,37], solvothermal method [38], sulfidation and selenization [39], free intercalation method [40], electrospinning technique, the successive ionic layer adsorption and reaction (SILAR) method [32], ultrasonic-exfoliation [23], electrodeposition method [27,41], photoreduction method [42], refluxing method [43], chemical vapor deposition (CVD) [44], template-free one-pot method [45], successive surface modification method [46], and microwave-assisted synthesis [47,48,49,50]. The following section will elaborate on some of the above-mentioned synthesis methods in detail.

2.1. Synthesis Methods

2.1.1. Hydrothermal/Solvothermal Method

The hydrothermal technique is one of the frequently reported methods to produce nanomaterials with various morphologies and thicknesses [51]. This method involves a heterogeneous reaction in the presence of an aqueous solvent or mineralizers under high pressure and low temperature to dissolve and recrystallize the materials that are moderately insoluble under ordinary conditions in closed steel containers called autoclaves. If the water is used as the solvent, this process is called the hydrothermal method whereas in the solvothermal method, organic solvents are used [36,52]. This method can be used to synthesize nanomaterials with high crystallinities, favorable crystal facets, tunable morphologies, and controllable sizes by varying experimental factors such as temperature, solvent, time, etc. Hydrothermal synthesis consumes relatively low energy, and most importantly, it is an environmentally friendly method [53,54], and hence is a very common method. Figure 2 illustrates an example of the synthesis of MoS2.
Figure 2 depicts the flow diagram of the synthesis of MoS2 by the hydrothermal method. Here, ammonium-hepta molybdate tetrahydrate and sulfur (S) powder were used as precursors. The black precipitates were collected and dried to obtain the required MoS2 sample, which exhibited a layered structure [55].
In a different study, Jing Yu et al. synthesized SnS2 nanosheets by the solvothermal method. The mechanism of their synthesis is illustrated in Figure 3, where Sn2+ from SnCl2 reacts with S2− to form orthorhombic SnS micro flowers stacked with decentralized sheets, and further addition of a S2− source results in SnS2 with an increase in the reaction time [38].
In addition, surfactants such as SDS (sodium dodecyl sulfate), CTAB (cetyltrimethylammonium bromide), PVP (polyvinyl pyrrolidone) [56], etc. have been used to increase the crystal growth of the nanomaterials, and a hydrothermal synthesis of a WS2 nanosphere by the addition of CTAB is discussed in Figure 4.
Figure 4a shows that the CTAB molecules were randomly dispersed in solution as single molecules (not aggregates) at low concentrations under the hydrothermal condition. After the amount of CTAB was adjusted to 1.5 mmol, a lamellar micelle was formed to minimize the free energy of the system. The external surface of the lamellar micelle contains a hydrophilic head, whereas the tail, which is hydrophobic, points to the micelle center [58]. In Figure 4b, WS42− ions tend to form lamellar micelles due to complementary electrostatic interactions between the CTA+ and WS42− To keep the free energy of the system low, the lamellar micelle was formed as an ordered nano-microstructure, as indicated in Figure 4c. Figure 4d shows 3-D assemblies formed through van der Waals forces. Similar methods have been utilized to synthesize FeS2, NiS2, and CuS in a different work [37]. The WS2 nanocrystal was formed from the reduction of WS42− ions, as indicated in Figure 4e. The aggregation of WS2 colloids reduced the surface free energy and formed nanosheets. The flower-like WS2 nanospheres were obtained as indicated in Figure 4f, and it was concluded that hydrothermal synthesis in the presence of a surfactant such as CTAB and careful control of the precursor concentration can be utilized to obtain the desired flower-like structure [57].
Apart from this hydrothermal method, there is a clean, ultrafast (60 s) synthetic technique that can be conducted using microwave heating without any inert gas protection to achieve uniform TMCs [48,59].

2.1.2. Microwave-Assisted Synthesis

The microwave-assisted synthesis method has also been used to prepare TMCs with controllable sizes and shapes [60]. Figure 5 illustrates the microwave-assisted synthesis of a MoS2/graphene composite in the presence of ammonium tetrathiomolybdate (ATTM) and CS2 on a graphene support. The reaction was carried out in a glass vial with equimolar amounts of ATTM and graphene. The reactants were mixed at a constant speed, and then the CS2 solvent (200 μL) was added and mixed well. The solvent was evaporated after 10 min of air drying. Then, the mixture was kept under microwave irradiation. Here, the graphene served as a substrate to absorb the microwave energy and convert the incident radiation energy to heat. During the process, ATTM was reduced first to MoO2, and then converted to MoS2, as indicated in Figure 5 [47].
In different works, SnS2 [61], CdS [62], CuS [63], NiS2 [64], WS2 [59], and CoS2 [65] nanomaterials were synthesized using the microwave-assisted synthesis method.

2.1.3. Electrodeposition Method

Electrodeposition is a well-established technique for the preparation of nanomaterials for a wide range of materials (e.g., metal nanoparticles, nanosheet arrays [66], nanowires [67], nanofilms [68], laminated composites [69], multilayered composite coatings [70], and nanoparticle-reinforced composite coatings [71]) and this is a rapid and single-step process. It can be used to produce homogeneous and high-purity crystalline materials at the cathode of the electrochemical system during the coating process because of the external current (direct or pulsed) applied to the electrodeposition system [72,73,74].
Figure 6 illustrates the synthesis of MoS2 quantum dots (QDs) on TiO2 nano-tube arrays (NTAs) to form heterostructures using a facile electrodeposition method with a three-electrode system. TiO2 NTAs/Ti foil, Pt foil, and saturated Ag/AgCl electrodes were used as the working electrode, counter electrode, and reference electrode, respectively. MoS2 QD solution and citric acid–sodium citrate buffer solution (0.1 M, pH 4.5) with a volume ratio of 1:1 was used as the electrolyte of this system. MoS2 QDs were deposited onto TiO2 NTA photoelectrodes by the electrodeposition process, and the obtained samples were rinsed with deionized water and dried at ambient temperature after different electrodeposition times of 5, 10, 20, and 30 min., respectively [27].

2.1.4. Photoreduction

Photoreduction, for example, the photodeposition of MoS2 on rGOx/CdS was also utilized to prepare metal chalcogenide materials. An appropriate amount of precursor molecule (MoS42−) and surfactants were added and the pH of the solution was adjusted by adding an acid or base. Then, the suspension was bubbled with nitrogen to maintain an inert environment, then irradiated with a 350 W Xe lamp equipped with a cutoff filter (λ ≥ 420 nm) for a particular time interval. The incident intensity was 34 mWcm−1. The MoS42− ions were reduced to MoS2 by the photoexcited electrons of rGOx/CdS. The final product, MoS2/rGOx/CdS, was obtained by filtration. The loaded amount of MoS2 was determined by measuring the MoS42− concentrations in both the reacted solution and the washed solution on a spectrophotometer at the wavelength of 468 nm [42].

2.1.5. Electron Beam Evaporation

There have been various methods used in the preparation of Au deposited on a NP/ZnxCd1−xS NW heterostructure. Ternary ZnxCd1−xS nanowires were synthesized by using silicon (100) wafers as substrates (Figure 7). First, Au film was deposited on a Si substrate by the electron beam evaporation method. The substrates coated with 10 nm of Au film were placed into the downstream region of a three-zone horizontal quartz tube furnace and then annealed at 500 °C for 1 h.
The pressure was maintained at 0.01 torr throughout the annealing process. Ternary ZnxCd1−xS nanowires were synthesized by a simple one-step process. Before growth, commercial-grade ZnS powder (99.995%) CdS powder (99.999%), and a piece of Si (100) substrate coated with 10 nm of Au film as a catalyst were annealed. Argon flow was purged during the growth as a carrier gas and the pressure was maintained. After the heating process, the samples were cooled to room temperature in the furnace. The key factor in tuning the x ratio of the ZnxCd1−xS nanowires was found to be temperature [75].

2.1.6. Sulfidation and Selenization

Figure 8a schematically illustrates the experimental setup used for the growth of WO3 nanowires (NWs) as well as WS2, WSe2, and their ternary WS2(1−x)Se2x nanotubes (NTs) on highly flexible carbon fibers (CFs) in a horizontal two-zone tube furnace [39]. High-density WO3 NWs were used to cover the entire surfaces of the CFs vertically and uniformly, as shown in Figure 8c. The diameter and length of typical WO3 NWs were found to be ca. 100 nm and the length was measured to be 5 μm. Then, the WS2 and WSe2 NTs were synthesized through sulfidization and selenization of WO3 NWs at 800 °C, as demonstrated in Figure 8d,e.
For the synthesis of WS2(1−x)Se2x NTs, the ratio of S and Se precursors were successfully altered for the controllable element component of ternary NTs (Figure 8f). As WO3 NWs were used as a conversion template, the diameter and length of the WS2 and WSe2 NTs were found to be comparable to those of WO3 NWs. Synthesis of the WS2 NTs was conducted by the sulfidization of the WO3 NWs performed at 800 °C in an Ar atmosphere. The chemical reaction is described as
2 WO 3 + 7 S     2 WS 2 + 3 SO 2

2.1.7. Successive Ionic Layer Adsorption and Reaction (SILAR) Method

The SILAR method is a simple, cost effective, and rapid technique for the deposition of binary semiconducting thin films when compared with other methods. It can be applied in the deposition of large-area thin films. Since this technique was introduced by Nicolau in the mid-1980s [76], this method has been employed to grow thin films, especially CdS, ZnS, and CdZnS. In particular, this method is appropriate for depositing thin films of chalcogenide groups I–VI, II–VI, III–VI, V–VI, VIII–VI binary, and I–III–VI, II–II–VI, II–III–VI, II–VI–VI and II–V–VI ternary chalcogenides and composite films. The SILAR method consists of the adsorption of the cation precursor, water rinsing, adsorption of the anion precursor, followed by the appropriate reaction, and rinsing. These steps may be repeated. The schematic diagram illustrated in Figure 9 explains the mechanism of the SILAR growth.
In a typical process, the substrate is immersed separately in two precursor solutions and washed with water to remove the loosely bound species. By controlling the experimental conditions, the growth rates of thin films in the SILAR method can be varied between a quarter and a half of a monolayer [78,79,80]. Therefore, the aqua ligands remain at least partially intact during adsorption, thereby lowering the density of cations and anions in one layer. Furthermore, the growth of the thin film can be controlled at an accuracy of one SILAR cycle.
Valkonen et al. [80] studied the effect of different types of substrates by depositing ZnS, CdS films on glass, ITO-coated glass, and single crystals of GaAs (100). In this paper, they mainly focused on promoting interest in the SILAR method as applied to tin disulfide thin films. The authors used the SILAR method to prepare large-area (∼10 cm2) SnS2 thin films on glass substrates under optimized deposition conditions. In addition, they reported SnS2 thin films on a single crystal wafer of Si (111) to study the effect of substrate on the crystallinity of the films. In summary, these reports explain the feasibility of using the SILAR method for the preparation of TMC thin-films/coatings.

2.1.8. Refluxing Method

The refluxing method is a suitable method for the large-scale synthesis of TMCs such as SnS2. In their paper, Liu et al. reported a facile refluxing method without using any templates or surfactants to synthesize SnS2 nanocrystals with controllable morphology. It was found that the as-obtained SnS2 nanostructures exhibited good adsorption ability and enhanced visible light photocatalytic activity for the degradation of Methylene Blue.
In a typical procedure, SnS2 was prepared by using SnCl4·5H2O in the presence of ethylene glycol in a round bottom flask under stirring. The resultant mixture was heated to specific temperature followed by the addition of Na2S solution under vigorous stirring. The resulting solution was refluxed under different temperature (120 °C and 180 °C) to obtain different samples.
In another study, Aditha et al. synthesized Cu2ZnSnS4 using the refluxing method. The experimental setup they used in their study is illustrated in Figure 10. A round bottom flask was fitted with a condenser connected to a water inlet and outlet to condense the vapor back into the reaction vessel to keep the total volume constant. In this method, the reaction temperature is the boiling point of the solvent in the reaction vessel [81].
CuCl2, ZnCl2, SnCl2·2H2O, and thiourea were used as the precursor materials for producing Cu2ZnSnS4 and the refluxing time was 8 h. The reaction mechanism is given below.
2 CuCl 2 +   ZnCl 2 +   SnCl 2 · 2 H 2 O   + 4 SC ( NH 2 ) 2 + 8 H 2 { Cu 2 +   Cu + ,   Sn 2 + Sn 4 + }
2 Cu 2 + +   Zn 2 + +   Sn 2 + + 4 S 2   Cu 2 ZnSnS 4 +   CO 2 + 8 NH 4 Cl
Further to this, several other studies have also been performed by using the refluxing method to synthesize metal chalcogenide materials such as CuS [82], MoS2 [83], ZnS [84], and NiS2 [85].

2.1.9. Chemical Vapor Deposition (CVD) Method

A typical bottom-up method, the chemical vapor deposition (CVD) method, is the process during which gaseous or vapor materials react at a certain temperature in the gas phase or gas–solid interfaces to form solid products with low structural defects [44]. In the CVD method, parameters such as choice of the precursor, temperature, pressure, flow rate, time, the distance between precursor and substrate, and the choice of substrates influences the morphology, layer number, lateral size, orientation, and degree of defects [86,87,88]. Kwon et al. [89] provided a deep understanding of the thermodynamic and kinetic factors influencing the CVD method. Most importantly, the CVD method exhibits prominent advantages in controlling 2-D materials at an atomic-level, which allow for clarity of the structure–activity relationships.
Tan et al. utilized WS2 materials with a graphene electrode for ultrathin 2-D photo detectors [90]. In another study, a 2-D-layered tin disulfide was synthesized by Chang et al. using the CVD method for photo detection of transistors by the thickness-controlled interface dynamics (TCID) technique, which shows high performance and great potential for several other applications [91].
Cong et al. synthesized single crystalline WS2 monolayers by the direct sulfurization of WO3 powders with sulfur powder at 750 °C; 300 nm using SiO2/Si as the substrate. The substrate was covered with the WO3 sample and another substrate was placed face down above it. After that, the setup (Figure 11) was heated in a small end closed quartz tube, which in turn was placed inside a big quartz tube using a tubular furnace [92].
In addition, SnS2 [93], NiS [94], NiS2 [94], MoS2 [95], and FeS2 [96] nanomaterials were also synthesized by using this CVD method in different studies.

2.1.10. Liquid Cascade Centrifugation (LCC) Method

In addition to the above-mentioned methods, synthesis of TMC can also be performed by a liquid cascade centrifugation (LCC) method to achieve a highly efficient nanosheet with selected size and thickness [97]. LCC is an efficient, scalable, and versatile method that is based on a set of iterative centrifugation cascades designed to achieve highly efficient nanosheets and/or monolayer enriched dispersions. Mihai et al. synthesized liquid-exfoliated 2D WS2 and MoS2 nanosheets by the liquid centrifugation cascade method and the prepared nanosheets were used for neuromorphic computing applications [98]. Figure 12 illustrates the LCC method used in their study to separate the nanosheets efficiently according to their size and thickness.
The following Table 1 discusses the different synthesis methods used in the preparation of TMC nanoparticles and the specific advantage(s)/disadvantage(s).

2.2. Characterization Techniques

There have been various techniques such as X-ray diffraction spectroscopy (XRD), scanning electron microscopy (SEM), UV–Visible spectroscopy/diffuse reflectance spectroscopy (UV-DRS), and X-ray photoelectron spectroscopy (XPS) used to characterize the transition metal chalcogenides, which are discussed in the following section.

2.2.1. X-ray Diffraction Method (XRD)

X-ray diffraction method is a non-disruptive technique used to characterize the crystallite size, crystallinity, crystal structure, and phase composition of the material in a powder or film.
Figure 13 shows the XRD patterns of SnS2 nanosheets (NSs) that confirm the formation of 2T-type hexagonal SnS2 NSs (space group P 3 ¯ m 1 ) with cell parameters of a = b = 3.649 Å and c = 5.899 Å. A high crystalline and pure product was obtained when compared with the standard diffraction data of JCPDS no. 23-0677. The highest intensity peak was assigned to the (001) facet of the indicating hexagonal phase [38]. Figure 14 shows the XRD patterns for MoS2 and WS2 on carbon cloth at different thermolysis temperatures. The peaks at 26.8° and 43.5° (marked as “C”) can be ascribed to the carbon cloth in all samples. As shown in Figure 14a, the presence of the (002) reflection peak at 2θ = 14.2° indexed the periodicity along the direction perpendicular to the planar structure of MoS2. This was only observed for the MoS2 samples heated at 1000 °C, but not at lower temperatures. For the WS2 samples, peaks were observed at lower temperatures (i.e., 200 and 400 °C). As the thermolysis temperature increased, higher crystallinity of WS2 was observed. For WS2 samples prepared at 800 and 1000 °C, the reflection peak was found for WO3, which is marked with * in Figure 14b.
Figure 15 shows the XRD patterns indexed to g-C3N4 for different heating times and MoS2 samples. The bottom (green) XRD pattern was indexed to a hexagonal phase of MoS2 (JCPDS 37-1492) [100]. The peaks observed at 2θ = 32.6° and 58.3° corresponded to the (100) and (110) crystal planes of metal chalcogenide, respectively. In addition, the XRD patterns of CN2.5, CN10, and CN20 samples showed similar peaks, as indicated in Figure 15 (JCPDS 87-1526). The peak at 2θ = 27.4° was indexed to the (002) stacking layered structure of g-C3N4 and the 2θ = 13.0° peak corresponded to the (100) in-plane of the repeating units [101]. The CM0.05 heterostructure pattern showed the sole g-C3N4 peaks without the MoS2 peaks due to the low content of MoS2 [102,103,104].
In a different study, pure CoS2, pure TiO2, and 10 wt. % CoS2/TiO2 were successfully synthesized using the hydrothermal method [105]. XRD patterns confirmed the formation of these nanoparticles and the Figure 16 demonstrates the structural quality and phase orientation of the synthesized nanocomposite between the 2θ = 20° and 70° diffraction angles. The XRD confirmed the anatase phase formation for TiO2, whereas the CoS2 phase formation was confirmed by the appearance of peaks at the 2θ degrees of 26.04°, 31.58°, 37.08°, 40.34°, 45.34°, and 55.1° for the (111), (200), (210), (211), (220), and (311) diffraction planes, respectively (JCPDS Card No.: 9007682). Impregnation of CoS2 on TiO2 was confirmed by the combined XRD peaks obtained with the 10 wt. % of CoS2/TiO2 nanocomposite materials.
In another study, SnS2/TiO2 nanocomposites were synthesized by the hydrothermal method, and XRD analysis was carried out to characterize those materials. Figure 17 shows that the peaks at the 2-theta value of 25.29, 37.88, 48.25, 53.74, 55.14, 62.71, 68.86, 70.30, and 75.06 for the pure TiO2 nanoparticles were due to the (101), (004), (200), (105), (211), (204), (116), (220), and (215) diffraction planes of the anatase phase of TiO2, which was well matched with the JCPDS:00-004-0477. The observed peaks at 2θ values of 14.76, 27.97, 32.00, 41.79, 49.92, 52.41, 54.79, and 60.55 for pure SnS2 was correlated to the (001), (100), (101), (102), (110), (111), (103), and (201) diffraction planes (JCPDS:00-023-0677).
Combined peaks of both SnS2 and TiO2 were observed with the 5, 10, 15, and 20 wt. % SnS2/TiO2 (ST-5, ST-10, ST-15, and ST-20, respectively) nanocomposite and confirmed the good impregnation of SnS2 on TiO2 in their study [106].

2.2.2. Scanning Electron Microscopy (SEM)

Scanning electron microscopy (SEM) is an established technique used to analyze the morphology of the sample, especially given the simplicity of the sample preparation. Estimates of the sample composition can be obtained through energy dispersive X-ray spectroscopy (EDS).
In addition, the particle size distribution can also be easily obtained with relative ease, and thus, many of the reported works have used this simple characterization technique to study the TMC materials using SEM.
Figure 18 exhibits the phase and morphology transformation of intermediates during the formation of monodispersed SnS2 nanosheets; consequently, the effects of the reaction parameters including reaction time and the dosage of polyvinylpyrrolidone (PVP) on the morphology of SnS2 was also studied. PVP was used as a steric stabilizer or capping agent to minimize aggregation of the colloidal particles.
Figure 18a shows the structure of orthorhombic SnS microspheres stacked by sheets, which were first nucleated in the primary stage of a chemical reaction. When the reaction time increased to 2 h, small sheets expanded and grew onto larger sheets. At the same time, in Figure 18b, the spherical flower-like morphology remained and the regular nanosheets were deposited around micro-flowers. Figure 18c is the high-magnification SEM image with the reaction time of 2 h, which exhibits the deposited nanosheets. When the reaction time was prolonged to 5 h, uniform nanosheets were obtained with the single phase of hexagonal SnS2. Consequently, it was proven that the SnS phase would be oxidized into SnS2 gradually as the reaction progresses. Furthermore, as the reaction time increases, SnS2 nanosheets tend to grow and enhance their crystallinity [38], which was proven by SEM analysis.
In another study, a different TMC, NiS2 particles with a hollow structure, were synthesized and characterized by FESEM. Figure 19a,b indicates the FESEM images of NiS2 hollow microspheres prepared on a large scale. Figure 19b shows that the diameter of NiS2 hollow microspheres was about 3 μm and it was found that the surface of the microspheres was rough. The thickness of the wall was found to be approximately 200 nm (Figure 19c), and the hollow structure can be clearly seen in this image. Figure 20d shows a high-resolution TEM image that displays the lattice fringes of NiS2. The lattice distance was 0.286 nm, which corresponds to the (200) crystal plane [13].
In a different study, TMC doped TiO2 materials were synthesized and studied for hydrogen production. Sivagowri et al. synthesized a CoS2/TiO2 nanocomposite using hydrothermal methods. As shown in the SEM images of pure CoS2 (Figure 20a,b), aggregates stacked by flake-like structures were attained. The image for pure TiO2 (Figure 20c,d) showed an irregular 3-D block-like structure covered with spongy like particles, whereas CoS2/TiO2 indicates a hexagonal rod like structure decorated with sponge like materials, as illustrated in Figure 20e,f [105].
The same group, in a different study, reported that a different morphology was attained with SnS2–TiO2 nanocomposites when only the TMC was changed. In the SEM study, they observed that the pure TiO2 nanoparticle showed an irregular 3-D block-like structure covered with sponge-like particles as indicated in Figure 21a. The micrograph obtained in the higher magnification (Figure 21b) image indicates that the aggregated particles had spongy-like structures. A flower-like structure composed with nanosheets was attained for pristine SnS2 and is shown in Figure 21c,d.
The SEM images of different wt. % of SnS2 that included 5% (Figure 22e), 10% (Figure 22f), 15% (Figure 22g), and 20% (Figure 22h) embedded TiO2 nanocomposites are shown in Figure 22. The surface morphology of these SnS2/TiO2 nanocomposites at different wt. % possessed nanosheets, and it can be seen that the sponge-like materials covered the structure and, therefore, exhibited significant aggregation [106].

2.2.3. X-ray Photoelectron Spectroscopy (XPS) Studies

X-ray photoelectron spectroscopy (XPS) is based on the photoelectric effect and was used to investigate the chemical state and surface elemental composition of the samples. Among the literature collected, it was found that only a few reports have presented the XPS analysis of TMC material, which may be due to the necessity of the analysis for the structure–activity correlation. In a study by Bo Gao et al., a MoS2/candle soot composite was synthesized and analyzed using XPS. Figure 23 shows the existence of the elements C, Mo, and S with a small amount of O. The corresponding peaks indicate the binding energies at 284.6 (C 1s), 161.5 (S 2p), 231.0 (Mo 3d), and 529.1 eV (O 1s), respectively. In addition, the atomic ratio between Mo and S was found to be 0.43. Figure 23b illustrates the presence of peaks with binding energies of 232.8 and 229.7 eV, which may be ascribed to Mo 3d3/2 and 3d5/2 of 2H-MoS2 at the peaks at 228.9 eV for 3d5/2, and at 231.9 eV for 3d3/2, which was ascribed to 1T-MoS2 [107]. The peak located at 226.9 eV corresponded to the binding energy of S 2s. In addition, the two weak peaks at 233.6 and 236.1 eV were due to the presence of the MoO3 that formed during the sputtering process. From these results, they found that Mo4+ dominates the product. Similarly, the two doublet peaks located at 162.6 eV and 164.1 eV corresponded to the 2p3/2 and 2p1/2 orbitals of 2H-MoS2. Furthermore, the binding energies at 161.6 eV and 163.2 eV were due to 1TMoS2. The high-resolution spectra of XPS further provides evidence for the existence of MoS2 in the hybridized samples [108].
In a different study, Li and coworkers used XPS characterization to study the Pt doped CdS materials and Figure 24a illustrates the XPS spectrum of the Pt–CdS, indicating the co-existence of Cd, S, and Pt elements in the as-prepared Pt–CdS composite. The peaks at 405.6 and 412.3 eV corresponded to the Cd 3d5/2 and Cd 3d3/2 of CdS, respectively (Figure 24b) [25]. In Figure 24c, the S 2p1/2 and S 2p3/2 peaks are located at 163.2 and 161.9 eV, respectively.

2.2.4. UV–Visible Spectroscopy/Diffuse Reflectance Spectroscopy (DRS)

Like all nanomaterials, to study the bandgap and the optical behavior of the TMCs, researchers have utilized UV–Visible diffuse reflectance spectroscopy. In this regard, in the study of the facile hydrothermal synthesis of a 3-D flower-like La-MoS2 nanostructure for photocatalytic hydrogen energy production, Riaz et al. utilized this technique to study the bandgap of La doped MoS2 materials, as illustrated in Figure 25.
The absorption spectra of pure MoS2 and 3% La-MoS2 indicate that the absorption of LaMoS2 was significantly extended toward a longer wavelength than pure MoS2 into the visible light range. Bandgap energy for pure MoS2 calculated by using the equation (αhυ)2 = A(hυ − Eg) allowed for direct transitions, resulting in a value of 1.84 eV, which is illustrated in the plot of (αhν)2 vs. hν (eV) (Figure 25). This value is close to the bandgap energy of 1.8 eV for a monolayer of MoS2 [110]. However, 3% La-MoS2 showed a considerably small bandgap energy of 1.5 eV due to La3+ doping, which significantly improved the light absorption properties of MoS2 [109].
Gaur et al. used SnS2 nanoparticles as photocatalysts for the reduction of Cr(VI) and their optical properties were investigated by DRS studies. Figure 26a,b shows DRS and Tauc plots of SnS2 nanoparticles. Bandgap absorption in the range between 430 nm and 580 nm was observed. The bandgap calculated by using the Tauc plot indicates values between 2.31 eV and 2.94 eV [111].
In addition to these methods, Raman spectroscopic technique and photoluminescence studies (PL) were also utilized to analyze these TMC materials. However, there is a limited number of literature available, especially with the PL and AFM studies on the TMC materials toward hydrogen production application, and thus, were omitted in this review.

2.2.5. Raman Spectroscopy

L. Zhu et al. synthesized CoS2-based thin films as model catalysts for the oxygen reduction reaction, which were characterized by using Raman spectroscopy. Figure 27 shows characteristic peaks at 287 and 389 cm−1 as well as a shoulder around 410 cm−1 for the CoS2 powder. This was in close agreement with the study of a CoS2 single crystal [112]. The peak at 287 cm−1 was also associated with polysulfides. In contrast, the Raman spectrum obtained from the magnetron-sputtered thin film of CoS2 showed a pronounced peak at 391 cm−1, a broad shoulder at around 410 cm−1, and another peak at 287 cm−1, which exhibited a general similarity to the spectrum obtained for the standard CoS2. Furthermore, the Raman spectrum of commercial NiS2 was also reported in their study and it showed distinctive peaks at 282 and 477 cm−1, and a shoulder at 488 cm−1 (Figure 27). From this study, they confirmed the formation of CoS2 with the expected characteristics.

3. Hydrogen Evolution Reactions

Production of hydrogen through water splitting is seen as an extremely important renewable and clean energy [4]. Since water is abundant, and as novel materials are synthesized and the technology of water splitting becomes mature, hydrogen can be produced from water by different methods such as thermolysis (e.g., thermochemical cycles), electrolysis (e.g., electrocatalysis), photolysis (e.g., photocatalysis), and biolysis (e.g., dark fermentation) [113].
Among them, electrolysis of water, which includes the cathodic hydrogen evolution reaction (HER) [113] and anodic oxygen evolution reaction (OER) [113], remains limited by the high cost of noble metals such as Pt [114] and Ru [115] based materials, which are typically used as electrocatalysts in different electrolyte materials that include acid and base electrolytes [116]. In addition, different photocatalytic materials are being used to utilize solar energy for splitting water [37,105,117]. Photoelectrochemical and photocatalytic are the two main methods for water splitting. In the photoelectrochemical method, a semiconductor photoelectrode is typically used as the working electrode and is immersed in an aqueous electrolyte. A counter and reference electrode are used to complete the cell. An electrochemical potential is then applied to the photoelectrode. A light source is used to excite the semiconductor electrode. If an n-type semiconductor is used, the photoexcited electrons generated in the photoanode travel to the counter electrode (via an external wire) and reduce protons to form hydrogen. At the working electrode, the holes oxidize water to produce oxygen and an anodic photocurrent is produced. Thus, the products, hydrogen and oxygen, are produced at separate electrodes.
In the photocatalytic method, the semiconductor powder is suspended in water. The slurry is irradiated by an appropriate light source. Bandgap excitation leads to the generation of electrons and holes. These charge carriers migrate to the surface and dissipate their charges and produce hydrogen and oxygen, respectively. In photocatalysis, the product separation is required. Nowadays, researchers are putting great effort into maximizing the production of hydrogen by water splitting with the eventual aim to generate hydrogen through sustainable ways [13,118,119,120].
The following table (Table 2) lists the different hydrogen production methods and their advantages and disadvantages.
In recent years, the structure of the synthesized catalyst and the hydrogen evolution activities have been investigated theoretically using density functional theory (DFT) [129]. In this regard, theoretical studies of metal chalcogenide materials have been conducted. In a recent paper, Som et al. investigated the structural property of ZrS2, Janus ZrSSe, and ZrSe2 and their catalytic activity from hydrogen evolution reactions [130]. From these calculations, they found that ZrS2 possessed a larger bandgap energy of 1.58 eV, whereas ZrSSe and ZrSe2 exhibited bandgap energies of 1.08 eV and 0.84 eV, respectively. In addition, Janus ZrSSe (DG ¼ 1:19 eV; DG: Gibbs free energy) was found to be the optimal catalyst [130] among the materials studied. Another study examined the electronic and optical properties and the potential applications of 2-D MoSe2/SnS2 van der Waals heterojunction as a photocatalyst for overall water splitting. The DFT calculations showed that the MoSe2/SnS2 heterostructure, as a direct Z-scheme photocatalyst, effectively absorbs visible light. The separation efficiency of photo-induced carriers was enhanced due to the induced built-in electric field. Moreover, when compared with the MoSe2 and SnS2 monolayers, the absorption intensity of the MoSe2/SnS2 heterojunction was reinforced in the visible light region [131].
Park et al. worked on Nb-, Ta-, and Bi-based oxides/chalcogenides. They studied the crystal structure, band dispersion, effective masses of the charge carriers, band alignment, and pH dependence of these materials for hydrogen production and CO2 reduction. They concluded that Bi-based chalcogenide materials, (KBiO2 and KBiS2) showed good performance over a wide range of pH conditions. In addition, the incorporation of chalcogenide narrows the bandgap and reduces the effective mass of the charge carriers when compared with their oxide counterparts [132]. Although these theoretical studies provide guidance toward the identification of materials required for efficient water splitting, it is necessary to test materials. The following sections summarize the reports of hydrogen evolution.

3.1. Photoelectrochemical Hydrogen Evolution

Electrochemical water splitting is a sustainable method for producing hydrogen on a large scale [133]. In photoelectrochemical hydrogen evolution, semiconductor material(s) are immersed in a water-based electrolyte and an external electric power/voltage is applied to dissociate water as hydrogen and oxygen molecules in the presence of a light source [126]. Electrochemical and photoelectrochemical hydrogen evolution have received much attention due to the fact that (i) the generation of O2 and H2 occurs at separate electrodes (Figure 28), which eliminates the separation issue(s), and (ii) the potential for operation under ambient conditions [134].
Fujishima and Honda first described the photoelectrochemical cell (PEC) splitting of water using a crystalline titanium dioxide anode and a platinum cathode under ultraviolet (UV) irradiation and an external bias [24,135]. The anodic and cathodic reactions are as indicated below [136].
Anode: 2   H 2 O   O 2 + 4   H + + 4   e
Cathode: 2   H 2 O + 4   H + + 4   e   2   H 2 + 2   O 2
Overall reaction: 2   H 2 O   2   H 2 +   O 2  
Different studies have been conducted in the field of electrochemical water splitting wherein Pt- and Ru-based materials [137] were used as electrocatalysts for the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER), respectively. In recent years, TMCs have been used. Ahn et al. used CoS2/rGO (2.3 mg/mL GO) hybrid materials that exhibited high electrocatalytic activity for hydrogen evolution reaction with an over potential of −150 mV versus RHE [138]. In another study, an impressive activity was attributed to the superior electrochemical performance of the Mo–CoSx skeleton, and the special configuration between MoS2 and CoSx, which provide more active sites [139]. Apart from these studies, there have been several metal chalcogenide materials that have been utilized for electrochemical hydrogen evolution, which are summarized in the Table 3 below.

3.2. Photocatalytic Hydrogen Evolution

Solar-assisted photocatalytic hydrogen evolution is an attractive technique to generate hydrogen and oxygen, because water and sunlight are abundant and renewable. Furthermore, the photocatalytic process can occur at ambient conditions [14]. Semiconductor materials are used as photocatalysts and their band structure consists of a filled valance band and an empty conduction band separated by a finite bandgap. Irradiation with incident energy greater than or equal to the bandgap of a material leads to excitation of electrons from the valence band to the conduction band edge [159,160,161]. This leaves holes in the valence band. These photoexcited electrons and holes are potent reducing and oxidizing agents and may react with adsorbed moieties on the surface of the photocatalytic material or may recombine to generate heat [9,136].
Photocatalytic reactions can be classified into two major categories that include uphill reaction, where the Gibbs free energy ( Δ G o ) is positive, or downhill reaction, where the ( Δ G o ) is negative (Figure 29). In this process, water molecules splitting into H2 and O2, are accompanied by a large positive change in the Gibbs free energy (i.e., an uphill reaction) [130]. During this process, light energy is transformed into chemical energy in a similar manner as the photosynthesis process. Therefore, this reaction is called artificial photosynthesis [136,162].
The photocatalytic hydrogen evolution through water splitting is a green route and has received significant attention toward simultaneously solving the current global environmental pollution and energy crisis [128]. Various photocatalysts have been well developed for the generation of hydrogen whereas TiO2 is still considered as a golden standard due to its exceptional stability under irradiation conditions, ease of synthesis, and cost [163]. However, a big disadvantage of TiO2 is that it requires UV light to be activated, thus limiting its utility since visible light cannot be used. Hence, other oxide and chalcogenide materials are constantly being explored and evaluated for photocatalytic hydrogen production. In addition, ZnO [33], mixed oxides [164], metal chalcogenides such as CdS [165], WS2 [166], g-C3N4 [30], and CdLa2S4 [34] have also been utilized for the hydrogen evolution reaction.

3.2.1. Principle and Mechanism of Photocatalytic Hydrogen Evolution

In general, photocatalytic hydrogen evolution reaction takes place through three major steps. The first step is the absorption of light by a catalyst to create exciton pair containing electrons and holes. In the second step, the excitons may separate and migrate to the surface of the photocatalyst, and finally, in the third step, the surface reactions take place. The electrons will reduce the protons to generate hydrogen molecules, and holes oxidize water molecules to produce oxygen, as illustrated in the Figure 30 below [167].
The primary reactions (1–4) occurring during the process are listed below, and as indicated, the overall reaction involves four electrons. Table 4 lists the different metal chalcogenide materials used for hydrogen production and the amount of hydrogen evolved.
Photocatalyst + 4hν → Photocatalyst + 4h+ + 4e
4e + 4H+ → 2H + 2H → 2H2
4h+ + 4H2O → 4OH + 4H+
2OH + 2OH → 2H2O + O2

3.2.2. Measurements of the Rate of Hydrogen Evolution (Quantum Yield)

Generally, the rate of hydrogen evolution is denoted with the unit of µmolh−1g−1, which indicates the number of moles of hydrogen evolved in unit time using a unit amount of catalyst. This amount depends on several experimental conditions such as reaction vessel, light source, intensity of radiation, amount of catalyst, sacrificial agent, etc. [120]. However, it is noteworthy to mention here that there are inconsistencies among researchers in reporting the rate of hydrogen evolution and hence care must be exercised in making comparisons from one report to another. To provide a general quantification, quantum yield measurement is important in photolytic reactions to remove any variations in experimental conditions and for making meaningful comparisons among different photocatalysts and provide consistent benchmarking. Although the number of incident photons can be measured using a thermopile, it is difficult to determine the actual number of photons absorbed by a photocatalyst because of scattering and some reflection in a dispersed system. Photocatalytic/luminescence materials convert excitation energy into photons, and this conversion is usually not 100% efficient due to various losses through alternative deactivation paths. Therefore, two additional parameters such as the quantum yield (QY) and the quantum (energy) efficiency [191] have to be defined. In preparative photochemistry, the QY is defined as the product yield of a reaction calculated by the number of moles of reactant reacted per unit time. This type of QY has been preferred as it can be measured with small expenditure. However, it depends on the time section of the reaction in which it is determined. Therefore, the QY can be considered to be an apparent quantum yield (AQY), and this yield is mainly dependent on the time at which they are considered [192]. AQY is the ratio between the number of reacted electrons and the number of incident photons, and is used for the quantification of hydrogen evolution. The following Equation (5) explains the percentage of AQY calculation [9].
AQY ( % )   =   the   number   of   reacted   electrons the   number   of   incident   photons   × 100 %
= the   number   of   evolved   H 2   molecules   ×   2   the   number   of   incident   photons × 100 %
The turnover frequency (TOF) of photocatalysts based on the amount of a particular molecule can be calculated using Equation (6) [193]:
TOF = moles   of   produced   hydrogen moles   of   a   particular   molecule   in   photocatalyst   × reaction   time   ( h )
Mangiri R. et al. used a dual co-catalyst of MoS2/CoSe2 on CdS for the hydrogen evolution reaction (Figure 31), and the system evolved a higher rate of hydrogen with the yield of 191.5 mmol g−1 h−1, which was 76.6 times higher than that of pristine CdS [182].
In their study, CdS was used as a semiconductor, which can harvest the light and eject electrons from the conduction band of CdS to its valance band. Then, the photo-generated electrons in the conduction band of CdS can be easily transferred to MoS2 and CoSe2 due to the favorable band edge positions of the three materials. Thus, the rate of charge-carrier recombination is minimized and photocatalytic activity is enhanced in the composite photocatalyst in comparison to the pristine CdS materials. Here, the MoS2 and CoSe2 act as reactive sites to produce hydrogen. The holes produced can react with lactic acid and oxidize it to pyruvic acid [182].
In a different study, Riaz et al. used 3% La3+ doped MoS2, and it exhibited an enhanced photocatalytic performance for hydrogen evolution with a 5.2 times higher hydrogen evolution rate compared with pure MoS2 [109] (Figure 32). This has been attributed to the presence of trapping energy levels formed below the conduction band of MoS2 due to the lanthanum ion-doping and minimizing recombination of electrons and holes. Surface adsorbed La3+ atoms also ensured the rapid transfer of electrons from MoS2 to capture protons for H2 production [109].
Transition metal chalcogenides with and without incorporation of other materials such as metals [168], graphene [172], graphene oxide [42], carbon nitrides [169], dyes [194], and metal oxides [33] have been utilized for hydrogen production (Table 4). It can be seen that the use of composite photocatalysts, in general, leads to a higher rate of hydrogen production than the pristine transition metal chalcogenides. This has been attributed to the presence of heterojunctions that favor vectorial transfer of electrons that minimize electron-hole recombination.

4. Summary

This review details the recent approaches to the hydrogen evolution reaction using TMCs, which are being considered as potential catalysts for the production of an energy carrier, hydrogen, due to their favorable electronic and optical properties. The different synthesis methods of TMCs and their characteristic properties reported in the literature are discussed in detail, providing guidance for interested researchers to explore TMCs as exploratory materials. Furthermore, the catalytic activities of TMCs for hydrogen evolution reactions were also elaborated herein. Overall, this review provides an overview of the recent approaches to hydrogen production using TMCs.

Author Contributions

Conceptualization, Y.S.; Methodology, S.S. and E.Y.; Investigation, S.S., Y.S., P.R., D.V. and R.K.; Resources, Y.S., P.R., D.V. and R.K.; Original draft preparation, S.S. and E.Y.; Writing—review & editing, S.S., E.Y., Y.S. and R.K.; Supervision, Y.S., P.R. and D.V. All authors have read and agreed to the published version of the manuscript.

Funding

National Research Council, Sri Lanka (NRC Grant No: 20-110) and the Capacity Building and Establishment of a Research Consortium (CBERC) project, grant number LKA-3182-HRNCET and Higher Education and Research Collaboration on Nanomaterials for Clean Energy Technologies (HRNCET) project, grant number NORPART/2016/10237.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors appreciate the support of the National Research Council, Sri Lanka (NRC, Grant No: 20-110) for funding the research.

Conflicts of Interest

There is no conflict of interest.

References

  1. Ma, Q.; Hu, C.; Liu, K.; Hung, S.-F.; Ou, D.; Chen, H.M.; Fu, G.; Zheng, N. Identifying the electrocatalytic sites of nickel disulfide in alkaline hydrogen evolution reaction. Nano Energy 2017, 41, 148–153. [Google Scholar] [CrossRef]
  2. Bai, J.; Meng, T.; Guo, D.; Wang, S.; Mao, B.; Cao, M. Co9S8@MoS2 Core–Shell Heterostructures as Trifunctional Electrocatalysts for Overall Water Splitting and Zn–Air Batteries. ACS Appl. Mater. Interfaces 2018, 10, 1678–1689. [Google Scholar] [CrossRef] [PubMed]
  3. Naik, S.G.; Rabinal, M.H.K. Molybdenum Disulphide Heterointerfaces as Potential Materials for Solar Cells, Energy Storage, and Hydrogen Evolution. Energy Technol. 2020, 8. [Google Scholar] [CrossRef]
  4. Feng, K.; Huang, D.; Li, L.; Wang, K.; Li, J.; Harada, T.; Ikeda, S.; Jiang, F. MoSx-CdS/Cu2ZnSnS4-based thin film photocathode for solar hydrogen evolution from water. Appl. Catal. Environ. 2020, 268, 118438. [Google Scholar] [CrossRef]
  5. Züttel, A. Hydrogen storage methods. Naturwissenschaften 2004, 91, 157–172. [Google Scholar] [CrossRef] [PubMed]
  6. Züttel, A. Materials for hydrogen storage. Mater. Today 2003, 6, 24–33. [Google Scholar] [CrossRef]
  7. Paquin, F.; Rivnay, J.; Salleo, A.; Stingelin, N.; Silva, C. Multi-phase semicrystalline microstructures drive exciton dissociation in neat plastic semiconductors. J. Mater. Chem. 2015, 3, 10715–10722. [Google Scholar] [CrossRef]
  8. Lin, L.; Sherrell, P.; Liu, Y.; Lei, W.; Zhang, S.; Zhang, H.; Wallace, G.G.; Chen, J. Engineered 2D Transition Metal Dichalcogenides—A Vision of Viable Hydrogen Evolution Reaction Catalysis. Adv. Energy Mater. 2020, 10. [Google Scholar] [CrossRef]
  9. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  10. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  11. Dincer, I. Green methods for hydrogen production. Int. J. Hydrogen Energy 2012, 37, 1954–1971. [Google Scholar] [CrossRef]
  12. Ma, L.-C.; Dominguez, B.C.; Kazantzis, N.K.; Ma, Y.H. Integration of membrane technology into hydrogen production plants with CO2 capture: An economic performance assessment study. Int. J. Greenh. Gas Control. 2015, 42, 424–438. [Google Scholar] [CrossRef] [Green Version]
  13. Luo, P.; Zhang, H.; Liu, L.; Zhang, Y.; Deng, J.; Xu, C.; Hu, N.; Wang, Y. Targeted Synthesis of Unique Nickel Sulfide (NiS, NiS2) Microarchitectures and the Applications for the Enhanced Water Splitting System. ACS Appl. Mater. Interfaces 2017, 9, 2500–2508. [Google Scholar] [CrossRef] [PubMed]
  14. Leung, D.Y.C.; Fu, X.; Wang, C.; Ni, M.; Leung, M.K.H.; Wang, X.; Fu, X. Hydrogen Production over Titania-Based Photocatalysts. ChemSusChem 2010, 3, 681–694. [Google Scholar] [CrossRef] [PubMed]
  15. Yin, Y.; Jin, Z.; Hou, F. Enhanced solar water-splitting efficiency using core/sheath heterostructure CdS/TiO2 nanotube arrays. Nanotechnology 2007, 18, 495608. [Google Scholar] [CrossRef]
  16. Kim, S.B.; Hong, S.C. Kinetic study for photocatalytic degradation of volatile organic compounds in air using thin film TiO2 photocatalyst. Appl. Catal. Environ. 2002, 35, 305–315. [Google Scholar] [CrossRef]
  17. Matthews, R. Photooxidative degradation of coloured organics in water using supported catalysts. TiO2 on sand. Water Res. 1991, 25, 1169–1176. [Google Scholar] [CrossRef]
  18. Chakrabarti, S.; Dutta, B.K. Photocatalytic degradation of model textile dyes in wastewater using ZnO as semiconductor catalyst. J. Hazard. Mater. 2004, 112, 269–278. [Google Scholar] [CrossRef]
  19. Mugunthan, E.; Saidutta, M.B.; JagadeeshBabu, P.E. Photocatalytic degradation of diclofenac using TiO2–SnO2 mixed oxide catalysts. Environ. Technol. 2019, 40, 929–941. [Google Scholar] [CrossRef] [PubMed]
  20. Kim, H.G.; Hwang, D.W.; Kim, J.; Kim, Y.G.; Lee, J.S. Highly donor-doped (110) layered perovskite materials as novel photocatalysts for overall water splitting. Chem. Commun. 1999, 2, 1077–1078. [Google Scholar] [CrossRef] [Green Version]
  21. Kanazawa, T.; Nozawa, S.; Lu, D.; Maeda, K. Structure and Photocatalytic Activity of PdCrOx Cocatalyst on SrTiO3 for Overall Water Splitting. Catalysts 2019, 9, 59. [Google Scholar] [CrossRef] [Green Version]
  22. Kong, C.; Min, S.; Lu, G. Dye-Sensitized NiSx Catalyst Decorated on Graphene for Highly Efficient Reduction of Water to Hydrogen under Visible Light Irradiation. ACS Catal. 2014, 4, 2763–2769. [Google Scholar] [CrossRef]
  23. Zhang, W.; Xiao, X.; Zeng, X.; Li, Y.; Zheng, L.; Wan, C. Enhanced photocatalytic activity of TiO2 nanoparticles using SnS2/RGO hybrid as co-catalyst: DFT study and photocatalytic mechanism. J. Alloys Compd. 2016, 685, 774–783. [Google Scholar] [CrossRef]
  24. Jafari, T.; Moharreri, E.; Amin, A.S.; Miao, R.; Song, W.; Suib, S.L. Photocatalytic Water Splitting—The Untamed Dream: A Review of Recent Advances. Molecules 2016, 21, 900. [Google Scholar] [CrossRef]
  25. Li, Y.; Zhou, X.; Xing, Y. In situ thermal-assisted loading of monodispersed Pt nanoclusters on CdS nanoflowers for efficient photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 506, 144933. [Google Scholar] [CrossRef]
  26. Zheng, L.; Zhang, W.; Xiao, X. Preparation of titanium dioxide/tungsten disulfide composite photocatalysts with enhanced photocatalytic activity under visible light. Korean J. Chem. Eng. 2015, 33, 107–113. [Google Scholar] [CrossRef]
  27. Wang, Q.; Huang, J.; Sun, H.; Ng, Y.H.; Zhang, K.-Q.; Lai, Y. MoS2 Quantum Dots@TiO2 Nanotube Arrays: An Extended-Spectrum-Driven Photocatalyst for Solar Hydrogen Evolution. ChemSusChem 2018, 11, 1708–1721. [Google Scholar] [CrossRef]
  28. Ali, A.; Oh, W.-C. Photocatalytic Performance of CoS2-Graphene-TiO2 Ternary Composites for Reactive Black B (RBB) Degradation. J. Korean Ceram. Soc. 2017, 54, 308–313. [Google Scholar] [CrossRef] [Green Version]
  29. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 Nanoparticles Grown on Graphene: An Advanced Catalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef] [Green Version]
  30. He, K.; Xie, J.; Li, M.; Li, X. In situ one-pot fabrication of g-C3N4 nanosheets/NiS cocatalyst heterojunction with intimate interfaces for efficient visible light photocatalytic H2 generation. Appl. Surf. Sci. 2018, 430, 208–217. [Google Scholar] [CrossRef]
  31. Jing, D.; Guo, L. WS2 sensitized mesoporous TiO2 for efficient photocatalytic hydrogen production from water under visible light irradiation. Catal. Commun. 2007, 8, 795–799. [Google Scholar] [CrossRef]
  32. Wang, Y.; Yang, X.; Wang, Z.; Lv, X.; Jia, H.; Kong, J.; Yu, M. CdS and SnS2 nanoparticles co-sensitized TiO2 nanotube arrays and the enhanced photocatalytic property. J. Photochem. Photobiol. Chem. 2016, 325, 55–61. [Google Scholar] [CrossRef]
  33. Ma, D.; Shi, J.-W.; Zou, Y.; Fan, Z.; Ji, X.; Niu, C. Highly Efficient Photocatalyst Based on a CdS Quantum Dots/ZnO Nanosheets 0D/2D Heterojunction for Hydrogen Evolution from Water Splitting. ACS Appl. Mater. Interfaces 2017, 9, 25377–25386. [Google Scholar] [CrossRef] [PubMed]
  34. Yuan, Y.-P.; Cao, S.-W.; Yin, L.-S.; Xu, L.; Xue, C. NiS2 Co-catalyst decoration on CdLa2S4 nanocrystals for efficient photocatalytic hydrogen generation under visible light irradiation. Int. J. Hydrogen Energy 2013, 38, 7218–7223. [Google Scholar] [CrossRef]
  35. Sun, B.; Liu, A.; Li, J.; Wang, J.; Wang, S. Development of novel highly stable synergistic quaternary photocatalyst for the efficient hydrogen evolution reaction. Appl. Surf. Sci. 2020, 510, 145498. [Google Scholar] [CrossRef]
  36. Makwana, N.M.; Tighe, C.J.; Gruar, R.I.; McMillan, P.F.; Darr, J.A. Pilot plant scale continuous hydrothermal synthesis of nano-titania; effect of size on photocatalytic activity. Mater. Sci. Semicond. Process. 2016, 42, 131–137. [Google Scholar] [CrossRef]
  37. Huerta-Flores, A.M.; Torres-Martínez, L.M.; Moctezuma, E.; Singh, A.P.; Wickman, B. Green synthesis of earth-abundant metal sulfides (FeS2, CuS, and NiS2) and their use as visible-light active photocatalysts for H2 generation and dye removal. J. Mater. Sci. Mater. Electron. 2018, 29, 11613–11626. [Google Scholar] [CrossRef]
  38. Yu, J.; Xu, C.; Ma, F.; Hu, S.-P.; Zhang, Y.-W.; Zhen, L. Monodisperse SnS2 Nanosheets for High-Performance Photocatalytic Hydrogen Generation. ACS Appl. Mater. Interfaces 2014, 6, 22370–22377. [Google Scholar] [CrossRef]
  39. Xu, K.; Wang, F.; Wang, Z.; Zhan, X.; Wang, Q.; Cheng, Z.; Safdar, M.; He, J. Component-Controllable WS2(1–x)Se2x Nanotubes for Efficient Hydrogen Evolution Reaction. ACS Nano 2014, 8, 8468–8476. [Google Scholar] [CrossRef]
  40. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.; Asefa, T.; Shenoy, V.B.; Eda, G.; Chhowalla, M. Conducting MoS2 Nanosheets as Catalysts for Hydrogen Evolution Reaction. Nano Lett. 2013, 13, 6222–6227. [Google Scholar] [CrossRef] [PubMed]
  41. Sun, Y.; Liu, C.; Grauer, D.C.; Yano, J.; Long, J.R.; Yang, P.; Chang, C.J. Electrodeposited Cobalt-Sulfide Catalyst for Electrochemical and Photoelectrochemical Hydrogen Generation from Water. J. Am. Chem. Soc. 2013, 135, 17699–17702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Li, Y.; Wang, H.; Peng, S. Tunable Photodeposition of MoS2 onto a Composite of Reduced Graphene Oxide and CdS for Synergic Photocatalytic Hydrogen Generation. J. Phys. Chem. C 2014, 118, 19842–19848. [Google Scholar] [CrossRef]
  43. Liu, J.-H.; Huang, G.-F.; Huang, W.-Q.; Miao, H.; Zhou, B.-X. Morphology-controlled SnS2 nanostructures synthesized by refluxing method with high photocatalytic activity. Mater. Lett. 2015, 161, 480–483. [Google Scholar] [CrossRef]
  44. Wang, Q.; Lei, Y.; Wang, Y.; Liu, Y.; Song, C.; Zeng, J.; Song, Y.; Duan, X.; Wang, D.; Li, Y. Atomic-scale engineering of chemical-vapor-deposition-grown 2D transition metal dichalcogenides for electrocatalysis. Energy Environ. Sci. 2020, 13, 1593–1616. [Google Scholar] [CrossRef]
  45. Luo, M.; Liu, Y.; Hu, J.; Liu, H.; Li, J. One-Pot Synthesis of CdS and Ni-Doped CdS Hollow Spheres with Enhanced Photocatalytic Activity and Durability. ACS Appl. Mater. Interfaces 2012, 4, 1813–1821. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, Y.; Li, L.; Yao, W.; Song, S.; Sun, J.T.; Pan, J.; Ren, X.; Li, C.; Okunishi, E.; Wang, Y.-Q.; et al. Monolayer PtSe2, a New Semiconducting Transition-Metal-Dichalcogenide, Epitaxially Grown by Direct Selenization of Pt. Nano Lett. 2015, 15, 4013–4018. [Google Scholar] [CrossRef] [PubMed]
  47. Sarwar, S.; Nautiyal, A.; Cook, J.; Yuan, Y.; Li, J.; Uprety, S.; Shahbazian-Yassar, R.; Wang, R.; Park, M.; Bozack, M.J.; et al. Facile microwave approach towards high performance MoS2/graphene nanocomposite for hydrogen evolution reaction. Sci. China Mater. 2020, 63, 62–74. [Google Scholar] [CrossRef] [Green Version]
  48. Jiao, F.; Yen, H.; Hutchings, G.S.; Yonemoto, B.; Lua, Q.; Kleitz, F. Synthesis, structural characterization, and electrochemical performance of nanocast mesoporous Cu-/Fe-based oxides. J. Mater. Chem. A 2014, 2, 3065–3071. [Google Scholar] [CrossRef]
  49. Panigrahi, P.K.; Pathak, A. Microwave-assisted synthesis of WS2 nanowires through tetrathiotungstate precursors. Sci. Technol. Adv. Mater. 2008, 9, 045008. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, H.; Su, Y.; Chen, P.; Wang, Y. Microwave-assisted solvothermal synthesis of 3D carnation-like SnS2 nanostructures with high visible light photocatalytic activity. J. Mol. Catal. Chem. 2013, 378, 285–292. [Google Scholar] [CrossRef]
  51. Rao, B.G.; Mukherjee, D.; Reddy, B.M. Novel Approaches for Preparation of Nanoparticles; Elsevier Inc.: Amsterdam, The Netherlands, 2017; ISBN 9780323461481. [Google Scholar]
  52. Kumar, A. Different Methods Used for the Synthesis of TiO2 Based Nanomaterials: A Review. Am. J. Nano Res. Appl. 2018, 6, 1–10. [Google Scholar] [CrossRef]
  53. Aliofkhazraei, M. (Ed.) Handbook of Nanoparticles; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar]
  54. Pokrovski, G.S.; Roux, J.; Hazemann, J.-L.; Borisova, A.; Gonchar, A.A.; Lemeshko, M.P. In situ X-ray absorption spectroscopy measurement of vapour-brine fractionation of antimony at hydrothermal conditions. Miner. Mag. 2008, 72, 667–681. [Google Scholar] [CrossRef]
  55. Gautam, A.K.; Faraz, M.; Khare, N. Enhanced thermoelectric properties of MoS2 with the incorporation of reduced graphene oxide (RGO). J. Alloys Compd. 2020, 838, 155673. [Google Scholar] [CrossRef]
  56. Bricha, M.; Belmamouni, Y.; Essassi, E.M.; Ferreira, J.; El Mabrouk, K. Surfactant-Assisted Hydrothermal Synthesis of Hydroxyapatite Nanopowders. J. Nanosci. Nanotechnol. 2012, 12, 8042–8049. [Google Scholar] [CrossRef]
  57. Cao, S.; Liu, T.; Zeng, W.; Hussain, S.; Peng, X.; Pan, F. Synthesis and characterization of flower-like WS2 nanospheres via a facile hydrothermal route. J. Mater. Sci. Mater. Electron. 2014, 25, 4300–4305. [Google Scholar] [CrossRef]
  58. Naskar, B.; Dan, A.; Ghosh, S.; Aswal, V.; Moulik, S.P. Revisiting the self-aggregation behavior of cetyltrimethylammonium bromide in aqueous sodium salt solution with varied anions. J. Mol. Liq. 2012, 170, 1–10. [Google Scholar] [CrossRef]
  59. Vollath, D.; Szabó, D.V. Synthesis of nanocrystalline MoS2 and WS2 in a microwave plasma. Mater. Lett. 1998, 35, 236–244. [Google Scholar] [CrossRef]
  60. Shanmugaratnam, S.; Rasalingam, S. Transition Metal Chalcogenide (TMC) Nanocomposites for Environmental Remediation Application over Extended Solar Irradiation. In Nanocatalysts; IntechOpen: London, UK, 2012. [Google Scholar] [CrossRef]
  61. Park, S.; Park, J.; Selvaraj, R.; Kim, Y. Facile microwave-assisted synthesis of SnS2 nanoparticles for visible-light responsive photocatalyst. J. Ind. Eng. Chem. 2015, 31, 269–275. [Google Scholar] [CrossRef]
  62. Choubey, S.K.; Tiwary, K.P. Microwave assisted synthesis of CdS nanoparticles for structural and optical characterization. IJIRSET 2014, 3, 10670–10674. [Google Scholar]
  63. Nethravathi, C.; Rajamathi, J.T.; Rajamathi, M. Microwave-Assisted Synthesis of Porous Aggregates of CuS Nanoparticles for Sunlight Photocatalysis. ACS Omega 2019, 4, 4825–4831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Pang, H.; Wei, C.; Li, X.; Li, G.; Ma, Y.; Li, S.; Chen, J.; Zhang, J. Microwave-assisted synthesis of NiS2 nanostructures for supercapacitors and cocatalytic enhancing photocatalytic H2 production. Sci. Rep. 2014, 4, 1–8. [Google Scholar] [CrossRef] [Green Version]
  65. Amaresh, S.; Karthikeyan, K.; Jang, I.-C.; Lee, Y.S. Single-step microwave mediated synthesis of the CoS2 anode material for high rate hybrid supercapacitors. J. Mater. Chem. A 2014, 2, 11099–11106. [Google Scholar] [CrossRef]
  66. Liu, B.; Qu, S.; Kou, Y.; Liu, Z.; Chen, X.; Wu, Y.; Han, X.; Deng, Y.; Hu, W.; Zhong, C. In Situ Electrodeposition of Cobalt Sulfide Nanosheet Arrays on Carbon Cloth as a Highly Efficient Bifunctional Electrocatalyst for Oxygen Evolution and Reduction Reactions. ACS Appl. Mater. Interfaces 2018, 10, 30433–30440. [Google Scholar] [CrossRef] [PubMed]
  67. Ghahremaninezhad, A.; Asselin, E.; Dixon, D.G. Electrodeposition and Growth Mechanism of Copper Sulfide Nanowires. J. Phys. Chem. C 2011, 115, 9320–9334. [Google Scholar] [CrossRef]
  68. Mammadov, M.N.; Sh Aliyev, A.; Elrouby, M. Electrodeposition of cadmium sulfide. Int. J. Thin Film Sci. Tec. 2012, 1, 43–53. [Google Scholar]
  69. Huang, J.; Daryadel, S.; Minary-Jolandan, M. Low-Cost Manufacturing of Metal–Ceramic Composites through Electrodeposition of Metal into Ceramic Scaffold. ACS Appl. Mater. Interfaces 2019, 11, 4364–4372. [Google Scholar] [CrossRef]
  70. Maharana, H.; Basu, A.; Mondal, K. Structural and tribological correlation of electrodeposited solid lubricating Ni-WSe2 composite coating. Surf. Coatings Technol. 2018, 349, 328–339. [Google Scholar] [CrossRef]
  71. Safavi, M.S.; Fathi, M.; Charkhesht, V.; Jafarpour, M.; Ahadzadeh, I. Electrodeposition of Co-P Coatings Reinforced by MoS2 + Y2O3 Hybrid Ceramic Nanoparticles for Corrosion-Resistant Applications: Influences of Operational Parameters. Met. Mater. Trans. A 2020, 51, 6740–6758. [Google Scholar] [CrossRef]
  72. Jayakrishnan, D.S. Electrodeposition: The Versatile Technique for Nanomaterials; Woodhead Publishing Limited: Sawston, UK, 2012. [Google Scholar]
  73. Aruna, S.; Bindu, C.; Selvi, V.E.; Grips, V.W.; Rajam, K. Synthesis and properties of electrodeposited Ni/ceria nanocomposite coatings. Surf. Coatings Technol. 2006, 200, 6871–6880. [Google Scholar] [CrossRef]
  74. Shariza, S.; Anand, T.J.S. Effect of deposition time on the structural and optical properties of molybdenum chalcogenides thin films. Chalcogenide Lett. 2011, 8, 529–539. [Google Scholar]
  75. Chen, Y.-C.; Huang, Y.-S.; Huang, H.; Su, P.-J.; Perng, T.-P.; Chen, L.-J. Photocatalytic enhancement of hydrogen production in water splitting under simulated solar light by band gap engineering and localized surface plasmon resonance of ZnxCd1−xS nanowires decorated by Au nanoparticles. Nano Energy 2020, 67, 104225. [Google Scholar] [CrossRef]
  76. Sankapal, B.R.; Mane, R.S.; Lokhande, C.D. Successive ionic layer adsorption and reaction (SILAR) method for the deposition of large area (~10 cm2) tin disulfide (SnS2) thin films. Mater. Res. Bull. 2000, 35, 2027–2035. [Google Scholar] [CrossRef]
  77. Pathan, H.M.; Lokhande, C.D. Deposition of metal chalcogenide thin films by successive ionic layer adsorption and reaction (SILAR) method. Bull. Mater. Sci 2004, 27, 85–111. [Google Scholar]
  78. Leskell, M. Deposition of manganese-doped successive ionic layer adsorption zinc sulfide thin films by the and reaction (SILAR) method. Thin Solid Film. 1995, 263, 79–84. [Google Scholar]
  79. Lindroos, S.; Kanniainen, T.; Leskelä, M. Growth of zinc sulfide thin films by the successive ionic layer adsorption and reaction (SILAR) method on polyester substrates. Mater. Res. Bull. 1997, 32, 1631–1636. [Google Scholar] [CrossRef]
  80. Valkonen, M.P.; Kanniainen, T.; Lindroos, S.; Leskelä, M.; Rauhala, E. Growth of ZnS, CdS and multilayer ZnS/CdS thin films by SILAR technique. Appl. Surf. Sci. 1997, 115, 386–392. [Google Scholar] [CrossRef]
  81. Aditha, S.K.; Kurdekar, A.D.; Chunduri, L.A.A.; Patnaik, S.; Kamisetti, V. Aqueous based reflux method for green synthesis of nanostructures: Application in CZTS synthesis. MethodsX 2016, 3, 35–42. [Google Scholar] [CrossRef] [PubMed]
  82. Phuruangrat, A.; Thoonchalong, P.; Thongtem, S.; Thongtem, T. Synthesis of cus with different morphologies by refluxing method: Nanopaticles in clusters and nanoflakes in sponge-like clusters. Chalcogenide Lett. 2012, 9, 421–426. [Google Scholar]
  83. Wang, L.-C.; Bao, S.-K.; Luo, J.; Wang, Y.-H.; Nie, Y.-C.; Zou, J.-P. Efficient exfoliation of bulk MoS2 to nanosheets by mixed-solvent refluxing method. Int. J. Hydrogen Energy 2016, 41, 10737–10743. [Google Scholar] [CrossRef]
  84. Phuruangrat, A.; Karthik, K.; Kuntalue, B.; Dumrongrojthanath, P.; Thongtem, S.; Thongtem, T.; Yai, H.; Nadu, T.; Science, M. Refuxing synthesis and characterization of zns nanoparticles and their photocatalytic properties. Chalcogenide Lett. 2019, 16, 387–393. [Google Scholar]
  85. Dai, Z.; Zang, X.; Yang, J.; Sun, C.; Si, W.; Huang, W.; Dong, X. Template Synthesis of Shape-Tailorable NiS2 Hollow Prisms as High-Performance Supercapacitor Materials. ACS Appl. Mater. Interfaces 2015, 7, 25396–25401. [Google Scholar] [CrossRef] [PubMed]
  86. Li, J.; Yang, X.; Liu, Y.; Huang, B.; Wu, R.; Zhang, Z.; Zhao, B.; Ma, H.; Dang, W.; Wei, Z.; et al. General synthesis of two-dimensional van der Waals heterostructure arrays. Nat. Cell Biol. 2020, 579, 368–374. [Google Scholar] [CrossRef] [PubMed]
  87. Gong, Y.; Lin, J.; Wang, X.; Shi, G.; Lei, S.; Lin, Z.; Zou, X.; Ye, G.; Vajtai, R.; Yakobson, B.I.; et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 2014, 13, 1135–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Hao, S.; Yang, B.; Gao, Y. Orientation-specific transgranular fracture behavior of CVD-grown monolayer MoS2 single crystal. Appl. Phys. Lett. 2017, 110, 153105. [Google Scholar] [CrossRef]
  89. Kim, S.; Kwak, J.; Ciobanu, C.V.; Kwon, S. Recent Developments in Controlled Vapor-Phase Growth of 2D Group 6 Transition Metal Dichalcogenides. Adv. Mater. 2019, 31, e1804939. [Google Scholar] [CrossRef]
  90. Tan, H.; Fan, Y.; Zhou, Y.; Chen, Q.; Xu, W.; Warner, J.H. Ultrathin 2D Photodetectors Utilizing Chemical Vapor Deposition Grown WS2 With Graphene Electrodes. ACS Nano 2016, 10, 7866–7873. [Google Scholar] [CrossRef] [PubMed]
  91. Chang, R.-J.; Tan, H.; Wang, X.; Porter, B.; Chen, T.; Sheng, Y.; Zhou, Y.; Huang, H.; Bhaskaran, H.; Warner, J.H. High-Performance All 2D-Layered Tin Disulfide: Graphene Photodetecting Transistors with Thickness-Controlled Interface Dynamics. ACS Appl. Mater. Interfaces 2018, 10, 13002–13010. [Google Scholar] [CrossRef] [PubMed]
  92. Cong, C.; Shang, J.; Wu, X.; Cao, B.; Peimyoo, N.; Qiu, C.; Sun, L.; Yu, T. Synthesis and Optical Properties of Large-Area Single-Crystalline 2D Semiconductor WS2Monolayer from Chemical Vapor Deposition. Adv. Opt. Mater. 2014, 2, 131–136. [Google Scholar] [CrossRef]
  93. Ye, G.; Gong, Y.; Lei, S.; He, Y.; Li, B.; Zhang, X.; Jin, Z.; Dong, L.; Lou, J.; Vajtai, R.; et al. Synthesis of large-scale atomic-layer SnS2 through chemical vapor deposition. Nano Res. 2017, 10, 2386–2394. [Google Scholar] [CrossRef]
  94. Dai, C.; Li, B.; Li, J.; Zhao, B.; Wu, R.; Ma, H.; Duan, X. Controllable synthesis of NiS and NiS2 nanoplates by chemical vapor deposition. Nano Res. 2020, 13, 2506–2511. [Google Scholar] [CrossRef]
  95. Ardahe, M.; Hantehzadeh, M.R.; Ghoranneviss, M. Effect of Growth Temperature on Physical Properties of MoS2 Thin Films Synthesized by CVD. J. Electron. Mater. 2019, 49, 1002–1008. [Google Scholar] [CrossRef]
  96. Samad, L.; Cabán-Acevedo, M.; Shearer, M.; Park, K.; Hamers, R.; Jin, S. Direct Chemical Vapor Deposition Synthesis of Phase-Pure Iron Pyrite (FeS2) Thin Films. Chem. Mater. 2015, 27, 3108–3114. [Google Scholar] [CrossRef]
  97. Backes, C.; Szydłowska, B.M.; Harvey, A.; Yuan, S.; Vega-Mayoral, V.; Davies, B.R.; Zhao, P.-L.; Hanlon, D.; Santos, E.J.G.; Katsnelson, M.I.; et al. Production of Highly Monolayer Enriched Dispersions of Liquid-Exfoliated Nanosheets by Liquid Cascade Centrifugation. ACS Nano 2016, 10, 1589–1601. [Google Scholar] [CrossRef]
  98. Mihai, C.; Sava, F.; Galca, A.C.; Velea, A. Low power non-volatile memory switching in monolayer-rich 2D WS2 and MoS2 devices. AIP Adv. 2020, 10, 025102. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, T.-Y.; Chang, Y.-H.; Hsu, C.-L.; Wei, K.-H.; Chiang, C.-Y.; Li, L.-J. Comparative study on MoS2 and WS2 for electrocatalytic water splitting. Int. J. Hydrogen Energy 2013, 38, 12302–12309. [Google Scholar] [CrossRef]
  100. Ma, L.; Chen, W.-X.; Xu, Z.-D.; Xia, J.-B.; Li, X. Carbon nanotubes coated with tubular MoS2layers prepared by hydrothermal reaction. Nanotechnology 2006, 17, 571–574. [Google Scholar] [CrossRef]
  101. Li, Q.; Zhang, N.; Yang, Y.; Wang, G.; Ng, D.H.L. High Efficiency Photocatalysis for Pollutant Degradation with MoS2/C3N4 Heterostructures. Langmuir 2014, 30, 8965–8972. [Google Scholar] [CrossRef] [PubMed]
  102. Lukowski, M.A.; Daniel, A.S.; Meng, F.; Forticaux, A.; Li, L.; Jin, S. Enhanced Hydrogen Evolution Catalysis from Chemically Exfoliated Metallic MoS2 Nanosheets. J. Am. Chem. Soc. 2013, 135, 10274–10277. [Google Scholar] [CrossRef]
  103. Peng, Y.; Meng, Z.; Zhong, C.; Lu, J.; Yu, W.; Jia, Y.; Qian, Y. Hydrothermal synthesis and characterization of single-molecular-layer MoS2 and MoSe2. Chem. Lett. 2001, 30, 772–773. [Google Scholar] [CrossRef]
  104. Afanasiev, P.; Xia, G.-F.; Berhault, G.; Jouguet, B.; Lacroix, M. Surfactant-Assisted Synthesis of Highly Dispersed Molybdenum Sulfide. Chem. Mater. 1999, 11, 3216–3219. [Google Scholar] [CrossRef]
  105. Shanmugaratnam, S.; Velauthapillai, D.; Ravirajan, P.; Christy, A.A.; Shivatharsiny, Y. CoS2/TiO2 Nanocomposites for Hydrogen Production under UV Irradiation. Materials 2019, 12, 3882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Shanmugaratnam, S.; Selvaratnam, B.; Baride, A.; Koodali, R.; Ravirajan, P.; Velauthapillai, D.; Shivatharsiny, Y. SnS2/TiO2 Nanocomposites for Hydrogen Production and Photodegradation under Extended Solar Irradiation. Catalysts 2021, 11, 589. [Google Scholar] [CrossRef]
  107. Cao, J.; Zhou, J.; Zhang, Y.; Wang, Y.; Liu, X. Dominating Role of Aligned MoS2/Ni3S2 Nanoarrays Supported on Three-Dimensional Ni Foam with Hydrophilic Interface for Highly Enhanced Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2018, 10, 1752–1760. [Google Scholar] [CrossRef]
  108. Gao, B.; Du, X.; Ma, Y.; Li, Y.; Li, Y.; Ding, S.; Song, Z.; Xiao, C. 3D flower-like defected MoS2 magnetron-sputtered on candle soot for enhanced hydrogen evolution reaction. Appl. Catal. Environ. 2020, 263, 117750. [Google Scholar] [CrossRef]
  109. Nadeem Riaz, K.; Yousaf, N.; Bilal Tahir, M.; Israr, Z.; Iqbal, T. Facile hydrothermal synthesis of 3D flower-like La-MoS2 nanostructure for photocatalytic hydrogen energy production. Int. J. Energy Res. 2019, 43, 491–499. [Google Scholar] [CrossRef] [Green Version]
  110. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  111. Gaur, R.; Jeevanandam, P. Synthesis of SnS2 Nanoparticles and Their Application as Photocatalysts for the Reduction of Cr(VI). J. Nanosci. Nanotechnol. 2018, 18, 165–177. [Google Scholar] [CrossRef]
  112. Zhu, L.; Susac, D.; Teo, M.; Wong, K.; Wong, P.; Parsons, R.; Bizzotto, D.; Mitchell, K.; Campbell, S. Investigation of CoS2-based thin films as model catalysts for the oxygen reduction reaction. J. Catal. 2008, 258, 235–242. [Google Scholar] [CrossRef]
  113. Wang, J.; Yin, Y. Fermentative hydrogen production using pretreated microalgal biomass as feedstock. Microb. Cell Factories 2018, 17, 1–16. [Google Scholar] [CrossRef] [Green Version]
  114. Kye, J.; Shin, M.; Lim, B.; Jang, J.-W.; Oh, I.; Hwang, S. Platinum Monolayer Electrocatalyst on Gold Nanostructures on Silicon for Photoelectrochemical Hydrogen Evolution. ACS Nano 2013, 7, 6017–6023. [Google Scholar] [CrossRef]
  115. Mckone, J.R.; Pieterick, A.P.; Gray, H.B.; Lewis, N.S. Supporting Information: Hydrogen Evolution from Pt /Ru-Coated p-Type WSe2 Photocathodes. J. Am. Chem. Soc. 2013, 135, 223–231. [Google Scholar] [CrossRef] [PubMed]
  116. Wiensch, J.D.; John, J.; Velazquez, J.M.; Torelli, D.A.; Pieterick, A.P.; McDowell, M.T.; Sun, K.; Zhao, X.; Brunschwig, B.S.; Lewis, N.S. Comparative Study in Acidic and Alkaline Media of the Effects of pH and Crystallinity on the Hydrogen-Evolution Reaction on MoS2 and MoSe2. ACS Energy Lett. 2017, 2, 2234–2238. [Google Scholar] [CrossRef] [Green Version]
  117. Sun, B.; Wang, H.; Wu, J.; Geng, Y.; Xu, J.; Wang, Y.; Li, Y.; Lin, H.; Wang, L. Designed synthesis of unique ZnS@CdS@Cd0.5Zn0.5S-MoS2 hollow nanospheres for efficient visible-light-driven H2 evolution. CrystEngComm 2020, 22, 2743–2755. [Google Scholar] [CrossRef]
  118. Komba, N.; Zhang, G.; Pu, Z.; Wu, M.; Rosei, F.; Sun, S. MoS2-supported on free-standing TiO2-nanotubes for efficient hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 4468–4480. [Google Scholar] [CrossRef]
  119. Zhao, X.; Ma, X.; Sun, J.; Li, D.; Yang, X. Enhanced Catalytic Activities of Surfactant-Assisted Exfoliated WS2 Nanodots for Hydrogen Evolution. ACS Nano 2016, 10, 2159–2166. [Google Scholar] [CrossRef] [PubMed]
  120. Mao, Q.; Chen, J.; Chen, H.; Chen, Z.; Chen, J.; Li, Y. Few-layered 1T-MoS2-modified ZnCoS solid-solution hollow dodecahedra for enhanced photocatalytic hydrogen evolution. J. Mater. Chem. A 2019, 7, 8472–8484. [Google Scholar] [CrossRef]
  121. Oertel, M.; Schmitz, J.; Weirich, W.; Jendryssek-Neumann, D.; Schulten, R. Vapour-liquid Equilibrium Data Collec-tion’ dechema Frankfurt. Izv. Vyss. Uchebn. Zaved. Khim. Tekhnol. 1987, 27, 686–691. [Google Scholar]
  122. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar] [CrossRef]
  123. Dehghanimadvar, M.; Shirmohammadi, R.; Sadeghzadeh, M.; Aslani, A.; Ghasempour, R. Hydrogen production technologies: Attractiveness and future perspective. Int. J. Energy Res. 2020, 44, 8233–8254. [Google Scholar] [CrossRef]
  124. FreedomCAR and Fuel Partnership Hydrogen Offers Sustainable Solutions to Our Nation’s Energy and Climate Challenges. Table of Contents Sustainability Urban Air Quality. 2009. Available online: https://www1.eere.energy.gov/hydrogenandfuelcells/pdfs/h2_tech_roadmap.pdf (accessed on 11 October 2021).
  125. Luo, X.; Wu, T.; Shi, K.; Song, M.; Rao, Y. Biomass Gasification: An Overview of Technological Barriers and Socio-Environmental Impact. Gasif. Low-Grade Feedstock 2018, 3–18. [Google Scholar] [CrossRef] [Green Version]
  126. Ding, Q.; Song, B.; Xu, P.; Jin, S. Efficient Electrocatalytic and Photoelectrochemical Hydrogen Generation Using MoS2 and Related Compounds. Chem 2016, 1, 699–726. [Google Scholar] [CrossRef] [Green Version]
  127. Chu, D.; Zhang, C.; Yang, P.; Du, Y.; Lu, C. WS2 as an Effective Noble-Metal Free Cocatalyst Modified TiSi2 for Enhanced Photocatalytic Hydrogen Evolution under Visible Light Irradiation. Catalysts 2016, 6, 136. [Google Scholar] [CrossRef] [Green Version]
  128. Zhang, J.; Hu, W.; Cao, S.; Piao, L. Recent progress for hydrogen production by photocatalytic natural or simulated seawater splitting. Nano Res. 2020, 13, 2313–2322. [Google Scholar] [CrossRef]
  129. Liu, J.; Yu, G.; Zhang, R.; Huang, X.; Chen, W. Theoretical predication of the high hydrogen evolution catalytic activity for the cubic and tetragonal SnP systems. Phys. Chem. Chem. Phys. 2019, 21, 5521–5530. [Google Scholar] [CrossRef] [PubMed]
  130. Som, N.N.; Jha, P.K. Hydrogen evolution reaction of metal di-chalcogenides: ZrS2, ZrSe2 and Janus ZrSSe. Int. J. Hydrogen Energy 2020, 45, 23920–23927. [Google Scholar] [CrossRef]
  131. Wang, B.; Wang, X.; Yuan, H.; Zhou, T.; Chang, J.; Chen, H. Direct Z-scheme photocatalytic overall water splitting on two dimensional MoSe2/SnS2 heterojunction. Int. J. Hydrogen Energy 2020, 45, 2785–2793. [Google Scholar] [CrossRef]
  132. Park, H.; Akande, A.; Sanvito, S.; El-Mellouhi, F. The rise of Nb-, Ta-, and Bi-based oxides/chalcogenides for photocatalytic applications. Int. J. Hydrogen Energy 2021, 1–13. [Google Scholar] [CrossRef]
  133. Ran, N.; Qiu, W.; Song, E.; Wang, Y.; Zhao, X.; Liu, Z.; Liu, J. Bond Electronegativity as Hydrogen Evolution Reaction Catalyst Descriptor for Transition Metal (TM = Mo, W) Dichalcogenides. Chem. Mater. 2020, 32, 1224–1234. [Google Scholar] [CrossRef]
  134. Faraji, M.; Yousefi, M.; Yousefzadeh, S.; Zirak, M.; Naseri, N.; Jeon, T.H.; Choi, W.; Moshfegh, A.Z. Two-dimensional materials in semiconductor photoelectrocatalytic systems for water splitting. Energy Environ. Sci. 2019, 12, 59–95. [Google Scholar] [CrossRef]
  135. Huang, D.; Wen, M.; Zhou, C.; Li, Z.; Cheng, M.; Chen, S.; Xue, W.; Lei, L.; Yang, Y.; Xiong, W.; et al. ZnxCd1−xS based materials for photocatalytic hydrogen evolution, pollutants degradation and carbon dioxide reduction. Appl. Catal. Environ. 2020, 267, 118651. [Google Scholar] [CrossRef]
  136. Kudo, A.; Kato, H.; Tsuji, I. Strategies for the Development of Visible-light-driven Photocatalysts for Water Splitting. Chem. Lett. 2004, 33, 1534–1539. [Google Scholar] [CrossRef]
  137. Li, K.; Li, Y.; Wang, Y.; Ge, J.; Liu, C.; Xing, W. Enhanced electrocatalytic performance for the hydrogen evolution reaction through surface enrichment of platinum nanoclusters alloying with ruthenium in situ embedded in carbon. Energy Environ. Sci. 2018, 11, 1232–1239. [Google Scholar] [CrossRef]
  138. Ahn, S.; Yang, J.; Lim, H.; Shin, H.S. Selective synthesis of pure cobalt disulfide on reduced graphene oxide sheets and its high electrocatalytic activity for hydrogen evolution reaction. Nano Converg. 2016, 3, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Gao, C.; Hua, H.; Du, M.; Liu, J.; Wu, X.; Pu, Y.; Li, X. 1T/2H MoS2 nanoflowers decorated amorphous Mo-CoSx skeleton: A ZIF-based composite electrocatalyst for the hydrogen evolution reaction. Appl. Surf. Sci. 2020, 515, 145842. [Google Scholar] [CrossRef]
  140. Bhat, K.S.; Nagaraja, H. Performance evaluation of molybdenum dichalcogenide (MoX2; X= S, Se, Te) nanostructures for hydrogen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 17878–17886. [Google Scholar] [CrossRef]
  141. Li, H.; Yang, C.; Wang, X.; Zhang, J.; Xi, J.; Du, G.; Ji, Z. Mixed 3D/2D dimensional TiO2 nanoflowers/MoSe2 nanosheets for enhanced photoelectrochemical hydrogen generation. J. Am. Ceram. Soc. 2020, 103, 1187–1196. [Google Scholar] [CrossRef]
  142. Sharma, M.D.; Mahala, C.; Basu, M. 2D Thin Sheet Heterostructures of MoS2 on MoSe2 as Efficient Electrocatalyst for Hydrogen Evolution Reaction in Wide pH Range. Inorg. Chem. 2020, 59, 4377–4388. [Google Scholar] [CrossRef]
  143. Xia, Y.; Wu, W.; Wang, H.; Rao, S.; Zhang, F.; Zou, G. Amorphous RuS2 electrocatalyst with optimized active sites for hydrogen evolution. Nanotechnology 2020, 31, 145401. [Google Scholar] [CrossRef] [PubMed]
  144. Zhang, D.; Mou, H.; Chen, L.; Wang, D.; Song, C. Design and in-situ synthesis of unique catalyst via embedding graphene oxide shell membrane in NiS2 for efficient hydrogen evolution. Appl. Surf. Sci. 2020, 510, 145483. [Google Scholar] [CrossRef]
  145. Manjunathaae, C.; Patil, R.S.; Sudeep, M.; Srinivasa, N.; Kumar, R.C.; Aan, M.S.; Ashoka, S. Rational design and synthesis of hetero-nanostructured electrospun PU@PANI@FeS2: A surface tailored hybrid catalyst for H2 production via electrochemical splitting of water. Surf. Interfaces 2020, 18, 100445. [Google Scholar] [CrossRef]
  146. Zeng, X.; Jiao, Q.; Li, N.; Wang, J. Facile synthesis of CuxS coated electrodes for the efficient hydrogen evolution reaction. Appl. Surf. Sci. 2020, 513, 145785. [Google Scholar] [CrossRef]
  147. Li, Z.; Liu, P.; Yang, G. Efficient All-2D Amorphous Cobalt Sulfide Nanosheets/Multilayered Molybdenum Disulfide Hybrid Heterojunction Catalyst for Electrochemical Hydrogen Evolution. Glob. Chall. 2020, 4, 1900066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Zhu, X.; Dai, J.; Li, L.; Zhao, D.; Wu, Z.; Tang, Z.; Ma, L.-J.; Chen, S. Hierarchical carbon microflowers supported defect-rich Co3S4 nanoparticles: An efficient electrocatalyst for water splitting. Carbon 2020, 160, 133–144. [Google Scholar] [CrossRef]
  149. Aslan, E.; Sarilmaz, A.; Yanalak, G.; Chang, C.S.; Cinar, I.; Ozel, F.; Patir, I.H. Facile preparation of amorphous NiWSex and CoWSex nanoparticles for the electrocatalytic hydrogen evolution reaction in alkaline condition. J. Electroanal. Chem. 2019, 856, 113674. [Google Scholar] [CrossRef]
  150. Li, G.; Chen, Z.; Li, Y.; Zhang, D.; Yang, W.; Liu, Y.; Cao, L. Engineering Substrate Interaction to Improve Hydrogen Evolution Catalysis of Monolayer MoS2 Films beyond Pt. ACS Nano 2020, 14, 1707–1714. [Google Scholar] [CrossRef] [PubMed]
  151. Nguyen, V.-T.; Le, P.A.; Hsu, Y.-C.; Wei, K.-H. Plasma-Induced Exfoliation Provides Onion-Like Graphene-Surrounded MoS2 Nanosheets for a Highly Efficient Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 11533–11542. [Google Scholar] [CrossRef] [PubMed]
  152. Shi, Q.; Zhou, Y.; Sun, W.; Wang, H.; Li, S.; Zhang, X.; Wan, J. Hydrogen Evolution and Wastewater Treatment of Hydrangeal-like Catalyst Decroatedby the NiS Nanosheet and PdNanoparticle. ChemistrySelect 2020, 5, 1041–1046. [Google Scholar] [CrossRef]
  153. Zhao, A.; Zhang, L.; Xu, G.; Zhang, X.; Zhang, S.; Xiao, Y. Hollow ZnxCo1−xSe2 microcubes derived from Metal–Organic framework as efficient bifunctional electrocatalysts for hydrogen evolution and oxygen evolution reactions. Int. J. Hydrogen Energy 2020, 45, 2607–2616. [Google Scholar] [CrossRef]
  154. Zhou, Y.; Zhang, J.; Ren, H.; Pan, Y.; Yan, Y.; Sun, F.; Wang, X.; Wang, S.; Zhang, J. Mo doping induced metallic CoSe for enhanced electrocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 268, 118467. [Google Scholar] [CrossRef]
  155. Liang, Y.; Yang, Y.; Xu, K.; Yu, T.; Yao, S.; Peng, Q.; Yuan, C. Crystal plane dependent electrocatalytic performance of NiS2 nanocrystals for hydrogen evolution reaction. J. Catal. 2020, 381, 63–69. [Google Scholar] [CrossRef]
  156. Friedman, F.; Hastie, T.; John, T.; Weatherwax, R.L.; Epstein, D.; Wasserman, L.; Abu-El-Haija, S.; Do, C.B.; Batzoglou, S.; Resknik, C.; et al. Solution-processed GaSe nanoflake-based films for photoelectrochemical water splitting and photoelectrochemical-type. J. Chem. Inf. Model. 2013, 67, 1689–1699. [Google Scholar] [CrossRef]
  157. Ma, F.; Liang, Y.; Zhou, P.; Tong, F.; Wang, Z.; Wang, P.; Liu, Y.; Dai, Y.; Zheng, Z.; Huang, B. One-step synthesis of Co-doped 1T-MoS2 nanosheets with efficient and stable HER activity in alkaline solutions. Mater. Chem. Phys. 2020, 244, 122642. [Google Scholar] [CrossRef]
  158. Song, W.; Wang, K.; Jin, G.; Wang, Z.; Li, C.; Yang, X.; Chen, C. Two-Step Hydrothermal Synthesis of CoSe/MoSe 2 as Hydrogen Evolution Electrocatalysts in Acid and Alkaline Electrolytes. ChemElectroChem 2019, 6, 4842–4847. [Google Scholar] [CrossRef]
  159. Yang, J.-X.; Yu, W.-B.; Li, C.-F.; Dong, W.-D.; Jiang, L.-Q.; Zhou, N.; Zhuang, Z.-P.; Liu, J.; Hu, Z.-Y.; Zhao, H.; et al. PtO nanodots promoting Ti3C2 MXene in-situ converted Ti3C2/TiO2 composites for photocatalytic hydrogen production. Chem. Eng. J. 2021, 420, 129695. [Google Scholar] [CrossRef]
  160. Guayaquil-Sosa, J.; Rosales, B.S.; Valadés-Pelayo, P.; De Lasa, H. Photocatalytic hydrogen production using mesoporous TiO2 doped with Pt. Appl. Catal. Environ. 2017, 211, 337–348. [Google Scholar] [CrossRef]
  161. Zhao, H.; Li, C.-F.; Hu, Z.-Y.; Liu, J.; Li, Y.; Hu, J.; Van Tendeloo, G.; Chen, L.-H.; Su, B.-L. Size effect of bifunctional gold in hierarchical titanium oxide-gold-cadmium sulfide with slow photon effect for unprecedented visible-light hydrogen production. J. Colloid Interface Sci. 2021, 604, 131–140. [Google Scholar] [CrossRef] [PubMed]
  162. Yerga, R.M.N.; Alvarez-Galvan, M.C.; Del Valle, F.; De La Mano, J.A.V.; Fierro, J.L.G. Water Splitting on Semiconductor Catalysts under Visible-Light Irradiation. ChemSusChem 2009, 2, 471–485. [Google Scholar] [CrossRef] [PubMed]
  163. Wang, Q.; Huang, J.; Sun, H.; Zhang, K.-Q.; Lai, Y. Uniform carbon dots@TiO2 nanotube arrays with full spectrum wavelength light activation for efficient dye degradation and overall water splitting. Nanoscale 2017, 9, 16046–16058. [Google Scholar] [CrossRef]
  164. Onsuratoom, S.; Chavadej, S.; Sreethawong, T. Hydrogen production from water splitting under UV light irradiation over Ag-loaded mesoporous-assembled TiO2–ZrO2 mixed oxide nanocrystal photocatalysts. Int. J. Hydrogen Energy 2011, 36, 5246–5261. [Google Scholar] [CrossRef]
  165. Chen, J.; Wu, X.-J.; Yin, L.; Li, B.; Hong, X.; Fan, Z.; Chen, B.; Xue, C.; Zhang, H. One-pot Synthesis of CdS Nanocrystals Hybridized with Single-Layer Transition-Metal Dichalcogenide Nanosheets for Efficient Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2015, 54, 1210–1214. [Google Scholar] [CrossRef]
  166. Zong, X.; Han, J.; Ma, G.; Yan, H.; Wu, G.; Li, C. Photocatalytic H2 Evolution on CdS Loaded with WS2 as Cocatalyst under Visible Light Irradiation. J. Phys. Chem. C 2011, 115, 12202–12208. [Google Scholar] [CrossRef]
  167. Ran, J.; Zhang, J.; Yu, J.; Jaroniec, M.; Qiao, S.Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. [Google Scholar] [CrossRef] [PubMed]
  168. Yan, C.; Xue, X.; Zhang, W.; Li, X.; Liu, J.; Yang, S.; Hu, Y.; Chen, R.; Yan, Y.; Zhu, G.; et al. Well-designed Te/SnS2/Ag artificial nanoleaves for enabling and enhancing visible-light driven overall splitting of pure water. Nano Energy 2017, 39, 539–545. [Google Scholar] [CrossRef]
  169. Liu, E.; Chen, J.; Ma, Y.; Feng, J.; Jia, J.; Fan, J.; Hu, X. Fabrication of 2D SnS2/g-C3N4 heterojunction with enhanced H2 evolution during photocatalytic water splitting. J. Colloid Interface Sci. 2018, 524, 313–324. [Google Scholar] [CrossRef] [PubMed]
  170. Ke, X.; Dai, K.; Zhu, G.; Zhang, J.; Liang, C. In situ photochemical synthesis noble-metal-free NiS on CdS-diethylenetriamine nanosheets for boosting photocatalytic H2 production activity. Appl. Surf. Sci. 2019, 481, 669–677. [Google Scholar] [CrossRef]
  171. Khan, Z.; Chetia, T.R.; Vardhaman, A.K.; Barpuzary, D.; Sastri, C.V.; Qureshi, M. Visible light assisted photocatalytic hydrogen generation and organic dye degradation by CdS–metal oxide hybrids in presence of graphene oxide. RSC Adv. 2012, 2, 12122–12128. [Google Scholar] [CrossRef]
  172. Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J. MoS2/Graphene Cocatalyst for Efficient Photocatalytic H2 Evolution under Visible Light Irradiation. ACS Nano 2014, 8, 7078–7087. [Google Scholar] [CrossRef]
  173. Jiang, D.; Zhu, L.; Irfan, R.M.; Zhang, L.; Du, P. Integrating noble-metal-free NiS cocatalyst with a semiconductor heterojunction composite for efficient photocatalytic H 2 production in water under visible light. Chin. J. Catal. 2017, 38, 2102–2109. [Google Scholar] [CrossRef]
  174. Xin, X.; Song, Y.; Guo, S.; Zhang, Y.; Wang, B.; Yu, J.; Li, X. In-situ growth of high-content 1T phase MoS2 confined in the CuS nanoframe for efficient photocatalytic hydrogen evolution. Appl. Catal. Environ. 2020, 269, 118773. [Google Scholar] [CrossRef]
  175. Nagappagari, L.R.; Samanta, S.; Sharma, N.; Battula, V.R.; Kailasam, K. Synergistic effect of a noble metal free Ni(OH)2 co-catalyst and a ternary ZnIn2S4/g-C3N4 heterojunction for enhanced visible light photocatalytic hydrogen evolution. Sustain. Energy Fuels 2020, 4, 750–759. [Google Scholar] [CrossRef]
  176. Guo, X.; Guo, P.; Wang, C.; Chen, Y.; Guo, L. Few-layer WSe2 nanosheets as an efficient cocatalyst for improved photocatalytic hydrogen evolution over Zn0.1Cd0.9S nanorods. Chem. Eng. J. 2020, 383, 123183. [Google Scholar] [CrossRef]
  177. Li, Y.; Zhang, D.; Xi, X.; Shen, Y.; Chang, S.; Zhu, J.; Chen, Y.; Jiang, L.; Zhang, J. Noble metal-free bimetallic NiCo decorated Zn0.5Cd0.5S solid solution for enhanced photocatalytic H2 evolution under visible light. Int. J. Hydrogen Energy 2020, 45, 8300–8309. [Google Scholar] [CrossRef]
  178. Jia, J.; Bai, X.; Zhang, Q.; Hu, X.; Liu, E.; Fan, J. Porous honeycomb-like NiSe2/red phosphorus heteroarchitectures for photocatalytic hydrogen production. Nanoscale 2020, 12, 5636–5651. [Google Scholar] [CrossRef] [PubMed]
  179. Zhu, Y.; Chen, J.; Shao, L.; Xia, X.; Liu, Y.; Wang, L. Oriented facet heterojunctions on CdS nanowires with high photoactivity and photostability for water splitting. Appl. Catal. B Environ. 2020, 268, 118744. [Google Scholar] [CrossRef]
  180. Han, J.; Liu, Y.; Wang, Y.; Zhao, R.; Wang, L. Construction of ternary CdxMo1−xSe quantum dots for enhanced photocatalytic hydrogen production. J. Mater. Sci. 2019, 55, 1117–1125. [Google Scholar] [CrossRef]
  181. Beall, C.; Piipari, K.; Al-qassab, H.; Smith, M.A. MoS2/CdIn2S4 flower-like heterojunctions with enhanced photocatalytic degradation and H2 evolution activity. Biochem. J. 2010, 53, 1–14. [Google Scholar]
  182. Mangiri, R.; Reddy, D.A.; Subramanyam, K.; Kumar, K.S.; Sudharani, A.; Poornaprakash, B.; Vijayalakshmi, R. Decorating MoS2 and CoSe2 nanostructures on 1D-CdS nanorods for boosting photocatalytic hydrogen evolution rate. J. Mol. Liq. 2019, 289, 2–11. [Google Scholar] [CrossRef]
  183. Wang, B.; Ding, Y.; Deng, Z.; Li, Z. Rational design of ternary NiS/CQDs/ZnIn2S4 nanocomposites as efficient noble-metal-free photocatalyst for hydrogen evolution under visible light. Chin. J. Catal. 2019, 40, 335–342. [Google Scholar] [CrossRef]
  184. Chen, Z.; Gong, H.; Liu, Q.; Song, M.; Huang, C. NiSe2 Nanoparticles Grown in Situ on CdS Nanorods for Enhanced Photocatalytic Hydrogen Evolution. ACS Sustain. Chem. Eng. 2019, 7, 16720–16728. [Google Scholar] [CrossRef]
  185. Yan, A.; Shi, X.; Huang, F.; Fujitsuka, M.; Majima, T. Efficient photocatalytic H2 evolution using NiS/ZnIn2S4 heterostructures with enhanced charge separation and interfacial charge transfer. Appl. Catal. Environ. 2019, 250, 163–170. [Google Scholar] [CrossRef]
  186. Rangappa, A.P.; Kumar, D.P.; Gopannagari, M.; Reddy, D.A.; Hong, Y.; Kim, Y.; Kim, T.K. Highly efficient hydrogen generation in water using 1D CdS nanorods integrated with 2D SnS2 nanosheets under solar light irradiation. Appl. Surf. Sci. 2020, 508, 144803. [Google Scholar] [CrossRef]
  187. Ranjith, K.S.; Kumar, D.R.; Huh, Y.S.; Shanmugam, R.K.; Uyar, T.; Kumar, R.T.R. Promotional Effect of Cu2S–ZnS Nanograins as a Shell Layer on ZnO Nanorod Arrays for Boosting Visible Light Photocatalytic H2 Evolution. J. Phys. Chem. C 2020, 124, 3610–3620. [Google Scholar] [CrossRef]
  188. Puentes-Prado, E.; Garcia, C.; Oliva, J.; Galindo, R.; Bernal-Alvarado, J.; Torres, L.D.; Gomez-Solis, C. Enhancing the solar photocatalytic hydrogen generation of ZnS films by UV radiation treatment. Int. J. Hydrogen Energy 2020, 45, 12308–12317. [Google Scholar] [CrossRef]
  189. Li, X.; Song, S.; Gao, Y.; Ge, L.; Song, W.; Ma, T.; Liu, J. Identification of the Charge Transfer Channel in Cobalt Encapsulated Hollow Nitrogen-Doped Carbon Matrix@CdS Heterostructure for Photocatalytic Hydrogen Evolution. Small 2021, 17, 2101315. [Google Scholar] [CrossRef]
  190. Zhang, Z.; Kang, Y.; Yin, L.-C.; Niu, P.; Zhen, C.; Chen, R.; Kang, X.; Wu, F.; Liu, G. Constructing CdSe QDs modified porous g-C3N4 heterostructures for visible light photocatalytic hydrogen production. J. Mater. Sci. Technol. 2021, 95, 167–171. [Google Scholar] [CrossRef]
  191. Wong, K.-L.; Bünzli, J.-C.G.; Tanner, P.A. Quantum yield and brightness. J. Lumin. 2020, 224, 117256. [Google Scholar] [CrossRef]
  192. Bouas-Laurent, H.; Dürr, H. Organic photochromism (IUPAC Technical Report) Republication or reproduction of this report or its storage and/or dissemination by electronic means is permitted without the need for formal IUPAC permission on condition that an acknowledgment, with full re. Photochromism 2003, 73, 639–665. [Google Scholar] [CrossRef]
  193. Meng, X.; Zhang, C.; Dong, C.; Sun, W.; Ji, D.; Ding, Y. Carbon quantum dots assisted strategy to synthesize Co@NC for boosting photocatalytic hydrogen evolution performance of CdS. Chem. Eng. J. 2020, 389, 124432. [Google Scholar] [CrossRef]
  194. Yuan, Y.; Lu, H.; Ji, Z.; Zhong, J.; Ding, M.; Chen, D.; Li, Y.; Tu, W.; Cao, D.; Yu, Z.; et al. Enhanced visible-light-induced hydrogen evolution from water in a noble-metal-free system catalyzed by ZnTCPP-MoS2/TiO2 assembly. Chem. Eng. J. 2015, 275, 8–16. [Google Scholar] [CrossRef]
Figure 1. Web-diagram showing the primary energy sources for the production of hydrogen and its uses.
Figure 1. Web-diagram showing the primary energy sources for the production of hydrogen and its uses.
Energies 14 08265 g001
Figure 2. Schematic illustration of synthesis of MoS2 by the hydrothermal method. Reproduced with permission from [55].
Figure 2. Schematic illustration of synthesis of MoS2 by the hydrothermal method. Reproduced with permission from [55].
Energies 14 08265 g002
Figure 3. Formation mechanism of SnS2 nanostructures during the solvothermal process. Reproduced with permission from [38].
Figure 3. Formation mechanism of SnS2 nanostructures during the solvothermal process. Reproduced with permission from [38].
Energies 14 08265 g003
Figure 4. Schematic diagram of the formation process of flower-like WS2 nanospheres using CTAB as aa surfactant. Reproduced with permission from [57].
Figure 4. Schematic diagram of the formation process of flower-like WS2 nanospheres using CTAB as aa surfactant. Reproduced with permission from [57].
Energies 14 08265 g004
Figure 5. Schematic illustration of the microwave-assisted synthesis of the MoS2/graphene composite. Reproduced with permission from [47].
Figure 5. Schematic illustration of the microwave-assisted synthesis of the MoS2/graphene composite. Reproduced with permission from [47].
Energies 14 08265 g005
Figure 6. Schematic illustration of the preparation of TiO2 NTAs, MoS2 QDs, and MoS2 QDs@TiO2 NTA heterostructures. Reproduced with permission from [27].
Figure 6. Schematic illustration of the preparation of TiO2 NTAs, MoS2 QDs, and MoS2 QDs@TiO2 NTA heterostructures. Reproduced with permission from [27].
Energies 14 08265 g006
Figure 7. Schematic diagram of the fabrication of the Au NP/ZnxCd1−xS NW heterostructure. Reproduced with permission from [75].
Figure 7. Schematic diagram of the fabrication of the Au NP/ZnxCd1−xS NW heterostructure. Reproduced with permission from [75].
Energies 14 08265 g007
Figure 8. (a) Experimental setup of the preparation of WO3 nanowires and WS2, WSe2, and WS2(1−x)Se2x nanotubes. (b) Schematic diagram of WO3 nanowires (NWs) converting to WS2 and WSe2 nanotubes (NTs) on carbon fibers by sulfidization and selenization. SEM images of (c) WO3 NWs and (d) WS2, (e) WSe2, and (f) WS2(1−x)Se2x NTs. Reproduced with permission from [39].
Figure 8. (a) Experimental setup of the preparation of WO3 nanowires and WS2, WSe2, and WS2(1−x)Se2x nanotubes. (b) Schematic diagram of WO3 nanowires (NWs) converting to WS2 and WSe2 nanotubes (NTs) on carbon fibers by sulfidization and selenization. SEM images of (c) WO3 NWs and (d) WS2, (e) WSe2, and (f) WS2(1−x)Se2x NTs. Reproduced with permission from [39].
Energies 14 08265 g008
Figure 9. Schematic diagram of SILAR growth. (a) Adsorption of K+X and the formation of electrical double layer, (b) first rinsing to remove excess, unabsorbed K+ and X, (c) reaction of A with pre-adsorbed K+ ions to form KA, and (d) second rinsing to remove excess and unreacted species and form the solid solution KA on the surface of the substrate. Reproduced with permission from [77].
Figure 9. Schematic diagram of SILAR growth. (a) Adsorption of K+X and the formation of electrical double layer, (b) first rinsing to remove excess, unabsorbed K+ and X, (c) reaction of A with pre-adsorbed K+ ions to form KA, and (d) second rinsing to remove excess and unreacted species and form the solid solution KA on the surface of the substrate. Reproduced with permission from [77].
Energies 14 08265 g009
Figure 10. Schematic experimental setup of the refluxing method. Reproduced with permission from [81].
Figure 10. Schematic experimental setup of the refluxing method. Reproduced with permission from [81].
Energies 14 08265 g010
Figure 11. Schematic illustration of the chemical vapor deposition system used for the synthesis of WS2. Reproduced with permission from [92].
Figure 11. Schematic illustration of the chemical vapor deposition system used for the synthesis of WS2. Reproduced with permission from [92].
Energies 14 08265 g011
Figure 12. The LCC method used to separate the nanosheets. Reproduced with permission from [98].
Figure 12. The LCC method used to separate the nanosheets. Reproduced with permission from [98].
Energies 14 08265 g012
Figure 13. XRD pattern of SnS2 nanosheets (NSs) prepared at 220 °C for 12 h. Reproduced with permission from [38].
Figure 13. XRD pattern of SnS2 nanosheets (NSs) prepared at 220 °C for 12 h. Reproduced with permission from [38].
Energies 14 08265 g013
Figure 14. XRD pattern for the samples obtained at different thermolysis temperatures: (a) MoS2 on carbon cloth and (b) WS2 on carbon cloth. Peaks with notation C represent the carbon peaks from carbon cloth. (*) indicates the peaks from WO3. Reproduced with permission from [99].
Figure 14. XRD pattern for the samples obtained at different thermolysis temperatures: (a) MoS2 on carbon cloth and (b) WS2 on carbon cloth. Peaks with notation C represent the carbon peaks from carbon cloth. (*) indicates the peaks from WO3. Reproduced with permission from [99].
Energies 14 08265 g014
Figure 15. XRD patterns of CN2.5, CN10, CN20, CM0.05, and M (MoS2). Reproduced with permission from [101].
Figure 15. XRD patterns of CN2.5, CN10, CN20, CM0.05, and M (MoS2). Reproduced with permission from [101].
Energies 14 08265 g015
Figure 16. XRD pattern of pure CoS2, pure TiO2, and 10 wt. % of the CoS2/TiO2 nanocomposite. Reproduced with permission from [105].
Figure 16. XRD pattern of pure CoS2, pure TiO2, and 10 wt. % of the CoS2/TiO2 nanocomposite. Reproduced with permission from [105].
Energies 14 08265 g016
Figure 17. XRD image for pure SnS2 (ST-100), pure TiO2 (ST-0), different wt. % of SnS2/TiO2 materials 5% (ST-5), 10% (ST-10), 15% (ST-15), and 20% (ST-20). Reproduced with permission from [106].
Figure 17. XRD image for pure SnS2 (ST-100), pure TiO2 (ST-0), different wt. % of SnS2/TiO2 materials 5% (ST-5), 10% (ST-10), 15% (ST-15), and 20% (ST-20). Reproduced with permission from [106].
Energies 14 08265 g017
Figure 18. SEM images of the SnS2 samples prepared at different reaction times. (a) 1 h; (b,c) 2 h; (d) 5 h; (e) 12 h; and (f) 24 h. Reproduced with permission from [38].
Figure 18. SEM images of the SnS2 samples prepared at different reaction times. (a) 1 h; (b,c) 2 h; (d) 5 h; (e) 12 h; and (f) 24 h. Reproduced with permission from [38].
Energies 14 08265 g018
Figure 19. (ac) Field emission scanning electron microscopy (FESEM) images of NiS2 hollow microspheres at different magnifications. (d) High-resolution transmission electron microscope (TEM) image of NiS2. Reproduced with permission from [13].
Figure 19. (ac) Field emission scanning electron microscopy (FESEM) images of NiS2 hollow microspheres at different magnifications. (d) High-resolution transmission electron microscope (TEM) image of NiS2. Reproduced with permission from [13].
Energies 14 08265 g019
Figure 20. SEM images of (a,b)—pure CoS2, (c,d)—pure TiO2, and (e,f)—CoS2/TiO2 nanocomposites. Reproduced with permission from [105].
Figure 20. SEM images of (a,b)—pure CoS2, (c,d)—pure TiO2, and (e,f)—CoS2/TiO2 nanocomposites. Reproduced with permission from [105].
Energies 14 08265 g020
Figure 21. SEM images for (a,b)—pure TiO2 and (c,d)—pure SnS2. Reproduced with permission from [106].
Figure 21. SEM images for (a,b)—pure TiO2 and (c,d)—pure SnS2. Reproduced with permission from [106].
Energies 14 08265 g021
Figure 22. SEM images for different wt. % of SnS2/TiO2 (a,e) 5%, (b,f) 10%, (c,g) 15%, and (d,h) 20%. nanocomposites. Reproduced with permission from [106].
Figure 22. SEM images for different wt. % of SnS2/TiO2 (a,e) 5%, (b,f) 10%, (c,g) 15%, and (d,h) 20%. nanocomposites. Reproduced with permission from [106].
Energies 14 08265 g022
Figure 23. Full XPS spectra of (a) MoS2/candle soot composite, (bd) high-resolution XPS spectra of Mo 3d, S 2p, and C 1s for MoS2/candle soot composite. Reproduced with permission from [108].
Figure 23. Full XPS spectra of (a) MoS2/candle soot composite, (bd) high-resolution XPS spectra of Mo 3d, S 2p, and C 1s for MoS2/candle soot composite. Reproduced with permission from [108].
Energies 14 08265 g023
Figure 24. (a) XPS of Pt-CdS and the corresponding high-resolution XPS spectra of Pt–CdS, (b) Cd 3d, and (c) S 2p. Reproduced with permission from [25].
Figure 24. (a) XPS of Pt-CdS and the corresponding high-resolution XPS spectra of Pt–CdS, (b) Cd 3d, and (c) S 2p. Reproduced with permission from [25].
Energies 14 08265 g024
Figure 25. UV–Visible absorption spectra of 3-D flower-like MoS2 and 3% LaMoS2 nanoflake samples. Reproduced with permission from [109].
Figure 25. UV–Visible absorption spectra of 3-D flower-like MoS2 and 3% LaMoS2 nanoflake samples. Reproduced with permission from [109].
Energies 14 08265 g025
Figure 26. (a) DRS spectra and (b) Tauc plots for SnS2 nanoparticles synthesized by using different reactants and solvents. Reproduced with permission from [111].
Figure 26. (a) DRS spectra and (b) Tauc plots for SnS2 nanoparticles synthesized by using different reactants and solvents. Reproduced with permission from [111].
Energies 14 08265 g026
Figure 27. Raman spectra measured with HeNe laser for the CoS2, NiS2 and (Co,Ni)S2 thin films and compared with those measured for standard CoS2 and NiS2 samples. Reproduced with permission from [112].
Figure 27. Raman spectra measured with HeNe laser for the CoS2, NiS2 and (Co,Ni)S2 thin films and compared with those measured for standard CoS2 and NiS2 samples. Reproduced with permission from [112].
Energies 14 08265 g027
Figure 28. Schematic illustration of the experimental setup of electrochemical water splitting.
Figure 28. Schematic illustration of the experimental setup of electrochemical water splitting.
Energies 14 08265 g028
Figure 29. Types of photocatalytic reactions. Reproduced with permission from [130].
Figure 29. Types of photocatalytic reactions. Reproduced with permission from [130].
Energies 14 08265 g029
Figure 30. Schematic illustration of the process involved during water splitting with the photocatalyst under the illumination of sunlight.
Figure 30. Schematic illustration of the process involved during water splitting with the photocatalyst under the illumination of sunlight.
Energies 14 08265 g030
Figure 31. Schematic diagram of the mechanism of the hydrogen evolution reduction (HER) process. Reproduced with permission from [182].
Figure 31. Schematic diagram of the mechanism of the hydrogen evolution reduction (HER) process. Reproduced with permission from [182].
Energies 14 08265 g031
Figure 32. Schematic diagram of the proposed reaction mechanism for H2 production using the La-MoS2 photocatalyst. Reproduced with permission from [109].
Figure 32. Schematic diagram of the proposed reaction mechanism for H2 production using the La-MoS2 photocatalyst. Reproduced with permission from [109].
Energies 14 08265 g032
Table 1. Summary of the different types of the synthesis method of TMC.
Table 1. Summary of the different types of the synthesis method of TMC.
Synthesis MethodsAdvantages/Disadvantages
Hydrothermal/Solvothermal MethodProduce nanomaterials with various morphologies and thicknesses at high pressure and low temperature; can synthesize high quality large crystals; cost of equipment is high
Microwave-assisted SynthesisRequired less time/rapid process; size can be controlled
Electrodeposition MethodRapid and single-step process; used to produce homogeneous and high-purity crystalline materials at the cathode of the electrochemical system during the coating process
PhotoreductionRequire higher photon energy; can synthesize the materials with large surface area and many active sites
Electron Beam Evaporation0D, 1D, and 2D materials can be prepared; used for depositing materials with high melting point; as electrons can be focalized, it is possible to obtain a very localized heating on the material to evaporate with a high density of evaporation power
Sulfidation and SelenizationSolution-phase conversion; facile and selectable synthesis method
Successive Ionic Layer Adsorption and Reaction (SILAR) MethodSimple, cost effective, and rapid technique for the deposition of binary semiconducting thin films
Refluxing MethodLarge scale synthesis method; facile and cost effective
Chemical Vapor Deposition (CVD) MethodGas-phase aerosol process for producing high-purity nanoparticles; mainly used for large scale thin-film production
Liquid cascade centrifugation (LCC) methodUsed to achieve highly efficient nanosheet with selected size and thickness; most efficient, scalable, and versatile method based on a set of iterative centrifugation cascades
Table 2. Different hydrogen production methods and its advantages and disadvantages.
Table 2. Different hydrogen production methods and its advantages and disadvantages.
Hydrogen Production MethodsDescriptionAdvantagesDisadvantagesReference
Steam natural gas reformingDirect production of hydrogen gas by conversion of feed hydrocarbons at high temperature
  • Most viable technology to begin building hydrogen market
  • Existing feedstock infrastructure
  • Capital investment is high
  • High operation and maintenance expenses
  • CO2 byproduct
  • Dependence on fossil fuels
[121,122,123]
Coal gasificationPartial oxidation of coal with air. However, it generates the CO2 through traditional combustion
  • Low-cost
  • Uses abundant and affordable coal feedstock
  • High reactor costs
  • System efficiency
  • Feedstock impurities
  • Carbon capture and storage
[123,124]
Biomass gasificationBurning of biomass using limited supply of air generates combustible gases, such as CO, CO2, H2 CH4, H2O and N2 with contaminants like small char particles, ash and tars
  • CO2-neutral, abundant and cheap feedstock
  • Tar formation
  • Varying H2 content due to seasonal
  • Availability and feed stock impurities
[122,125]
ThermolysisUsing high temperature from concentrated solar energy and chemical reactions to generate the H2 and O2 from water.
  • Clean and Sustainable, abundant feedstock
  • O2 is the only byproduct; No greenhouse gas emissions
  • Element’s toxicity
  • Corrosive problems
  • High capital costs
[122,123]
ElectrolysisUsage of electricity to split water into hydrogen and oxygen
  • Emission-free, proven technology, existing infrastructure
  • Abundant feedstock
  • O2 is the only byproduct
  • Contributes to electricity storage option.
  • Low overall efficiency
  • High capital costs
[114,122]
PhotoelectrochemicalHydrogen is produced from water in the presence of sunlight and specialized semiconductors immersed in water-based electrolytes
  • Abundant feedstock
  • O2 is the only byproduct
  • Low conversion efficiency
  • Non-effective photocatalytic material
[10,114,126]
PhotocatalyticA simple method to produce hydrogen from water in the presence of catalyst and sunlight
  • Photocatalyst material directly suspended in to the water
  • Simple/facile method
  • Low yield
  • Separation of gases
  • Selection of suitable catalyst
  • Separation of suspended catalysts
[10,127,128]
Table 3. Comparison of electrochemical hydrogen evolution by using different TMC photocatalysts.
Table 3. Comparison of electrochemical hydrogen evolution by using different TMC photocatalysts.
MaterialCurrent DensityOver PotentialPhotocurrentOther Condition(s)References
MoS210 mAcm−2173 mV109.81 mVdec−1Nafion resin[140]
MoSe210 mAcm−2208 mV65.92 mVdec−1Nafion resin[140]
MoTe210 mAcm−2283 mV102.06 mVdec−1Nafion resin[140]
MoS2/candle soot/Ni foam10 mAcm−256 mVNA1.0 M KOH[108]
3D/2D TiO2/MoSe21.40 mAcm−2NANA1.0 M NaOH[141]
2D MoS2/MoSe210 mA/cm−2
147 μmol of H2 in ∼50 min.
NANAWide pH conditions[142]
Amorphous-RuS210 mAcm−2141 mV65.6 mVdec−10.5 M H2SO4[143]
MoS2/TiO210 mAcm−2170 mV70 mVdec−10.5 M H2SO4[118]
GO/NiS210 mAcm−257 mV-HER
294 mV-OER
NA1.0 M KOH[144]
PU@PANI@FeS210 mAcm−2
50 mAcm−2
266 mV
372 mV
NA0.5M H2SO4[145]
CuxSNA–90 mV100 mVdec−10.5 M H2SO4[146]
CoS/MoS2NA−147 mV126 mVdec−1 [147]
Carbon/Co3S410 mVcm−2250 mV and 140 mV for OER and HERNA1.0 M KOH[148]
NiWSex and CoWSex1 mAcm−2NiWSex (231 mV) CoWSex (281 mV) and WSex (404 mV)NA1.0 M KOH[149]
MoS2/Pt0.642 mAcm−2 NA52 mVdec−1 to 32 mVdec−10.5 M H2S O4[150]
G/MoS210 mAcm−2118 mV73 mVedec−10.5 M H2SO4[151]
NiS/Pd10 mAcm−2100 mV50 mVdec−1NA[152]
ZnxCo1−xSe210 mAcm−2196 mV-HER
308 mV-OER
NAHER—acid media
OER—alkaline media
[153]
Mo/CoSe100 mAcm−2186.1 mV58.7 mV dec−10.5 M H2SO4[154]
MoS2/graphene1000 mAcm−2250 mV43.3 mVdec−10.5 M
H2SO4
[47]
NiS210 mA/cm−2302 mVNANafion ethanol solution[155]
GaSe−9.3 µAcm−2 83.4 µAcm−2 +1.23 V vs. RHE NA0.5 M H2SO4[156]
1T-MoS210 mAcm−2240 mV68 mVdec−11 M KOH[157]
CoSe/MoSe210 mAcm−2192 and 115 mV in acidic and alkalineNANafion ethanol solution[158]
NA—Not Available.
Table 4. Photocatalytic hydrogen evolution by using different TMC photocatalysts, the experimental conditions used, and the respective references.
Table 4. Photocatalytic hydrogen evolution by using different TMC photocatalysts, the experimental conditions used, and the respective references.
MaterialsAmount of Hydrogen EvolvedLight SourceCo-CatalystScavenging Agents/Other Chemicals UsedReference
SnS21.06 mmol h−1 g−1UV–VisibleNANa2S and Na2SO3[38]
Te/SnS2/Ag332.4 and
166.2 µmol h−1 for H2 and O2
UV–VisibleNA-[168]
SnS2/g-C3N4972.6 µmol h−1g−1VisiblePtTriethanolamine[169]
CdS/WS20.42 mmol h−1VisibleNALactic acid[166]
NiS /CdS DETA230 µmolh−1VisibleNA-[170]
CdS/ZnO/GO,
CdS/Al2O3/GO
22.12 mmol h−1g−1500 W Phoenix tungsten halogen lampPtNa2SO3 and Na2S[171]
MoS2/RGO/CdS99 μmol h−1VisibleNALactic acid[42]
MoS2/G1.80 mmol h−1VisibleNANa2S and Na2S2O3[172]
Eosin Y/NiSx/G0.34 mmol h−1VisibleNATriethanolamine[22]
NiS/CdS/ZnS574 μmol h−1VisibleNANa2S/Na2S2O3[173]
MoS2110 µmol h−1VisibleLa3+Methanol[109]
CuS–MoS2–1T9648.7 μmol g −1h−1UV–VisibleNA0.3 M Na2SO3[174]
WS2/TiO2596.4 µmolg−1VisiblePt1 wt. % of WS2/TiO2[31]
ZnIn2S4/g-C3N4450 μmol g−1h−1VisibleNATriethanolamine[175]
WSe2/ Zn0.1Cd0.9S147.32 mmol g−1h−1VisibleNALactic acid[176]
NiCo/Zn0.5Cd0.5S34.7 mmol g−1h−1UV–VisibleNANa2S and Na2SO3[177]
NiSe2/RP1968.8μmol g−1h−1UV–VisibleNANa2S and Na2SO3[178]
CdS3072 μmol g−1h−1UV–VisibleNANa2S and Na2SO3[179]
ZnS/CdS/Cd0.5Zn0.5S/MoS250.65 mmol g−1h−1VisibleNALactic acid, Na2S, and Na2SO3[117]
CdxMo1−xSe911.1 mol in 7 hVisibleNANa2S and Na2SO3[180]
MoS2/CdIn2S41868.19 μmol g−1h−1UV–VisibleNANa2S and Na2SO3[181]
MoS2/CoSe2/1D-CdS191.5 mmol g−1h−1UV–VisibleNALactic acid[182]
NiS/CQDs/ZnIn2S428.2 μmol h−1VisibleNATrimethylamine[183]
NiSe2/CdS167.1 mmol g−1h−1VisibleNA [184]
NiS/ZnIn2S45.0 μmol h−1VisibleNAMethanol and lactic acid[185]
1T-MoS2-ZnCoS15.47 mmol h−1 g−1UV–VisibleNATrimethylamine, Eosin Y, and acetonitrile[120]
CdS/SnS220.2 mmol h−1 g−1UV–VisibleNALactic acid[186]
ZnO−ZnS−Cu2S436 μmol h−1 g−1VisibleNANa2S and Na2SO3[187]
ZnS films5202.4 μmol h−1 g−1UVNANa2SO3[188]
CoS2/TiO22.55 mmol g−1UVNAMethanol[105]
SnS2/TiO2195.0 μmol g−1VisibleNAMethanol[106]
(NC@Co-NCT)/(CdS)3.8 mmol h−1 g−1VisibleNAlactic acid[189]
CdSe QDs/g-C3N4192.3 μmol h−1VisibleH2PtCl6Triethanolamine,
Na2S, and Na2SO3
[190]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shanmugaratnam, S.; Yogenthiran, E.; Koodali, R.; Ravirajan, P.; Velauthapillai, D.; Shivatharsiny, Y. Recent Progress and Approaches on Transition Metal Chalcogenides for Hydrogen Production. Energies 2021, 14, 8265. https://0-doi-org.brum.beds.ac.uk/10.3390/en14248265

AMA Style

Shanmugaratnam S, Yogenthiran E, Koodali R, Ravirajan P, Velauthapillai D, Shivatharsiny Y. Recent Progress and Approaches on Transition Metal Chalcogenides for Hydrogen Production. Energies. 2021; 14(24):8265. https://0-doi-org.brum.beds.ac.uk/10.3390/en14248265

Chicago/Turabian Style

Shanmugaratnam, Sivagowri, Elilan Yogenthiran, Ranjit Koodali, Punniamoorthy Ravirajan, Dhayalan Velauthapillai, and Yohi Shivatharsiny. 2021. "Recent Progress and Approaches on Transition Metal Chalcogenides for Hydrogen Production" Energies 14, no. 24: 8265. https://0-doi-org.brum.beds.ac.uk/10.3390/en14248265

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop