Next Article in Journal
Mechanistic Insights into Graphene Oxide Driven Photocatalysis as Co-Catalyst and Sole Catalyst in Degradation of Organic Dye Pollutants
Next Article in Special Issue
Plasmon-Induced Semiconductor-Based Photo-Thermal Catalysis: Fundamentals, Critical Aspects, Design, and Applications
Previous Article in Journal
Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structural, Morphological and Optical Properties of MoS2-Based Materials for Photocatalytic Degradation of Organic Dye

by
Jadan Resnik Jaleel UC
1,
Madhushree R
1,
Sunaja Devi K R
1,*,
Dephan Pinheiro
1 and
Mothi Krishna Mohan
2
1
Department of Chemistry, CHRIST (Deemed to be University), Bangalore 560029, India
2
Department of Sciences and Humanities, School of Engineering and Technology, CHRIST (Deemed to be University), Kumbalagodu, Mysore Road, Bangalore 560074, India
*
Author to whom correspondence should be addressed.
Photochem 2022, 2(3), 628-650; https://doi.org/10.3390/photochem2030042
Submission received: 4 July 2022 / Revised: 3 August 2022 / Accepted: 4 August 2022 / Published: 8 August 2022
(This article belongs to the Special Issue Advance in Photocatalysis in Asia)

Abstract

:
Molybdenum disulfide (MoS2) is a transition metal dichalcogenide (TMDCs) having versatile properties and plays a great role in the photodegradation of organic dyes. MoS2 also finds applications in diverse fields such as catalysis, electronics, and nanomedicine transportation. MoS2 can be prepared by using chemical and physical methods such as hydrothermal, solvothermal, and chemical vapour deposition methods. The preparation method employed can produce subtle but significant changes in the morphology. To increase the efficiency of MoS2, it can be combined with different materials to produce composites that improve the photodegradation efficiency of MoS2. The various methods of preparation, the morphology of MoS2, and photodegradation activity of the MoS2-based nanocomposites are briefly discussed in this review.

1. Introduction

The discovery of Buckminster fullerene (C60) and carbon nanotubes (CNTs) in the late twentieth century marked a pivotal moment in nanomaterial research. These materials have since become a substitute for carbon-based bulk materials such as graphite and diamond. These materials have already made their presence felt and proved their utility in diverse fields [1]. A significant amount of studies have been done on the carbon nanotube as a versatile material in electronics and numerous other fields [2]. However, the inherent drawbacks in carbon nanotubes led to the discovery of materials like graphene [3], two-dimensional transition metal dichalcogenides (TMDCs), etc. The TMDCs are composed of a transition metal element (M) and chalcogen element (X) like S, Se, Te, etc. [4]. The layers are held together by the weak van der Waals forces. MoS2 is a TMDCs having single-layer/monolayer or nanolayers formed from bulk material via exfoliation. [5]. There are about 40 different types of TMDCs [6], such as WSe2, MoS2, WS2, etc. They are prepared by sandwiching a layer of transition metal (e.g., molybdenum, tungsten, niobium) atoms between two layers of chalcogen (e.g., sulphur, selenium, tellurium) atoms. MoS2 is three atoms thick with a structure comprising of (X-M-X) S-Mo-S [7,8]. There are four different polytypes of MoS2, namely, 1T, 1H, 2H, and 3R. These are classified based on the coordination of Mo atom and stacking orientation of the single layers [6]. MoS2 has found extensive applications in photocatalysis and electrocatalysis [9,10]. MoS2 and MoS2-based nanomaterials have attracted a lot of attention for their high charge carrier potency and superb optical absorption property [11,12]. For commercial photocatalysis applications, a stable, efficient, and low-cost photocatalyst with a light-harvesting spectrum spanning the entire solar spectrum is highly desirable [13]. Photocatalytic materials can transform sunlight into chemical energy and use this to degrade organic pollutants into non-toxic carbon dioxide and tiny inorganic molecules like water without causing secondary pollution [14].
The poor interactions between the layers enable the formation of bulk equivalents of multi-layered materials, which support the stability of the structure. The interesting properties of TMDCs such as tuneable band gap ranging between 0–3 eV, high surface-to-volume ratio, semiconductor property, and the high availability of active sites on the surface have led to a wide range of applications such as gas sensing, optoelectronics, catalysis, energy storage, etc. [15]. The physical properties of these materials are amenable to alteration by inter-lamellar space intercalating or doping [5]. MoS2 has a role in many fields such as catalysis, electronics [16,17,18], and nanomedicine transportation [19]. g-C3N4 is another material frequently utilized to alter the characteristics of MoS2 in order to boost its catalytic efficiency [20].
The characteristics of single-layered MoS2 differ significantly from those of their bulk counterparts. The large surface energy in MoS2 makes it prone to aggregation, which can adversely impact its stability and efficiency as a catalyst [21,22]. This can be a serious drawback when MoS2-based catalysts are used in the aqueous phase. This can be overcome by embedding MoS2 within layered materials to achieve benefits to both materials, and improved performances can be realized [23,24]. MoS2 is a photocatalyst for hydrogen evolution because it has a strong conductivity, wide basic surface region, sufficient band edge, solid lattice fit, and relatively high mobility. Layered MoS2 is thought to be a good candidate for pairing with g-C3N4 because of its p-type conductivity. g-C3N4 is also an important photocatalytic substance for fuel generation from sunlight via water splitting and environmental remediation via organic pollutant degradation. ZnO, and CdS, which are also effective photocatalysts, have been used to alter the MoS2 catalyst to improve its photocatalytic efficiency. In recent studies, MoS2 has been widely explored in fields other than photocatalytic degradation. One major application that was studied is towards H2 production [25]. MoS2/Ni3S2/reduced graphene oxide electrocatalyst was studied for alcohol fuel cells [26], and on bacterial wound treatment [27]. The bacterial capturing and deactivating using MoS2-based nanocomposites have also been studied.
Access to clean freshwater is a necessity for all human beings and, to a large extent, determines the quality of our lives. Pollution of the water bodies, particularly with industrial wastewater containing a variety of organic and inorganic pollutants, is an environmental issue worldwide [28]. According to the 2017 United Nations World Water Development report, 80% of wastewater is dumped into the environment without proper treatment [29]. The availability of various pollutants in water has serious health consequences for human health as well [30]. Further, the exploding population has placed an additional pressure on our diminishing freshwater resources [31]. Rapid industrialization and increasing agricultural activities have led to the generation of a large amount of wastewater [32]. Pollution of water also has adverse effects on the aquatic ecosystem. There is an urgent need to address this sorry state of affairs. Remediating wastewater through economic and eco-friendly protocols needs to be devised. One such technology for wastewater remediation is photocatalysis, which makes use of freely accessible sunlight to fuel the breakdown of organic contaminants present in wastewater [33]. Photocatalysis has risen in popularity in recent years because of its huge potential for resolving the growing energy crisis and environmental pollution. At moderate temperatures, solvothermal techniques can produce acceptable yields with varied controllability in the size of nanomaterials, which are successfully used in a variety of semiconductor applications [34].
This article briefly discusses the structure, geometry, methods of preparation, characterization, and application of MoS2-based nanomaterials. The research achievements made in improving the degradation performance of MoS2-based photocatalysts have been explored. The schematic illustration of the MoS2-based catalysts’ preparation, morphology, and applications are given in Figure 1.

2. Structure and Geometries of MoS2

MoS2 is a black-coloured substance, insoluble in water, and like graphene, has a layered structure. It is mostly found in nature in the form of molybdenite MoS2 has excellent thermal and chemical stability, which is a feature of layered transition metal compounds in general. As a result, it is employed in nanochemistry, catalysis, electrode materials, pharmaceuticals, nanomedical transportation, etc. MoS2 is a typical n-type semiconductor with a layered structure that is quite similar to graphene. The band gap of MoS2 grows as the number of the atomic layers decreases, resulting in desirable photoelectric characteristics. MoS2 is a stable 2D transition metal dichalcogenide, which means it is a semiconductor of the MX2 type, where M stands for transition metals and X is for chalcogen [29]. A single sheet is similar to a “sandwich” structure, where the Mo atom is sandwiched between two S atoms [35,36]. MoS2 layers, separated by 0.65 nm, are held together by van der Waals forces, and the layers are bound together by strong covalent bonds [37]. Figure 2a,b show the crystal structure of MoS2 [16,38,39].
MoS2 is polytypic [4], with three distinct configurations: 1T [40], 2H [41], and 3R [42], all of which belong to symmetric point groups D6h, D6d, and C3V, respectively [13]. The polymorphic structures are given in Figure 2c. The first digit in these polytypes denotes the number of layers in the make-up, while the alphabet denotes the crystallographic configuration. In these polytypes, the letters T, H, and R stand for trigonal, hexagonal, and rhombohedral arrangements. Various allotropes and morphologies, such as 3D (flowers, snowflakes, and dandelion patterns), 2D (nanosheets, nanostrips, and nanoribbons structure), 1D (nanowires and nanorods arrangement), and 0D (nanowires and nanorods) can be visualized, designed, and achieved by suitably altering the synthesis process. MoS2 is a naturally occurring 2H molecule with 3% 3R that is thermodynamically stable.
MoS2 monolayers show less activity in catalysis due to the limited charge mobility and a lack of metal edge sites. Increasing the number of exposed edge sites and using monolayered MoS2 sheets having greater electrical conductivity have both improved the catalytic performances of MoS2. However, the bulk of the metal on the basal plane of the monolayers still shows a lower catalytic activity [43]. From the studies conducted, it is evident that while the basal planes of 2H-MoS2 are catalytically inert, the sulphur edge sites, and vacancy defects are the active sites for dye degradation. The photocatalytic dye degradation shows an enhancement when the size of the 2H-MoS2 particles has been decreased to produce a higher density of edge sites. Another method for improving the catalytic activity of MoS2 is to alter the 2H crystal phase to produce a polymorph with improved electrical conductivity, which will facilitate the transport of electrons to the active sites [44,45]. As stated below, MoS2 comes in a variety of crystalline forms. The images given below are from the Materials projects database [46].
Hexagonal: The hexagonal structure of the MoS2 is given in Figure 3. The structure of P63/mmc space group is given the Figure 3a. It shows a band gap of 1.465 eV and a density of 4.05 g/cm3. Lattice parameters of the structure are a = b = 3.190 Å, and c = 14.879 Å. α = β = 90° γ = 120° and volume is 131.151 Å3. The structure of P6m2 is given in Figure 3b. It has a band gap of 1.661 eV and a density of 2.80 g/cm3. Lattice parameters of the structure are a = b = 3.190 Å and c = 17.440 Å. α = β = 90° γ = 120° and volume is 153.721 Å3 [46].
Trigonal: The trigonal structure of the MoS2 is given in Figure 4. The structure of the R3m space group is given in Figure 4a. It shows a band gap of 1.578 eV and a density of 4.24 g/cm3. Lattice parameters of the structure are a = b = c = 7.341 Å. α = β = γ = 25.103°, and volume is 62.649 Å3. The structure of MoS2 with the P3m1 space group is given in Figure 4b. It shows a band gap of 1.554 eV and a density of 2.42 g/cm3. Lattice parameters of the structure are a = b = 3.190 c = 24.879 Å. A = β = 90° γ = 120°. Volume is 131.151 Å3. The structure of the P6m2 space group is given in Figure 4c, having a band gap of 1.509 eV and a density of 2.8 g/cm3. Lattice parameters of the structure are a = b = 3.190 c = 32.319 Å. α = β = 90° γ = 120° and volume is 284.872 Å3.
Other geometries: The orthorhombic structure of the MoS2 is given in Figure 5a. It has a band gap of 1.578 eV and a density of 2.67 g/cm3. The structure belongs to the Pmmn space group. Lattice parameters of the structure are a = 3.163 b = 13.021 c = 43.774 Å. α = β = γ = 90°, volume = 1790.583 Å3. Figure 5b represents the cubic geometry of MoS2. This structure has a band gap of 1.562 eV and a density of 4.81 g/cm3. The structure belongs to the F43m space group. Lattice parameters of the structure are a = b = c = 6.778 Å. α = β = γ = 60°, and volume = 221.129 Å3. Figure 5c represents the tetragonal structure of MoS2, and it has a band gap of 1.785 eV and a density of 3.27 g/cm3. The structure belongs to the I42d space group. Lattice parameters of the above structure are a = b = c = 6.505 Å. α = β = 128.852°, γ = 75.248°, and volume = 162.514Å3 [46].

3. Different Methods of Preparation of MoS2

For the preparation of MoS2, a variety of physical and chemical approaches have been used. The generally used physical methods are (a) mechanical exfoliation, (b) sputtering, (c) epitaxy, (d) plasma, etc. The lattice structure of the material will not change or be destroyed in this method, provided the precursor materials are pure [38]. However, in recent studies, physical methods are not commonly used since they are not cost effective. The chemical methods include (a) chemical vapor deposition method, (b) hydrothermal method, and (c) solvothermal method. The most common methods of preparation are listed in Figure 6.

3.1. Physical Methods

Mechanical exfoliation: Mechanical exfoliation is an approach used in the preparation of 2D materials, similar to the method used in monolayer graphene. Yu et al., developed a simple, environmentally friendly, and scalable exfoliation approach for water-dispersible MoS2 [47]. In the pre-treatment of MoS2 powder, no intercalation agents have been used. Then, dispersion took place in ultrasound-assisted water—a typical exfoliation process for nanosheets—in combination with viscoelastic stinging for the preparation of MoS2. Funke et al. used an micromechanical exfoliation method where they did not find any evident grid defects in elliptic-polarization spectra imaging [48]. This method is useful to manage the materials’ surface and mechanical exfoliation. One drawback here is that the thickness and transverse dimensions are difficult to regulate and, therefore, the procedure is unsatisfactory and unsuitable for commercial manufacturing [49].

3.2. Chemical Methods

The chemical vapor deposition (CVD) method: This method can be used to prepare MoS2 and transition metal dichalcogenides of high quality on a large scale, mainly to make MoS2 films [50,51]. It is a bottom-up approach that preserves the defects, crystallinity, and morphology of the material [52]. CVD is a large-scale chemical reaction that involves vapors reacting with the substrate to produce thin films. Direct evaporation is used in the CVD process, commonly known as the vapour solid development technique. It creates a high-quality monolayer with fewer tiny flakes on the substrate. Precursors are charged and sublimated into a gaseous state which helps to undergo chemical reactions at high temperatures. Strong sediments grow on the matrix surface during condensation; hence it is commonly used to prepare monolayer nanocomposite systems and multilayer transition metal MoS2 with good control on the number of layers.
MoO3 powder is used as a molybdenum source and S powder as a sulfur source in the preparation of large-area monolayer MoS2 films. Monolayer films grow consistently with a uniform nature, but defects created during the growth process cause a reduction in the material consistency [53]. In the preparation, improvements were made by Zhang et al., who produced triangular monolayers of MoS2 with grain sizes up to 150 µm [54]. Alharbi et al. used a chemical vapor deposition process to create a large-area MoS2 film with the highest island density in the field [55]. Balendhran et al. developed a two-step bulk fabrication technique using an MoO3 precursor [56]. In a quartz channel, MoO3 is evaporated on every substrate, followed by sulfidation. A schematic representation of synthesizing MoS2 using the chemical vapour method and CVD growth of the MoS2 when Mo foil is used as a precursor is depicted in Figure 7a,b, respectively [57,58].
Hydrothermal method: The hydrothermal method is a promising wet chemical method that is performed using a hydrothermal bomb. The temperature used for this method is usually 180 °C [59]. Typically, the precursors used are sodium molybdate (Na2MoO4·2H2O) and ammonium molybdate tetrahydrate ((NH4)6Mo7O24⋅4H2O) for molybdenum and thiourea (CS(NH2)2) for sulfur. Nanosheets [60], nanospheres [61], nanoflowers [62], and nanotubes [63] can be prepared using hydrothermal methods. It is the most frequently used method for synthesis of flower-like MoS2 [64]. Figure 8 represents the preparation steps of MoS2 nanosheets. The molybdenum source was Na2MoO4·2H2O, and the sulfur source was thiourea in the synthesis proposed by Zou et al. [65]. The MoS2-RGO-3 was generated using a hydrothermal process and had a structure that resembles a cabbage, with distinct lattice fringes and a diameter range of 500 nm−1. 3D MoS2 ultrastructure was prepared by Anwer et al. [66], who used a controllable hydrothermal approach to achieve their results. These micro-sized marigold flower-like patterns were made up of 2D ultrathin MoS2 nanosheets that were randomly spaced but closely connected. Controlling the input concentrations of thiourea and the precursor of MoS2 resulted in the MoS2-microflower composition. Tong et al. [67] have used a previously published hydrothermal method for MoS2 preparation. The flower-like MoS2 had a hierarchical structure with a diameter of 500 nm and was subjected to an 18 h hydrothermal reaction at 220 °C. The MoS2 nanoflowers made by the hydrothermal process were composed of uniform crystallinity and size with minimum agglomeration. Zhou et al. [68] have replaced the sulfur source with thioacetamide in their studies. Xia et al., in their recent study, synthesized the CdS@MoS2 nanocomposite using the hydrothermal method, and evaluated the photocatalytic degradation of Rhodamine 6G (Rh 6G), and found that adsorption in the nanocomposite increases when the MoS2 amount is increased, showing a great efficacy in processing Rh 6G [69]. Jian et al. used a two-step hydrothermal method to prepare the nanocomposite MoS2/BiOBr, which is very efficient in Rhodamine B (RhB) dye degradation [70].
Additionally, the regulating parameters under hydrothermal parameters are: (a) the percentage of components, (b) the reaction temperature, (c) the solution, (d) the pH value, and (e) the solution concentration. These variables might be effectively controlled to monitor the chemical reaction and the shape of the substance prepared. MoS2 shows enhanced photothermal properties when synthesized using other processes such as hydrothermal process, a photo-deposition technique, and an in situ solid-state chemical reduction approach. Moreover, at low temperatures, the Ag/MoS2/TiO2-x composite had outstanding degradation properties. Its hydrogen production rate was considerably higher than that of pristine TiO2 [71].
Solvothermal method: The solvothermal and hydrothermal methods of syntheses are similar techniques [72]. An organic solvent is used in the latter technique instead of water. Thus, where hydrothermal method cannot be applied for compounds that are very easily hydrolyzed, the solvothermal method can be used. In the solvothermal process, non-aqueous solvents are utilized as pressure carriers, intermediates, and mineralizers. Simultaneously, several non-aqueous solvents with a variety of characteristics can be utilized, resulting in nanomaterials with a diverse set of properties. Solvothermal processes can be used to make core-shell composites. Bai and collaborators generated a Co9S8@MoS2 core-shell heterojunction via a solvothermal method, which enhanced the catalytically active sites [73]. Zhang et al. used a basic solvent thermal approach to make MoS2/carbon nanotubes with a core-shell structure having improved optical properties [74]. To build Fe3O4@SiO2@MoS2/g-C3N4 (FSMG), Lu et al. employed a unique and efficient method where the Fe3O4 spheres formed the inner core of the SiO2 shell—placed using the sol-gel process—and were made using the solvothermal method. FSMG demonstrated good degradation behaviour against RhB when exposed to visible light. Importantly, by adjusting the proportion of water in the solvent, the size of the gel can be easily managed. Gao et al. have reported a solvothermal preparation by mixing MoS2 from ammonium tetrathiomolybdate with dimethylformamide (DMF) and hydrazine [75]. The mixture was then autoclaved for 10 h with 200 °C. The preparation of MoS2 aggregate is given in Figure 9.
Other methods: The methods discussed below are some of the most novel ones, namely, (a) thermal-based sulfurization process, (b) ultrasonic-assisted homogenous magnetic stirring method, (c) ultrasound-assisted cracking system, (d) thermal evaporation technique-based systems, (e) in situ growth method, (f) hydrolysis method, and (g) dual-template approach. Hollow structures may be developed using the dual-template approach, where the method increases the active sites and characteristics of produced MoS2 nanosheets [76]. Thermal sulfurization has been used to create homogeneous MoS2 films [77]. This approach does not require any pre-processing, unlike chemical vapour deposition (CVD). Acid treatment and ultrasonic irradiation are required before the ultrasound-assisted cracking procedure in order to obtain ideal materials [78]. Hydrolysis is a process for preparing nanoparticles in which precursor molecules are appropriately hydrolyzed in an aqueous solution environment under specific conditions. Zhu et al. reported the synthesis of MoS2 nanosheets of uniform thickness using a quick and fast LiBH4 hydrolysis reaction [79].

4. Characterization of MoS2-Based Materials

4.1. Morphological Properties

The shape of MoS2-based photocatalysts affects their efficiency. Depending on the technique of preparation, MoS2 can have a variety of morphologies. Jaleel et al. created MoS2 with a nanoflower shape utilizing a hydrothermal procedure using ammonium molybdate as the Mo source and thiourea as the S source [59]. This nanocomposite was used for degrading the malachite green dye with great efficiency. The SEM image of the nanoflower MoS2 is given in Figure 10a [59]. Sun et al. used the CVD process to create nanosheets of MoS2 and the hydrothermal intercalation method to construct nanospheres and used them for the adsorption studies. Figure 10b represents the nanosphere structure of MoS2 [80]. Visic et al. and co-workers prepared MoS2 coaxial nanotubes by the sulfurization transition of M6S2I8 nanowires in an argon environment with a gas flow of H2/H2S mixture. They proposed this MoS2 as an analogous material to graphene. Figure 10c represents the nano-wired MoS2 [81]. Mahyavanshi et al. used a chemical vapour deposition technique to illustrate the directional growth of MoS2 monolayer ribbons (Figure 10e) in a sulfur-rich environment. An array of molybdenum and sulphur edge terminations with 60° and 120° angle formations are present in this nanocomposite, as evidenced from the SEM and TEM images. There may be new opportunities for electrical and electrochemical applications due to the directed development of MoS2 ribbons with specified edge structures under specific CVD conditions [82]. The MoS2 nanoflakes were synthesized by Wu et al., (Figure 10f) via effective grinding-assisted sonication exfoliation for photoluminescence studies of MoS2 [83]. SEM images of nanoflower, nanosheet, nanosphere, nanotube, nanoribbon, and nanoflake morphologies are depicted in Figure 10a–f, respectively [83]. The HRTEM images in Figure 11a,b shows a close interface between the three components, which could assist with electron transfer to the composite surface [84,85].

4.2. Structural Properties

The various characterization techniques used to determine the structure of MoS2 are discussed here [59,81,86,87,88].
X-ray diffraction (XRD) has a significant role in identifying the crystalline nature of the material in nanocomposite studies. MoS2 shows different peaks in the XRD as shown in Figure 12a. The major plane obtained is (002) at 14.38° [89]. This corresponds to about eight layers of MoS2. Besides this major peak, MoS2 also displays peaks that correspond to (100), (101) at 32°, (103) at 40°, (105) at 49°, (110) at 58.5°, (008), and (200) planes [81]. There is no peak corresponding to (001) plane as there is no MoO3 [90]. Infrared spectral analysis (FTIR) is the other major characterization tool used to identify the functional groups present in the nanomaterial. As shown in Figure 12b, the FTIR plot of MoS2 shows various peaks. The peak at 3182 cm−1 is due to the O−H group, while 639 cm−1, 893.39 cm−1, 1402.99 cm−1, and 1622.8 cm−1, are the broad absorption bands attributed to MoS2. And the band at 483.23 cm−1 and 931.39 cm−1 are due to S−S bond [88].
Thermogravimetric analysis (TGA) on the MoS2 shows the thermal stability of the material. The TGA plot of MoS2 is given in Figure 12. In Figure 12c,d, we can see a similar pattern of thermal decomposition of the MoS2. Bahuguna et al. [87] synthesized MoS2 using a hydrothermal process and then carried out TGA analysis. MoS2 shows an initial loss of mass at about 100 °C, which is caused by the evaporation of adsorbed water molecules, as seen in Figure 12c. Further, MoS2 loses mass at around 300 °C, due to the decomposition of sulfur from the compound, and the MoS2 oxidizes to MoO3, after which it remains stable up to 1000 °C. Figure 12d(i) shows the TGA analysis of the MoS2 and its modified system [59].

4.3. Optical Properties

The optical properties are usually characterized by two different characterizing tools (a) photoluminescence (PL) and (b) UV-diffuse reflectance spectroscopy (UV-DRS) study. From the UV-DRS analysis, and Tauc plot, the band gap is calculated using Kubelka–Munk function, which is given in Figure 12e,f. The PL spectrum has an important role in proposing a mechanism in organic dye degradation using nanocomposites. The band gap of MoS2 using Tauc plot has been calculated for MoS2 as 3.26 eV [86]. Zhang et al. prepared a ternary system of MoS2/g-C3N4/TiO2 and calculated the band gap as 2.76 eV. Chakrabarty et al. determined the band gap of RGO-MoS2 supported NiCo2O4 as 2.36 eV [91]. Nayak et al. prepared MoS2/NiFe layered double hydroxide (LDH), and using Tauc calculation, found 1.86 eV as the band gap [92]. In the modified MoS2 system, the UV-DRS of the composites displayed a red shift towards the visible light spectrum [93].

5. MoS2-Based System for Degradation of Organic Dyes

Because of its strong optical absorption, MoS2 has a wide range of applications in photochemistry, photocatalysis, and photoelectronic research. In electronics, the MoS2-based materials are used as photodetectors [94], field-effect transistors [95], solar cells [96], chemical sensors [97], and so on. MoS2 is also used in the transportation of drugs in cancer chemotherapy. Major applications of MoS2 and MoS2-based systems in various photocatalytic degradation reactions are discussed here.

5.1. Binary Systems of MoS2

Graphitic carbon nitride (g-C3N4) has been thoroughly explored for its small band gap (2.7 eV), nonmetal characteristics, nontoxicity, ease of availability, and excellent thermal and chemical stability since its debut in 2009 for photocatalytic hydrogen evolution. Many studies have been done by fabricating MoS2 on g-C3N4. Zhang et al. [98] prepared a nanocomposite by fabricating MoS2 and g-C3N4. Here, MoS2 was prepared by the hydrothermal method. The photocatalytic efficiency of the samples was assessed by measuring their ability to degrade the organic RhB and methyl orange (MO) dyes in aqueous solutions when exposed to visible light. The prepared catalysts showed a satisfactory amount of degradation of these dyes. Benavente et al. [99] synthesized a catalyst ZnO/MoS2 by a sol-gel method and investigated the degradation of 10 ppm of Methylene blue (MB) dye using visible light for 300 min, resulting in a 75% degradation.
Wang and their colleagues have prepared a binary system of MoS2 with TiO2. The MoS2 is synthesized using a simple hydrothermal method, and their ability to degrade RhB, MB, and MO dyes for 60 min showed activities of 99.6, 96.4, and 87.4%, respectively [100]. Selvaraj et al. [101] have prepared a type-II MoS2/ZnO composite via the hydrothermal method. It degraded the MB dye using UV light and demonstrated 99% degradation activity. In a recent study by Bargozideh et al. [102], MoS2 nanoflowers were made using a hydrothermal process and fabricated with BiFeO3, which showed degradation of 89% on RhB dye.
Ding and co-workers have modified MoS2 with graphene oxide (GO), and studied for the degradation of MB dye [103]. A total of 10 mg of the nanocomposite was used to study the degradation of MB dye in the presence of visible light. The composite showed a degradation of 99%, which is an enhancement of 20% activity over the pristine MoS2. In this study, they have synthesized MoS2 using solvothermal method using n-butyl lithium as solvent. The preparation of GO was carried out by the sol-gel method. Tang et al. [104] in a recent study, prepared modified MoS2 using SrZrO3 via the hydrothermal method and studied the photodegradation of the MB dye using visible light irradiation. The degradation efficacy was found to be 99.7% for 15 mg nanocomposite in 25 ppm of MB dye. MoS2/ZnS synthesized via hydrothermal method have been used for the degradation of MB dye. In total, 10 ppm dye solution along with 50 mg of nanocomposite gave 99.89% degradation activity [105]. In a recent study, Jian et al. constructed a BiOBr loaded MoS2 and studied the photocatalytic activity on the RhB dye [70]. Using 20 mg of nanocomposite in 10 ppm solution of RhB dye in Xenon arc light achieved 96% efficiency. Liu et al., synthesized a flower-like CeO2/MoS2 composite and studied its catalytic activity under visible light for the degradation of MO dye. The nanocomposite showed a 96.1% efficiency in 90 min [106]. Sharma et al., used a biosynthetic approach to construct a 3D/2D CeO2/MoS2 nanocomposite [107]. Here, 20 mg of the composite decomposed 20ppm of Methyl violet (MV) dye solution under a visible light source (300 W). The efficiency was found to be 96.25% in 90 min. Wang et al. [108], synthesied a CeO2/MoS2 2D nanostructure via hydrothermal method for the reduction study of aqueous Cr(VI). Then, 6 mg of catalyst and 5ppm of Cr(VI) resulted in a 99% removal of Cr(VI) in 120 min.
From all these studies, it can be seen that the fabrication of effective photocatalysts on MoS2 reduces the band gap of the final composite prepared. All of the photocatalysts discussed below are semiconductor photocatalysts, where the narrowing of the band gap considerably boosts catalytic activity.

5.2. Ternary Systems of MoS2

Different oxides have been doped with MoS2 to increase their degradation efficacy in photocatalysis. g-C3N4 is a metal-free semiconductor having a layered structure with great electron-proton transferability and has found applications as a photocatalyst [109]. Incorporating TiO2 into the binary system above to obtain MoS2/g-C3N4/TiO2 ternary system has led to a significant increase in the action of MoS2 and g-C3N4 photocatalysts.
Jo et al. synthesized a ternary MoS2/g-C3N4/TiO2 system in 2015 [110]. The composite was prepared over multiple steps. Initially, MoS2 was prepared using hydrothermal method and then exfoliated with g-C3N4. TiO2 was added by the impregnation method to obtain MoS2/g-C3N4/TiO2. In the photocatalytic experiment, to 10 ppm MB, 30 mg of prepared catalyst was added, and the reaction observed for 60 min under visible light irradiation. The system g-C3N4 (10%)/TiO2/MoS2 (0.2%) showed an activity of 99.5%. Atrazine herbicide degradation has also been studied under the same conditions for 300 min to obtain 86% degradation [110]. In 2016, Zhang et al. [111] fabricated TiO2/g-C3N4/MoS2 composite from g-C3N4 and exfoliated it with hydrothermally synthesized MoS2 to obtain TiO2/g-C3N4/MoS2 using solvothermal method and used for MO degradation (20 ppm of MO solution). The degradation of pure TiO2, pure g-C3N4, g-C3N4/MoS2, and TiO2/g-C3N4 was 28.9, 22.4, 26.7, and 90.1%, respectively (100 mg catalyst). In comparison, all TiO2/g-C3N4/MoS2 composites had better photocatalytic activity than pure TiO2, g-C3N4, and TiO2/g-C3N4 composites, demonstrating that TiO2 and g-C3N4/MoS2 hybrid had a synergistic impact. The Gaussian09 programme was used to perform DFT investigations on the as-prepared composite photocatalyst for both geometry optimization and frequency estimation [111]. Figure 13 shows the optimized TiO2/g-C3N4/MoS2 composite structure utilizing the LANL2DZ basis package at the density functional B3LYP level. Following the calculations of the TiO2, MoS2, g-C3N4, and TiO2/MoS2 moieties, the structures of potential TiO2/g-C3N4/MoS2 combinations were computed using the B3LYP density functional theory with the LANL2D basis set. The B3LYP-D3 method was used to optimize the adsorption energy of a TiO2/g-C3N4/MoS2 composite.
In 2020, Mahalakshmi et al. reported the g-C3N4/MoS2/TiO2 system for photodegradation [112]. The g-C3N4, synthesized from melamine was added to titanium isopropoxide (TTIP) with rapid stirring. Then the hydrothermally prepared MoS2 and the above-prepared solutions were added, and hydrothermal synthesis was carried out to obtain a C3N4/MoS2/TiO2 system. They investigated the catalyst’s activity at various weight percentages as well as the degradation of the MO dye. In 0.355 g of MO dye, 50 mg of catalyst was added. Under visible light irradiation, the reaction was carried out for 60 min. The catalyst showed an 88% degradation. The same hybrid catalyst has been used for the degradation of 4-nitrophenol (4-NP) to obtain a degradation efficiency of 74%. In a recent study, Jaleel et al. prepared the nanocomposite MoS2/g-C3N4/TiO2 using hydrothermal and exfoliation methods [59]. Here, 50 mg of prepared nanocomposite was added to 50 ppm of malachite green (MG) dye to obtain 97% efficiency within 60 min. The Box–Behnken design of the response surface approach is also used to examine the optimal experimental conditions for dye degradation [59].
Modification of MoS2 nanoparticles with the g-C3N4 and ZnO is also reported. Since the two photocatalysts are well-matched with overlapping band alignments, coupling ZnO with g-C3N4 could yield an excellent heterostructure where increased charge separation takes place [113]. Theoretically, electrons excited from the valence band (VB) to the conduction band (CB) of the g-C3N4 will then migrate to the CB of ZnO, such heterostructures have been shown to have enhanced photocatalytic activity. The MoS2/g-C3N4/ZnO nanocomposite is used to degrade the MB dye for 60 min, and the MB dye shows more than 90% degradation. Lu et al. [114] synthesized g-C3N4 and Ag on MoS2 and investigated the degradation activity on the RhB dye for 90 min using a Xenon arc lamp as a light source. The amount of MoS2 used in this analysis was 100 g, and the dye concentration was 20 ppm to yield a degradation of 95.8%. The MoS2 has a flower-like morphology and was synthesized using a hydrothermal process.
Beyhaqi and colleagues prepared a nanocomposite by modifying the MoS2 using g-C3N4 and WO3. The photodegradation study was performed on three dyes RhB, MO, and MB using Xenon lamps as the light source. The degradation efficiency obtained was 99.9, 83.4, and 91.8%, respectively, for RhB, MO, and MB. The same nanocomposite could degrade ciprofloxacin drug completely within 2 h of irradiation of Xenon arc as light source [115]. Pant et al. [116] modified MoS2 using CdS and TiO2, and it was used to examine the breakdown of MB dye on carbon nanotubes. The dye degraded completely within 15 min of irradiation with visible light. Vignesh et al. modified MoS2 using g-C3N4 and Bi2O3 and studied the reaction of MB degradation. When the visible light was irradiated for 90 min on the dye (20 ppm) using 50 mg catalyst, it gave a good degradation activity of 98.5% [117]. This nanocomposite was also used for the destruction of E. coli and S. aureus bacteria. Reusability plays an important role in catalysis. Most of the reports related to MoS2-based systems have shown good reusability up to five cycles [59,116,118,119,120]. Talukar et al., synthesized a ternary hybrid nanoflower CeO2-ZrO2@MoS2 to study the sonophotocatalytic degradation of naproxen (NPX) using a visible light source (250 W) LED lamp. In total, 50 ppm of NPX solution yielded an enhanced degradation of 96% in 40 min.
A summary of the MoS2 nanocomposites and the degradation of various dyes has been given in Table 1 [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141]. In all these studies, we see that the fabricating/doping agent changes the degradational activity of the nanocomposite. The band gap shows a red shift in the modified MoS2, which helps in the better utilization of visible light. The morphology and degradation activities have a direct relationship which in turn depends on the preparation technique. Many authors have done computational and statistical studies to understand the highest degradation efficacy and optimal conditions for the degradation of organic dyes.

6. Mechanism of Dye Degradation Using MoS2-Based Systems

MoS2, a semiconductor photocatalyst undergoes electron-hole pair recombination during the degradation reaction. The mechanism of the MoS2-based photocatalyst is described using electron pairs and holes [143]. Based on the different types of electron transfer, the mechanism is classified as traditional and Z-scheme mechanisms. The traditional mechanism of the binary system MoS2/Cu2O is given in Figure 14a [144], and the ternary system MoS2/CdS/TiO2 is given in Figure 14c [116]. The Z-scheme mechanism of the binary system MoS2/g-C3N4 [145], and ternary system MoS2/g-C3N4/TiO2 [59] are shown in Figure 14b,d, respectively. In the traditional mechanism, valence band electrons (e) absorb enough energy to migrate to the conduction band and form holes (h+). Because the conduction band of MoS2 is close to that of TiO2, photoelectrons released in TiO2 are transported to MoS2 more quickly, enabling carrier separation and minimizing the photogenerated electron and hole recombination [126]. In the Z-scheme mechanism [146], the electrons move to CB from VB, and then they will transfer to the VB of the nearest atom. Thus, a zig zag electron transfer is exhibited by these nanocomposites. Mahalakshmi et al. [112] presented an S-scheme mechanism, using MoS2/g-C3N4/TiO2 catalyst (Figure 14e). During the irradiation of light, electrons from VB of g-C3N4, MoS2, and TiO2 are stimulated singly to the CB in this process. In the hole of TiO2’s VB, the excited electrons recombine. The approach helped in the spatial isolation and extraction of photoexcited carriers with larger redox capabilities, and stored electrons of the MoS2/TiO2/g-C3N4 surface are helpful in the decomposition of organic dyes.
As mentioned earlier, the PL spectra have an important role in identifying the particular mechanism of dye degradation. Figure 15a represents the PL spectrum and its traditional mechanism using MoS2@TiO2 [86]. Jia et al. have proposed a Z-scheme mechanism based on the PL spectra for the Au-CoFe2O4/MoS2 nanocomposite [134]. When the spectral intensity is increased after the fabrication, the nanocomposite chooses a Z-scheme mechanism. At low intensities, the composite chooses a traditional mechanism. The PL spectra and the mechanism are given in Figure 15b. A reduction in PL strength indicates that the electron-hole pair recombination has been suppressed.

7. Conclusions and Future Perspective

This review has discussed the geometry, polymorphism, different methods of preparation, and characterization, of MoS2-based composites for the photocatalytic degradation of organic dyes. MoS2-based binary and ternary systems have been considered. The degradation ability depends on factors such as morphology and the preparation techniques employed. The chemical properties of the added components help in the formation of a heterojunction with MoS2, leading to an increase in the efficacy of the final catalyst towards the degradation of dyes. Based on the various studies conducted, the photocatalytic activity of MoS2-based nanomaterials has the following characteristics. The force of attraction between the various materials used to fabricate the hybrid catalyst, and the crystallographic plane where the fabrication is taking place on the nanomaterial are significant factors affecting their efficiency. The MoS2-based nanomaterials show great degradation activity towards organic dyes and toxic pollutants. This area of study offers significant opportunities for fabricating excellent photocatalysts to affect the degradation of organic pollutant dyes from wastewater.

Author Contributions

J.R.J.U.—Methodology, software, investigation, data curation, writing original draft. M.R.—Methodology, investigation, writing original draft. S.D.K.R.—Investigation, writing, review and editing, supervision, project administration. D.P.—Visualization, validation, writing review and editing. M.K.M.—Resource, investigation, data curation, writing, review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This study did not report any data.

Acknowledgments

The authors are grateful to CHRIST (Deemed to be University) for the esteemed support and encouragement.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kroto, H.W.; Heath, J.R.; O’Brien, S.C.; Curl, R.F.; Smalley, R.E. C60: Buckminsterfullerene. Nature 1985, 318, 162–163. [Google Scholar] [CrossRef]
  2. Bethune, D.S.; Kiang, C.H.; de Vries, M.S.; Gorman, G.; Savoy, R.; Vazquez, J.; Beyers, R. Cobalt-Catalysed Growth of Carbon Nanotubes with Single-Atomic-Layer Walls. Nature 1993, 363, 605–607. [Google Scholar] [CrossRef]
  3. Xie, X.; Kretschmer, K.; Wang, G. Advances in Graphene-Based Semiconductor Photocatalysts for Solar Energy Conversion: Fundamentals and Materials Engineering. Nanoscale 2015, 7, 13278–13292. [Google Scholar] [CrossRef] [Green Version]
  4. Wilson, J.A.; Yoffe, A.D. The Transition Metal Dichalcogenides Discussion and Interpretation of the Observed Optical, Electrical and Structural Properties. Adv. Phys. 1969, 18, 193–335. [Google Scholar] [CrossRef]
  5. Singh, A.K.; Kumar, P.; Late, D.J.; Kumar, A.; Patel, S.; Singh, J. 2D Layered Transition Metal Dichalcogenides (MoS2): Synthesis, Applications and Theoretical Aspects. Appl. Mater. Today 2018, 13, 242–270. [Google Scholar] [CrossRef]
  6. Zhang, X.; Teng, S.Y.; Loy, A.C.M.; How, B.S.; Leong, W.D.; Tao, X. Transition Metal Dichalcogenides for the Application of Pollution Reduction: A Review. Nanomaterials 2020, 10, 1012. [Google Scholar] [CrossRef]
  7. Monga, D.; Sharma, S.; Shetti, N.P.; Basu, S.; Reddy, K.R.; Aminabhavi, T.M. Advances in Transition Metal Dichalcogenide-Based Two-Dimensional Nanomaterials. Mater. Today Chem. 2021, 19, 100399. [Google Scholar] [CrossRef]
  8. Agarwal, V.; Chatterjee, K. Recent Advances in the Field of Transition Metal Dichalcogenides for Biomedical Applications. Nanoscale 2018, 10, 16365–16397. [Google Scholar] [CrossRef]
  9. Zhang, G.; Liu, H.; Qu, J.; Li, J. Two-Dimensional Layered MoS2: Rational Design, Properties and Electrochemical Applications. Energy Environ. Sci. 2016, 9, 1190–1209. [Google Scholar] [CrossRef]
  10. Gupta, U.; Rao, C.N.R. Hydrogen Generation by Water Splitting Using MoS2 and Other Transition Metal Dichalcogenides. Nano Energy 2017, 41, 49–65. [Google Scholar] [CrossRef]
  11. Yuan, Y.J.; Lu, H.W.; Yu, Z.T.; Zou, Z.G. Noble-Metal-Free Molybdenum Disulfide Cocatalyst for Photocatalytic Hydrogen Production. ChemSusChem 2015, 8, 4113–4127. [Google Scholar] [CrossRef]
  12. Jin, X.; Fan, X.; Tian, J.; Cheng, R.; Li, M.; Zhang, L. MoS2 Quantum Dot Decorated g-C3N4 Composite Photocatalyst with Enhanced Hydrogen Evolution Performance. RSC Adv. 2016, 6, 52611–52619. [Google Scholar] [CrossRef]
  13. Dhiman, V.; Kondal, N. MoS2–ZnO Nanocomposites for Photocatalytic Energy Conversion and Solar Applications. Phys. B 2022, 628, 413569. [Google Scholar] [CrossRef]
  14. Jiang, Y.; Wu, R.; Zhou, L.; Li, Z.; Tian, L.; Cao, R. Phase Engineering-Defective 1T @ 2H MoS2 Nanoflowers as Excellent Full Spectrum Photocatalyst. J. Alloys Compd. 2022, 910, 164898. [Google Scholar] [CrossRef]
  15. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent Development of Two-Dimensional Transition Metal Dichalcogenides and Their Applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  16. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  17. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef]
  18. Tongay, S.; Zhou, J.; Ataca, C.; Liu, J.; Kang, J.S.; Matthews, T.S.; You, L.; Li, J.; Grossman, J.C.; Wu, J. Broad-Range Modulation of Light Emission in Two-Dimensional Semiconductors by Molecular Physisorption Gating. Nano Lett. 2013, 13, 2831–2836. [Google Scholar] [CrossRef]
  19. Liu, T.; Wang, C.; Gu, X.; Gong, H.; Cheng, L.; Shi, X.; Feng, L.; Sun, B.; Liu, Z. Drug Delivery with PEGylated MoS2 Nano-Sheets for Combined Photothermal and Chemotherapy of Cancer. Adv. Mater. 2014, 26, 3433–3440. [Google Scholar] [CrossRef]
  20. Tian, Y.; Ge, L.; Wang, K.; Chai, Y. Synthesis of Novel MoS2/g-C3N4 Heterojunction Photocatalysts with Enhanced Hydrogen Evolution Activity. Mater. Charact. 2014, 87, 70–73. [Google Scholar] [CrossRef]
  21. Lu, Q.; Yu, Y.; Ma, Q.; Chen, B.; Zhang, H. 2D Transition-Metal-Dichalcogenide-Nanosheet-Based Composites for Photocatalytic and Electrocatalytic Hydrogen Evolution Reactions. Adv. Mater. 2016, 28, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  22. Guardia, L.; Paredes, J.I.; Munuera, J.M.; Villar-Rodil, S.; Ayán-Varela, M.; Martínez-Alonso, A.; Tascón, J.M.D. Chemically Exfoliated MoS2 Nanosheets as an Efficient Catalyst for Reduction Reactions in the Aqueous Phase. ACS Appl. Mater. Interfaces 2014, 6, 21702–21710. [Google Scholar] [CrossRef] [PubMed]
  23. Hill, E.H.; Zhang, Y.; Whitten, D.G. Aggregation of Cationic P-Phenylene Ethynylenes on Laponite Clay in Aqueous Dispersions and Solid Films. J. Colloid Interface Sci. 2015, 449, 347–356. [Google Scholar] [CrossRef]
  24. Das, P.; Malho, J.M.; Rahimi, K.; Schacher, F.H.; Wang, B.; Demco, D.E.; Walther, A. Nacre-Mimetics with Synthetic Nanoclays up to Ultrahigh Aspect Ratios. Nat. Commun. 2015, 6, 5967. [Google Scholar] [CrossRef] [Green Version]
  25. Dong, J.; Fang, W.; Yuan, H.; Xia, W.; Zeng, X.; Shangguan, W. Few-Layered MoS2/ZnCdS/ZnS Heterostructures with an Enhanced Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2022, 5, 4893–4902. [Google Scholar] [CrossRef]
  26. Salarizadeh, P.; Askari, M.B.; Di Bartolomeo, A. MoS2/Ni3S2/Reduced Graphene Oxide Nanostructure as an Electrocatalyst for Alcohol Fuel Cells. ACS Appl. Nano Mater. 2022, 5, 3361–3373. [Google Scholar] [CrossRef]
  27. Li, Y.; Fu, R.; Duan, Z.; Zhu, C.; Fan, D. Construction of Multifunctional Hydrogel Based on the Tannic Acid-Metal Coating Decorated MoS2 Dual Nanozyme for Bacteria-Infected Wound Healing. Bioact. Mater. 2022, 9, 461–474. [Google Scholar] [CrossRef]
  28. Shen, H.; Liao, S.; Jiang, C.; Zhang, J.; Wei, Q.; Ghiladi, R.A.; Wang, Q. In Situ Grown Bacterial Cellulose/ MoS2 Composites for Multi-Contaminant Wastewater Treatment and Bacteria Inactivation. Carbohydr. Polym. 2022, 277, 118853. [Google Scholar] [CrossRef]
  29. Sivaranjani, P.R.; Janani, B.; Thomas, A.; Raju, L.L.; Khan, S.S. Recent Development in MoS2-Based Nano-Photocatalyst for the Degradation of Pharmaceutical Active Compounds. J. Clean. Prod. 2022, 352, 131506. [Google Scholar] [CrossRef]
  30. Szkoda, M.; Zarach, Z.; Nadolska, M.; Trykowski, G.; Trzciński, K. SnO2 Nanoparticles Embedded onto MoS2 Nanoflakes-an Efficient Catalyst for Photodegradation of Methylene Blue and Photoreduction of Hexavalent Chromium. Electrochim. Acta 2022, 414, 140173. [Google Scholar] [CrossRef]
  31. Schwarzenbach, R.P.; Egli, T.; Hofstetter, T.B.; Von Gunten, U.; Wehrli, B. Global Water Pollution and Human Health. Annu. Rev. Environ. Resour. 2010, 35, 109–136. [Google Scholar] [CrossRef]
  32. Dwivedi, A.K. Researches in Water Pollution: A Review. Int. Res. J. Nat. Appl. Sci. 2017, 4, 118–142. [Google Scholar]
  33. Pirilä, M.; Saouabe, M.; Ojala, S.; Rathnayake, B.; Drault, F.; Valtanen, A.; Huuhtanen, M.; Brahmi, R.; Keiski, R.L. Photocatalytic Degradation of Organic Pollutants in Wastewater. Top. Catal. 2015, 58, 1085–1099. [Google Scholar] [CrossRef]
  34. Wu, F.; Guan, P.; Xu, X.; Kong, Y.; Cao, D.; Yang, J. High Aspect-Ratio of MoS2 Nanoribbons via a Single-Source Precursor Route for Photocatalytic Degradation. Mater. Lett. 2022, 315, 131987. [Google Scholar] [CrossRef]
  35. Liang, Y.; Yu, L. A New Class of Semiconducting Polymers for Bulk Heterojunction Solar Cells with Exceptionally High Performance. Acc. Chem. Res. 2010, 43, 1227–1236. [Google Scholar] [CrossRef] [PubMed]
  36. Kong, D.; Wang, H.; Cha, J.J.; Pasta, M.; Koski, K.J.; Yao, J.; Cui, Y. Synthesis of MoS2 and MoSe2 Films with Vertically Aligned Layers. Nano Lett. 2013, 13, 1341–1347. [Google Scholar] [CrossRef]
  37. Kumar, D.P.; Kim, E.H.; Park, H.; Chun, S.Y.; Gopannagari, M.; Bhavani, P.; Reddy, D.A.; Song, J.K.; Kim, T.K. Tuning Band Alignments and Charge-Transport Properties through MoSe2 Bridging between MoS2 and Cadmium Sulfide for Enhanced Hydrogen Production. ACS Appl. Mater. Interfaces 2018, 10, 26153–26161. [Google Scholar] [CrossRef]
  38. Yuan, Y.; Guo, R.; Hong, L.; Ji, X.; Li, Z.; Lin, Z.; Pan, W. Recent Advances and Perspectives of MoS2-Based Materials for Photocatalytic Dyes Degradation: A Review. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125836. [Google Scholar] [CrossRef]
  39. 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]
  40. Wypych, F.; Weber, T.; Prins, R. Scanning Tunneling Microscopic Investigation of Kx(H2O)Y MoS2. Surf. Sci. 1997, 380, L474–L478. [Google Scholar] [CrossRef]
  41. El-Mahalawy, S.H.; Evans, B.L. The Thermal Expansion of 2H-MoS2, 2H-MoSe2 and 2H-WSe2 between 20 and 800 °C. J. Appl. Crystallogr. 1976, 9, 403–406. [Google Scholar] [CrossRef]
  42. Arif Khalil, R.M.; Hussain, F.; Manzoor Rana, A.; Imran, M.; Murtaza, G. Comparative Study of Polytype 2H-MoS2 and 3R-MoS2 Systems by Employing DFT. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 106, 338–345. [Google Scholar] [CrossRef]
  43. Lin, L.; Miao, N.; Huang, J.; Zhang, S.; Zhu, Y.; Horsell, D.D.; Ghosez, P.; Sun, Z.; Allwood, D.A. A Photocatalyst of Sulphur Depleted Monolayered Molybdenum Sulfide Nanocrystals for Dye Degradation and Hydrogen Evolution Reaction. Nano Energy 2017, 38, 544–552. [Google Scholar] [CrossRef] [Green Version]
  44. Nguyen, T.N.; Kampouri, S.; Valizadeh, B.; Luo, W.; Ongari, D.; Planes, O.M.; Züttel, A.; Smit, B.; Stylianou, K.C. Photocatalytic Hydrogen Generation from a Visible-Light-Responsive Metal-Organic Framework System: Stability versus Activity of Molybdenum Sulfide Cocatalysts. ACS Appl. Mater. Interfaces 2018, 10, 30035–30039. [Google Scholar] [CrossRef] [Green Version]
  45. Rao, C.N.R.; Maitra, U.; Waghmare, U.V. Extraordinary Attributes of 2-Dimensional MoS2 Nanosheets. Chem. Phys. Lett. 2014, 609, 172–183. [Google Scholar] [CrossRef]
  46. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A Materials Genome Approach to Accelerating Materials Innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef] [Green Version]
  47. Yu, H.; Zhu, H.; Dargusch, M.; Huang, Y. A Reliable and Highly Efficient Exfoliation Method for Water-Dispersible MoS2 Nanosheet. J. Colloid Interface Sci. 2018, 514, 642–647. [Google Scholar] [CrossRef] [Green Version]
  48. Funke, S.; Miller, B.; Parzinger, E.; Thiesen, P.; Holleitner, A.W.; Wurstbauer, U. Imaging Spectroscopic Ellipsometry of MoS2. J. Phys. Condens. Matter 2016, 28, 385301. [Google Scholar] [CrossRef] [Green Version]
  49. Ghasemi, F.; Abdollahi, A.; Mohajerzadeh, S. Controlled Plasma Thinning of Bulk MoS2 Flakes for Photodetector Fabrication. ACS Omega 2019, 4, 19693–19704. [Google Scholar] [CrossRef] [Green Version]
  50. Zheng, J.; Yan, X.; Lu, Z.; Qiu, H.; Xu, G.; Zhou, X.; Wang, P.; Pan, X.; Liu, K.; Jiao, L. High-Mobility Multilayered MoS2 Flakes with Low Contact Resistance Grown by Chemical Vapor Deposition. Adv. Mater. 2017, 29, 1604540. [Google Scholar] [CrossRef]
  51. Zhou, Q.; Su, S.; Hu, D.; Lin, L.; Yan, Z.; Gao, X.; Zhang, Z.; Liu, J.-M. Ultrathin MoS2-Coated Ag@Si Nanosphere Arrays as an Efficient and Stable Photocathode for Solar-Driven Hydrogen Production. Nanotechnology 2018, 29, 105402. [Google Scholar] [CrossRef] [PubMed]
  52. Agafonov, V.; Nargelienė, V.; Balakauskas, S.; Bukauskas, V.; Kamarauskas, M.; Lukša, A.; Mironas, A.; Rėza, A.; Šetkus, A. Single Variable Defined Technology Control of the Optical Properties in MoS2 Films with Controlled Number of 2D-Layers. Nanotechnology 2020, 31, 025602. [Google Scholar] [CrossRef] [PubMed]
  53. Amani, M.; Burke, R.A.; Ji, X.; Zhao, P.; Lien, D.H.; Taheri, P.; Ahn, G.H.; Kirya, D.; Ager, J.W.; Yablonovitch, E.; et al. High Luminescence Efficiency in MoS2 Grown by Chemical Vapor Deposition. ACS Nano 2016, 10, 6535–6541. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, X.; Nan, H.; Xiao, S.; Wan, X.; Ni, Z.; Gu, X.; Ostrikov, K. Shape-Uniform, High-Quality Monolayered MoS2 Crystals for Gate-Tunable Photoluminescence. ACS Appl. Mater. Interfaces 2017, 9, 42121–42130. [Google Scholar] [CrossRef] [PubMed]
  55. Alharbi, A.; Armstrong, D.; Alharbi, S.; Shahrjerdi, D. Physically Unclonable Cryptographic Primitives by Chemical Vapor Deposition of Layered MoS2. ACS Nano 2017, 11, 12772–12779. [Google Scholar] [CrossRef] [PubMed]
  56. Balendhran, S.; Ou, J.Z.; Bhaskaran, M.; Sriram, S.; Ippolito, S.; Vasic, Z.; Kats, E.; Bhargava, S.; Zhuiykov, S.; Kalantar-Zadeh, K. Atomically Thin Layers of MoS2 via a Two Step Thermal Evaporation-Exfoliation Method. Nanoscale 2012, 4, 461–466. [Google Scholar] [CrossRef] [Green Version]
  57. Qi, H.; Wang, L.; Sun, J.; Long, Y.; Hu, P.; Liu, F.; He, X. Production Methods of Van Der Waals Heterostructures Based on Transition Metal Dichalcogenides. Crystals 2018, 8, 35. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, J.; Li, T.; Wang, Q.; Wang, W.; Shi, R.; Wang, N.; Amini, A.; Cheng, C. Controlled Growth of Atomically Thin Transition Metal Dichalcogenides via Chemical Vapor Deposition Method. Mater. Today Adv. 2020, 8, 100098. [Google Scholar] [CrossRef]
  59. Jaleel, U.C.; Devi, K.R.; Madhushree, P. Statistical and Experimental Studies of MoS2/g-C3N4/TiO2: A Ternary Z-Scheme Hybrid Composite. J. Mater. Sci. 2021, 56, 6922–6944. [Google Scholar] [CrossRef]
  60. Chaudhary, N.; Khanuja, M.; Abid; Islam, S.S. Hydrothermal Synthesis of MoS2 Nanosheets for Multiple Wavelength Optical Sensing Applications. Sens. Actuators A Phys. 2018, 277, 190–198. [Google Scholar] [CrossRef]
  61. Li, G.; Li, C.; Tang, H.; Cao, K.; Chen, J.; Wang, F.; Jin, Y. Synthesis and Characterization of Hollow MoS2 Microspheres Grown from MoO3 Precursors. J. Alloys Compd. 2010, 501, 275–281. [Google Scholar] [CrossRef]
  62. Wang, D.; Pan, Z.; Wu, Z.; Wang, Z.; Liu, Z. Hydrothermal Synthesis of MoS2 Nanoflowers as Highly Efficient Hydrogen Evolution Reaction Catalysts. J. Power Sources 2014, 264, 229–234. [Google Scholar] [CrossRef]
  63. Song, X.C.; Zheng, Y.F.; Zhao, Y.; Yin, H.Y. Hydrothermal Synthesis and Characterization of CNT@ MoS2 Nanotubes. Mater. Lett. 2006, 60, 2346–2348. [Google Scholar] [CrossRef]
  64. Liu, C.; Wang, Q.; Jia, F.; Song, S. Adsorption of Heavy Metals on Molybdenum Disulfide in Water: A Critical Review. J. Mol. Liq. 2019, 292, 111390. [Google Scholar] [CrossRef]
  65. Zou, X.; Zhang, J.; Zhao, X.; Zhang, Z. MoS2/RGO Composites for Photocatalytic Degradation of Ranitidine and Elimination of NDMA Formation Potential under Visible Light. Chem. Eng. J. 2020, 383, 123084. [Google Scholar] [CrossRef]
  66. Anwer, S.; Huang, Y.; Li, B.; Govindan, B.; Liao, K.; Cantwell, W.J.; Wu, F.; Chen, R.; Zheng, L. Nature-Inspired, Graphene-Wrapped 3D MoS2 Ultrathin Microflower Architecture as a High-Performance Anode Material for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 22323–22331. [Google Scholar] [CrossRef]
  67. Xie, J.; Zhang, J.; Li, S.; Grote, F.; Zhang, X.; Zhang, H.; Wang, R.; Lei, Y.; Pan, B.; Xie, Y. Controllable Disorder Engineering in Oxygen-Incorporated MoS2 Ultrathin Nanosheets for Efficient Hydrogen Evolution. J. Am. Chem. Soc. 2013, 135, 17881–17888. [Google Scholar] [CrossRef]
  68. Zhou, X.; Xu, B.; Lin, Z.; Shu, D.; Ma, L. Hydrothermal Synthesis of Flower-like MoS2 nanospheres for Electrochemical Supercapacitors. J. Nanosci. Nanotechnol. 2014, 14, 7250–7254. [Google Scholar] [CrossRef]
  69. Xia, J.; Ren, X.; Zhao, L.; Tang, M.; Tan, L.; Fu, C.; Wu, Q.; Ren, J.; Ding, M.; Meng, X. Synthesis of MoS2 Nanoflowers on CdS Nanorods with a Simple Route and Their Application in Removal of Dyes. J. Nanoparticle Res. 2022, 24, 1–9. [Google Scholar] [CrossRef]
  70. Jian, R.; Liu, J.; Wu, Z.; Zheng, H.; Zhu, K.; Sun, Z. Constructing Z-Scheme Structure by Loading BiOBr with (010) Exposure on the Surface of MoS2 and Its Enhanced Photocatalytic Property for Degrading RhB. J. Mater. Sci. Mater. Electron. 2022, 33, 6722–6733. [Google Scholar] [CrossRef]
  71. Zhao, T.; Xing, Z.; Xiu, Z.; Li, Z.; Chen, P.; Zhu, Q.; Zhou, W. Synergistic Effect of Surface Plasmon Resonance, Ti3+ and Oxygen Vacancy Defects on Ag/ MoS2/TiO2-x Ternary Heterojunctions with Enhancing Photothermal Catalysis for Low-Temperature Wastewater Degradation. J. Hazard. Mater. 2019, 364, 117–124. [Google Scholar] [CrossRef] [PubMed]
  72. Aliofkhazraei, M. Handbook of Nanoparticles; Aliofkhazraei, M., Ed.; Springer International Publishing: Cham, Switzerland, 2016; ISBN 978-3-319-15337-7. [Google Scholar]
  73. 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]
  74. Zhang, X.; Selkirk, A.; Zhang, S.; Huang, J.; Li, Y.; Xie, Y.; Dong, N.; Cui, Y.; Zhang, L.; Blau, W.J.; et al. MoS2/Carbon Nanotube Core-Shell Nanocomposites for Enhanced Nonlinear Optical Performance. Chem.-A Eur. J. 2017, 23, 3321–3327. [Google Scholar] [CrossRef] [PubMed]
  75. Gao, M.R.; Liang, J.X.; Zheng, Y.R.; Xu, Y.F.; Jiang, J.; Gao, Q.; Li, J.; Yu, S.H. An Efficient Molybdenum Disulfide/Cobalt Diselenide Hybrid Catalyst for Electrochemical Hydrogen Generation. Nat. Commun. 2015, 6, 5982. [Google Scholar] [CrossRef]
  76. Zheng, S.; Zheng, L.; Zhu, Z.; Chen, J.; Kang, J.; Huang, Z.; Yang, D. MoS2 Nanosheet Arrays Rooted on Hollow RGO Spheres as Bifunctional Hydrogen Evolution Catalyst and Supercapacitor Electrode. Nano-Micro Lett. 2018, 10, 62. [Google Scholar] [CrossRef] [Green Version]
  77. Antonelou, A.; Syrrokostas, G.; Sygellou, L.; Leftheriotis, G.; Dracopoulos, V.; Yannopoulos, S.N. Facile, Substrate-Scale Growth of Mono- and Few-Layer Homogeneous MoS2 Films on Mo Foils with Enhanced Catalytic Activity as Counter Electrodes in DSSCs. Nanotechnology 2015, 27, 45404. [Google Scholar] [CrossRef]
  78. Zheng, X.; Zhu, L.; Yan, A.; Bai, C.; Xie, Y. Ultrasound-Assisted Cracking Process to Prepare MoS2 Nanorods. Ultrason. Sonochem. 2004, 11, 83–88. [Google Scholar] [CrossRef]
  79. Zhu, J.; Wang, H.; Liu, J.; Ouyang, L.; Zhu, M. Exfoliation of MoS2 and H-BN Nanosheets by Hydrolysis of LiBH4. Nanotechnology 2017, 28, 115604. [Google Scholar] [CrossRef]
  80. Sun, Y.; Xu, J.; Zhu, Z.; Wang, Y.; Xia, H.; You, Z.; Lee, C.; Tu, C. Comparison of MoS2 Nanosheets and Hierarchical Nanospheres in the Application of Pulsed Solid-State Lasers. Opt. Mater. Express 2015, 5, 2924. [Google Scholar] [CrossRef]
  81. Visic, B.; Dominko, R.; Gunde, M.K.; Hauptman, N.; Skapin, S.D.; Remskar, M. Optical Properties of Exfoliated MoS2 Coaxial Nanotubes-Analogues of Graphene. Nanoscale Res. Lett. 2011, 6, 593. [Google Scholar] [CrossRef] [Green Version]
  82. Mahyavanshi, R.D.; Kalita, G.; Sharma, K.P.; Kondo, M.; Dewa, T.; Kawahara, T.; Tanemura, M. Synthesis of MoS2 Ribbons and Their Branched Structures by Chemical Vapor Deposition in Sulfur-Enriched Environment. Appl. Surf. Sci. 2017, 409, 396–402. [Google Scholar] [CrossRef]
  83. Wu, J.-Y.; Lin, M.-N.; Wang, L.-D.; Zhang, T. Photoluminescence of MoS2 Prepared by Effective Grinding-Assisted Sonication Exfoliation. J. Nanomater. 2014, 2014, 1–7. [Google Scholar]
  84. Shi, Y.; Li, H.; Wong, J.I.; Zhang, X.; Wang, Y.; Song, H.; Yang, H.Y. MoS2 Surface Structure Tailoring via Carbonaceous Promoter. Sci. Rep. 2015, 5, 10378. [Google Scholar] [CrossRef] [Green Version]
  85. Gao, D.; Si, M.; Li, J.; Zhang, J.; Zhang, Z.; Yang, Z.; Xue, D. Ferromagnetism in Freestanding MoS2 Nanosheets. Nanoscale Res. Lett. 2013, 8, 129. [Google Scholar] [CrossRef] [Green Version]
  86. Mahalakshmi, G.; Rajeswari, M.; Ponnarasi, P. Fabrication of Dandelion Clock-Inspired Preparation of Core-Shell TiO2 @ MoS2 Composites for Unprecedented High Visible Light-Driven Photocatalytic Performance. J. Mater. Sci. Mater. Electron. 2020, 31, 22252–22264. [Google Scholar] [CrossRef]
  87. Bahuguna, A.; Kumar, S.; Sharma, V.; Reddy, K.L.; Bhattacharyya, K.; Ravikumar, P.C.; Krishnan, V. Nanocomposite of MoS2-RGO as Facile, Heterogeneous, Recyclable, and Highly Efficient Green Catalyst for One-Pot Synthesis of Indole Alkaloids. ACS Sustain. Chem. Eng. 2017, 5, 8551–8567. [Google Scholar] [CrossRef]
  88. Lalithambika, K.C.; Shanmugapriya, K.; Sriram, S. Photocatalytic Activity of MoS2 Nanoparticles: An Experimental and DFT Analysis. Appl. Phys. A Mater. Sci. Process. 2019, 125, 817. [Google Scholar] [CrossRef]
  89. Rasamani, K.D.; Alimohammadi, F.; Sun, Y. Interlayer-Expanded MoS2. Mater. Today 2017, 20, 83–91. [Google Scholar] [CrossRef]
  90. Balakumar, S.; Rakkesh, R.A.; Prasad, A.K.; Dash, S.; Tyagi, A.K. Nanoplatelet Structures of MoO3 for H2 Gas Sensors. In Proceedings of the International Conference on Nanoscience, Engineering and Technology (ICONSET 2011), Chennai, India, 28–30 November 2011; IEEE: Chennai, India, 2011; pp. 514–517. [Google Scholar]
  91. Chakrabarty, S.; Mukherjee, A.; Basu, S. RGO- MoS2 Supported NiCo2O4 Catalyst toward Solar Water Splitting and Dye Degradation. ACS Sustain. Chem. Eng. 2018, 6, 5238–5247. [Google Scholar] [CrossRef]
  92. Nayak, S.; Swain, G.; Parida, K. Enhanced Photocatalytic Activities of RhB Degradation and H2 Evolution from in Situ Formation of the Electrostatic Heterostructure MoS2/NiFe LDH Nanocomposite through the Z-Scheme Mechanism via p–n Heterojunctions. ACS Appl. Mater. Interfaces 2019, 11, 20923–20942. [Google Scholar] [CrossRef]
  93. Rubio, E.J.; Ramana, C.V. Tungsten-Incorporation Induced Red-Shift in the Bandgap of Gallium Oxide Thin Films. Appl. Phys. Lett. 2013, 102, 191913. [Google Scholar] [CrossRef]
  94. Zhang, W.; Chuu, C.P.; Huang, J.K.; Chen, C.H.; Tsai, M.L.; Chang, Y.H.; Liang, C.T.; Chen, Y.Z.; Chueh, Y.L.; He, J.H.; et al. Ultrahigh-Gain Photodetectors Based on Atomically Thin Graphene- MoS2 Heterostructures. Sci. Rep. 2015, 4, 3826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Das, S.; Chen, H.Y.; Penumatcha, A.V.; Appenzeller, J. High Performance Multilayer MoS2 Transistors with Scandium Contacts. Nano Lett. 2013, 13, 100–105. [Google Scholar] [CrossRef] [PubMed]
  96. Tsai, M.L.; Su, S.H.; Chang, J.K.; Tsai, D.S.; Chen, C.H.; Wu, C.I.; Li, L.J.; Chen, L.J.; He, J.H. Monolayer MoS2 Heterojunction Solar Cells. ACS Nano 2014, 8, 8317–8322. [Google Scholar] [CrossRef]
  97. Late, D.J.; Huang, Y.K.; Liu, B.; Acharya, J.; Shirodkar, S.N.; Luo, J.; Yan, A.; Charles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing Behavior of Atomically Thin-Layered MoS2 Transistors. ACS Nano 2013, 7, 4879–4891. [Google Scholar] [CrossRef]
  98. 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]
  99. Benavente, E.; Durán, F.; Sotomayor-Torres, C.; González, G. Heterostructured Layered Hybrid ZnO/MoS2 Nanosheets with Enhanced Visible Light Photocatalytic Activity. J. Phys. Chem. Solids 2018, 113, 119–124. [Google Scholar] [CrossRef]
  100. Wang, C.; Zhan, Y.; Wang, Z. TiO2, MoS2, and TiO2/MoS2 Heterostructures for Use in Organic Dyes Degradation. ChemistrySelect 2018, 3, 1713–1718. [Google Scholar] [CrossRef]
  101. Selvaraj, R.; Kalimuthu, K.R.; Kalimuthu, V. A Type-II MoS2 /ZnO Heterostructure with Enhanced Photocatalytic Activity. Mater. Lett. 2019, 243, 183–186. [Google Scholar] [CrossRef]
  102. Bargozideh, S.; Tasviri, M.; Kianifar, M. Construction of Novel Magnetic BiFeO3/MoS2 Composite for Enhanced Visible-Light Photocatalytic Performance towards Purification of Dye Pollutants. Int. J. Environ. Anal. Chem. 2020, 2020, 1–15. [Google Scholar] [CrossRef]
  103. Ding, Y.; Zhou, Y.; Nie, W.; Chen, P. MoS2–GO Nanocomposites Synthesized via a Hydrothermal Hydrogel Method for Solar Light Photocatalytic Degradation of Methylene Blue. Appl. Surf. Sci. 2015, 357, 1606–1612. [Google Scholar] [CrossRef]
  104. Tang, J.; Shi, Y.; Cai, W.; Liu, F. Construction of Embedded Heterostructured SrZrO3 /Flower-like MoS2 with Enhanced Dye Photodegradation under Solar-Simulated Light Illumination. ACS Omega 2020, 5, 9576–9584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Harish, S.; Archana, J.; Navaneethan, M.; Shimomura, M.; Ikeda, H. Applied Surface Science Synergistic Interaction of 2D Layered MoS2/ZnS Nanocomposite for Highly Efficient Photocatalytic Activity under Visible Light Irradiation. Appl. Surf. Sci. 2019, 488, 36–45. [Google Scholar] [CrossRef]
  106. Liu, X.; Meng, F.; Yu, B.; Wu, H. Self-Assembly Synthesis of Flower-like CeO2/MoS2 Heterojunction with Enhancement of Visible Light Photocatalytic Activity for Methyl Orange. J. Mater. Sci. Mater. Electron. 2020, 31, 6690–6697. [Google Scholar] [CrossRef]
  107. Sharma, Y. Biosynthesis of 3D/2D CeO2/MoS2 Nanocomposites with Enhanced Photocatalytic Activity to Degrade Organic Dye in Wastewater and Statistical Optimization of Reaction Parameters. Res. Sq. 2021, 1, 1–21. [Google Scholar]
  108. Wang, H.; Wen, F.; Li, X.; Gan, X.; Yang, Y.; Chen, P.; Zhang, Y. Cerium-Doped MoS2 Nanostructures: Efficient Visible Photocatalysis for Cr(VI) Removal. Sep. Purif. Technol. 2016, 170, 190–198. [Google Scholar] [CrossRef]
  109. Inagaki, M.; Tsumura, T.; Kinumoto, T.; Toyoda, M. Graphitic Carbon Nitrides (g-C3N4) with Comparative Discussion to Carbon Materials. Carbon N. Y. 2019, 141, 580–607. [Google Scholar] [CrossRef]
  110. Jo, W.K.; Adinaveen, T.; Vijaya, J.J.; Sagaya Selvam, N.C. Synthesis of MoS2 Nanosheet Supported Z-Scheme TiO2/g-C3N4 Photocatalysts for the Enhanced Photocatalytic Degradation of Organic Water Pollutants. RSC Adv. 2016, 6, 10487–10497. [Google Scholar] [CrossRef]
  111. Zhang, W.; Xiao, X.; Li, Y.; Zeng, X.; Zheng, L.; Wan, C. Liquid-Exfoliation of Layered MoS2 for Enhancing Photocatalytic Activity of TiO2/g-C3N4 Photocatalyst and DFT Study. Appl. Surf. Sci. 2016, 389, 496–506. [Google Scholar] [CrossRef]
  112. Mahalakshmi, G.; Rajeswari, M.; Ponnarasi, P. Synthesis of Few-Layer g-C3N4 Nanosheets-Coated MoS2/TiO2 Heterojunction Photocatalysts for Photo-Degradation of Methyl Orange (MO) and 4-Nitrophenol (4-NP) Pollutants. Inorg. Chem. Commun. 2020, 120, 108146. [Google Scholar] [CrossRef]
  113. Zhang, J.; Zhang, M.; Sun, R.-Q.; Wang, X. A Facile Band Alignment of Polymeric Carbon Nitride Semiconductors to Construct Isotype Heterojunctions. Angew. Chemie Int. Ed. 2012, 51, 10145–10149. [Google Scholar] [CrossRef] [PubMed]
  114. Lu, D.; Wang, H.; Zhao, X.; Kondamareddy, K.K.; Ding, J.; Li, C.; Fang, P. Highly Efficient Visible-Light-Induced Photoactivity of Z-Scheme g-C3N4/Ag/ MoS2 Ternary Photocatalysts for Organic Pollutant Degradation and Production of Hydrogen. ACS Sustain. Chem. Eng. 2017, 5, 1436–1445. [Google Scholar] [CrossRef]
  115. Beyhaqi, A.; Zeng, Q.; Chang, S.; Wang, M.; Taghi Azimi, S.M.; Hu, C. Construction of g-C3N4/WO3/MoS2 Ternary Nanocomposite with Enhanced Charge Separation and Collection for Efficient Wastewater Treatment under Visible Light. Chemosphere 2020, 247, 125784. [Google Scholar] [CrossRef] [PubMed]
  116. Pant, B.; Park, M.; Park, S.J. MoS2/CdS/TiO2 Ternary Composite Incorporated into Carbon Nanofibers for the Removal of Organic Pollutants from Water. Inorg. Chem. Commun. 2019, 102, 113–119. [Google Scholar] [CrossRef]
  117. Shanmugam, V.; Jeyaperumal, K.S.; Mariappan, P.; Muppudathi, A.L. Fabrication of Novel g-C3N4 Based MoS2 and Bi2O3 Nanorod Embedded Ternary Nanocomposites for Superior Photocatalytic Performance and Destruction of Bacteria. New J. Chem. 2020, 44, 13182–13194. [Google Scholar] [CrossRef]
  118. Ikram, M.; Khan, M.I.; Raza, A.; Imran, M.; Ul-Hamid, A.; Ali, S. Outstanding Performance of Silver-Decorated MoS2 Nanopetals Used as Nanocatalyst for Synthetic Dye Degradation. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 124, 114246. [Google Scholar] [CrossRef]
  119. Mohammed, R.; Eid, M.; Ali, M.; Abdel-moniem, S.M.; Ibrahim, H.S. Nano-Structures & Nano-Objects Reusable and Highly Stable MoS2 Nanosheets for Photocatalytic, Sonocatalytic and Thermocatalytic Degradation of Organic Dyes: Comparative Study. Nano-Struct. Nano-Objects 2022, 31, 100900. [Google Scholar]
  120. Sindhu, A.S.; Shinde, N.B.; Harish, S.; Navaneethan, M.; Eswaran, S.K. Recoverable and Reusable Visible-Light Photocatalytic Performance of CVD Grown Atomically Thin MoS2 Films. Chemosphere 2022, 287, 132347. [Google Scholar] [CrossRef]
  121. Rani, A.; Singh, K.; Sharma, P. Investigation of Visible Light Photocatalytic Degradation of Organic Dyes by MoS2 Nanosheets Synthesized by Different Routes. Bull. Mater. Sci. 2022, 45, 63. [Google Scholar] [CrossRef]
  122. Yuan, B.; Yang, L.; Yang, H.; Bai, L.; Wang, W.; Wei, D.; Liang, Y.; Chen, H. Flexible Vacancy-Mediated MoS2-x Nanosheet Arrays for Solar-Driven Interfacial Water Evaporation, Photothermal-Enhanced Photodegradation, and Thermoelectric Generation. Energy Convers. Manag. 2022, 252, 115070. [Google Scholar] [CrossRef]
  123. Li, H.; Shen, H.; Duan, L.; Liu, R.; Li, Q.; Zhang, Q.; Zhao, X. Enhanced Photocatalytic Activity and Synthesis of ZnO Nanorods/MoS2 Composites. Superlattices Microstruct. 2018, 117, 336–341. [Google Scholar] [CrossRef]
  124. Abinaya, R.; Archana, J.; Harish, S.; Navaneethan, M.; Ponnusamy, S.; Muthamizhchelvan, C.; Shimomura, M.; Hayakawa, Y. Ultrathin Layered MoS2 Nanosheets with Rich Active Sites for Enhanced Visible Light Photocatalytic Activity. RSC Adv. 2018, 8, 26664–26675. [Google Scholar] [CrossRef] [Green Version]
  125. Sabarinathan, M.; Harish, S.; Archana, J.; Navaneethan, M.; Ikeda, H.; Hayakawa, Y. Highly Efficient Visible-Light Photocatalytic Activity of MoS2-TiO2 Mixtures Hybrid Photocatalyst and Functional Properties. RSC Adv. 2017, 7, 24754–24763. [Google Scholar] [CrossRef] [Green Version]
  126. Jiang, Q.; Wang, S.; Li, X.; Han, Z.; Zhao, C.; Di, T.; Liu, S.; Cheng, Z. Controllable Growth of MoS2 nanosheets on TiO2 burst Nanotubes and Their Photocatalytic Activity. RSC Adv. 2020, 10, 40904–40915. [Google Scholar] [CrossRef]
  127. Tama, A.M.; Das, S.; Dutta, S.; Bhuyan, M.D.I.; Islam, M.N.; Basith, M.A. MoS2 Nanosheet Incorporated α-Fe2O3/ZnO Nanocomposite with Enhanced Photocatalytic Dye Degradation and Hydrogen Production Ability. RSC Adv. 2019, 9, 40357–40367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Sharma, M.; Mohapatra, P.K.; Bahadur, D. Improved Photocatalytic Degradation of Organic Dye Using Ag3PO4/MoS2 Nanocomposite. Front. Mater. Sci. 2017, 11, 366–374. [Google Scholar] [CrossRef]
  129. Raja, V.R.; Rosaline, D.R.; Suganthi, A.; Rajarajan, M. Facile Fabrication of PbS/ MoS2 Nanocomposite Photocatalyst with Efficient Photocatalytic Activity under Visible Light. Solid State Sci. 2017, 67, 99–108. [Google Scholar] [CrossRef]
  130. Zhang, S.; Wang, L.; Liu, C.; Luo, J.; Crittenden, J.; Liu, X.; Cai, T.; Yuan, J.; Pei, Y.; Liu, Y. Photocatalytic Wastewater Purification with Simultaneous Hydrogen Production Using MoS2 QD-Decorated Hierarchical Assembly of ZnIn2S4 on Reduced Graphene Oxide Photocatalyst. Water Res. 2017, 121, 11–19. [Google Scholar] [CrossRef]
  131. Rani, A.; Singh, K.; Patel, A.S.; Chakraborti, A.; Kumar, S.; Ghosh, K.; Sharma, P. Visible Light Driven Photocatalysis of Organic Dyes Using SnO2 Decorated MoS2 Nanocomposites. Chem. Phys. Lett. 2020, 738, 136874. [Google Scholar] [CrossRef]
  132. Lu, Y.; Xu, L.; Liu, C. Magnetically Separable and Recyclable Photocatalyst MoS2-SrFe12O19 with p-n Heterojunction: Fabrication, Characterization, and Photocatalytic Mechanism. Appl. Organomet. Chem. 2020, 34, 1–15. [Google Scholar] [CrossRef]
  133. Gawari, D.; Pandit, V.; Jawale, N.; Kamble, P. Proceedings Layered MoS2 for Photocatalytic Dye Degradation. Mater. Today Proc. 2022, 53, 10–14. [Google Scholar] [CrossRef]
  134. Jia, Y.; Ma, H.; Liu, C. Au Nanoparticles Enhanced Z-Scheme Au-CoFe2O4/MoS2 Visible Light Photocatalyst with Magnetic Retrievability. Appl. Surf. Sci. 2019, 463, 854–862. [Google Scholar] [CrossRef]
  135. Gao, J.; Hu, J.; Wang, Y.; Zheng, L.; He, G.; Deng, J.; Liu, M.; Li, Y.; Liu, Y.; Zhou, H. Fabrication of Z-Scheme TiO2/SnS2/MoS2 Ternary Heterojunction Arrays for Enhanced Photocatalytic and Photoelectrochemical Performance under Visible Light. J. Solid State Chem. 2022, 307, 122737. [Google Scholar] [CrossRef]
  136. Khaing, K.K.; Yin, D.; Ouyang, Y.; Xiao, S.; Liu, B.; Deng, L.; Li, L.; Guo, X.; Wang, J.; Liu, J.; et al. Fabrication of 2D–2D Heterojunction Catalyst with Covalent Organic Framework (COF) and MoS2 for Highly Efficient Photocatalytic Degradation of Organic Pollutants. Inorg. Chem. 2020, 59, 6942–6952. [Google Scholar] [CrossRef]
  137. Lv, H.; Liu, Y.; Tang, H.; Zhang, P.; Wang, J. Synergetic Effect of MoS2 and Graphene as Cocatalysts for Enhanced Photocatalytic Activity of BiPO4 Nanoparticles. Appl. Surf. Sci. 2017, 425, 100–106. [Google Scholar] [CrossRef]
  138. Zhang, B.; Shi, H.; Hu, X.; Wang, Y.; Liu, E.; Fan, J. A Novel S-Scheme MoS2 /CdIn2S4 Flower-like Heterojunctions with Enhanced Photocatalytic Degradation and H2 Evolution Activity. J. Phys. D Appl. Phys. 2020, 53, 205101. [Google Scholar] [CrossRef]
  139. Ren, B.; Shen, W.; Li, L.; Wu, S.; Wang, W. 3D CoFe2O4 Nanorod/Flower-like MoS2 Nanosheet Heterojunctions as Recyclable Visible Light-Driven Photocatalysts for the Degradation of Organic Dyes. Appl. Surf. Sci. 2018, 447, 711–723. [Google Scholar] [CrossRef]
  140. Lu, M.; Xiao, X.; Wang, Y.; Chen, J. Construction of Novel BiOIO3/MoS2 2D/2D Heterostructures with Enhanced Photocatalytic Activity. J. Alloys Compd. 2020, 831, 154789. [Google Scholar] [CrossRef]
  141. Rathnasamy, R.; Santhanam, M.; Alagan, V. Anchoring of ZnO Nanoparticles on Exfoliated MoS2 Nanosheets for Enhanced Photocatalytic Decolorization of Methyl Red Dye. Mater. Sci. Semicond. Process. 2018, 85, 59–67. [Google Scholar] [CrossRef]
  142. Talukdar, K.; Saravanakumar, K.; Kim, Y.; Fayyaz, A.; Kim, G.; Yoon, Y.; Park, C.M. Rational Construction of CeO2–ZrO2@ MoS2 Hybrid Nanoflowers for Enhanced Sonophotocatalytic Degradation of Naproxen: Mechanisms and Degradation Pathways. Compos. Part B Eng. 2021, 215, 108780. [Google Scholar] [CrossRef]
  143. Ji, R.; Zhu, Z.; Ma, W.; Tang, X.; Liu, Y.; Huo, P. A Heterojunction Photocatalyst Constructed by the Modification of 2D-CeO2 on 2D-MoS2 Nanosheets with Enhanced Degrading Activity. Catal. Sci. Technol. 2020, 10, 788–800. [Google Scholar] [CrossRef]
  144. Zhang, X.; Xia, M.; Wang, F.; Lei, W. Cu2O/ MoS2 Composites: A Novel Photocatalyst for Photocatalytic Degradation of Organic Dyes under Visible Light. Ionics 2020, 26, 6359–6369. [Google Scholar] [CrossRef]
  145. Fu, Y.; Li, Z.; Liu, Q.; Yang, X.; Tang, H. Construction of Carbon Nitride and MoS2 Quantum Dot 2D/0D Hybrid Photocatalyst: Direct Z-Scheme Mechanism for Improved Photocatalytic Activity. Cuihua Xuebao/Chin. J. Catal. 2017, 38, 2160–2170. [Google Scholar] [CrossRef]
  146. Feng, H.; Zhou, W.; Zhang, X.; Zhang, S.; Liu, B.; Zhen, D. Synthesis of Z-Scheme Mn-CdS/ MoS2/TiO2 Ternary Photocatalysts for High-Efficiency Sunlight-Driven Photocatalysis. Adv. Compos. Lett. 2019, 28, 1–10. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Representation of the preparation, morphology, and applications of MoS2-based catalysts.
Figure 1. Representation of the preparation, morphology, and applications of MoS2-based catalysts.
Photochem 02 00042 g001
Figure 2. (a) Lateral view [16], (b) vertical view [38] of MoS2 crystal structure, and (c) Crystal structure of MoS2, 2H (left), 3R (centre), and 1T (right), respectively. Atom colour: Red-Mo, yellow-S [39]. Reprinted with permission from ref. [16,38,39]. Copyright 2011-Nature research, Copyright 2021-Elsevier, and Copyright 2016-Cell press.
Figure 2. (a) Lateral view [16], (b) vertical view [38] of MoS2 crystal structure, and (c) Crystal structure of MoS2, 2H (left), 3R (centre), and 1T (right), respectively. Atom colour: Red-Mo, yellow-S [39]. Reprinted with permission from ref. [16,38,39]. Copyright 2011-Nature research, Copyright 2021-Elsevier, and Copyright 2016-Cell press.
Photochem 02 00042 g002
Figure 3. The geometrical representation of Hexagonal MoS2 with their band structure and density of states of space group (a) P63/mmc, (b) P6m2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under copyleft, a CC BY 4.0 open source license.
Figure 3. The geometrical representation of Hexagonal MoS2 with their band structure and density of states of space group (a) P63/mmc, (b) P6m2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under copyleft, a CC BY 4.0 open source license.
Photochem 02 00042 g003
Figure 4. The geometrical representation of trigonal MoS2 with their band structure and density of states of space group (a) R3m, (b) P3m1, (c) P6m2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under a copyleft CC BY 4.0 open source license.
Figure 4. The geometrical representation of trigonal MoS2 with their band structure and density of states of space group (a) R3m, (b) P3m1, (c) P6m2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under a copyleft CC BY 4.0 open source license.
Photochem 02 00042 g004
Figure 5. The geometrical representation of (a) Orthorhombic, (b) Cubic, (c) Tetragonal MoS2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under a copyleft CC BY 4.0 opensource license.
Figure 5. The geometrical representation of (a) Orthorhombic, (b) Cubic, (c) Tetragonal MoS2. Reprinted with Creative Commons Attribution 4.0 License (CC BY 4.0) [46]. This work is licensed under a copyleft CC BY 4.0 opensource license.
Photochem 02 00042 g005
Figure 6. Different methods of preparation of MoS2-based systems.
Figure 6. Different methods of preparation of MoS2-based systems.
Photochem 02 00042 g006
Figure 7. (a) Representation diagram of the synthesis of MoS2 by chemical vapor deposition method [57] (b) Representation for the CVD growth of MoS2 using arched oxidized Mo foil as the precursor [58]. Reprinted with permission from ref. [57,58]. Copyright 2018-MDPI, Copyright 2020-Elsevier.
Figure 7. (a) Representation diagram of the synthesis of MoS2 by chemical vapor deposition method [57] (b) Representation for the CVD growth of MoS2 using arched oxidized Mo foil as the precursor [58]. Reprinted with permission from ref. [57,58]. Copyright 2018-MDPI, Copyright 2020-Elsevier.
Photochem 02 00042 g007
Figure 8. Schematic representation of synthesis of MoS2 using hydrothermal method [60]. Reprinted with permission from ref. [60]. Copyright 2018-Elsevier.
Figure 8. Schematic representation of synthesis of MoS2 using hydrothermal method [60]. Reprinted with permission from ref. [60]. Copyright 2018-Elsevier.
Photochem 02 00042 g008
Figure 9. Schematic diagram of solvothermal synthesis of MoS2 nanosheet aggregates [75]. Reprinted with permission from ref. [75]. Copyright 2018-Nature research.
Figure 9. Schematic diagram of solvothermal synthesis of MoS2 nanosheet aggregates [75]. Reprinted with permission from ref. [75]. Copyright 2018-Nature research.
Photochem 02 00042 g009
Figure 10. (a) Nanoflower [59], (b) Nanosheets [80], (c) Nanosphere [80], (d) Nanotube [81], (e) Nanoribbon [82], (f) Nanoflakes [83]. Reprinted with permission from ref. [59,80,81,82,83]. Copyright 2021-Springer, Copyright 2015-OSA publication, Copyright 2011-Springer, Copyright 2017-Elsevier, Copyright 2014-Hindawi publication.
Figure 10. (a) Nanoflower [59], (b) Nanosheets [80], (c) Nanosphere [80], (d) Nanotube [81], (e) Nanoribbon [82], (f) Nanoflakes [83]. Reprinted with permission from ref. [59,80,81,82,83]. Copyright 2021-Springer, Copyright 2015-OSA publication, Copyright 2011-Springer, Copyright 2017-Elsevier, Copyright 2014-Hindawi publication.
Photochem 02 00042 g010
Figure 11. (a,b) TEM images of MoS2-based system. Reprinted with permission from ref. [84,85] 2015-Nature research, Copyright 2013-Springer.
Figure 11. (a,b) TEM images of MoS2-based system. Reprinted with permission from ref. [84,85] 2015-Nature research, Copyright 2013-Springer.
Photochem 02 00042 g011
Figure 12. (a) XRD plot of MoS2, (b) FTIR plot of MoS2, (c,d) TGA curves of MoS2, (e) UV-DRS graph of MoS2, (f) Tauc plot of MoS2. Reprinted with permission from ref. [59,81,86,87,88]. Copyright 2011-Springer, Copyright 2019-Springer. Copyright 2020-Springer, Copyright 2017-American Chemical Society. Copyright 2020-Springer.
Figure 12. (a) XRD plot of MoS2, (b) FTIR plot of MoS2, (c,d) TGA curves of MoS2, (e) UV-DRS graph of MoS2, (f) Tauc plot of MoS2. Reprinted with permission from ref. [59,81,86,87,88]. Copyright 2011-Springer, Copyright 2019-Springer. Copyright 2020-Springer, Copyright 2017-American Chemical Society. Copyright 2020-Springer.
Photochem 02 00042 g012
Figure 13. The optimized structure of TiO2/g-C3N4/MoS2 nanocomposite. (Blue-N, Red-O, White-Ti, Grey-C, Yellow-Mo, Cyan-S) Reprinted with permission from ref. [111]. Copyright 2018-Elsevier.
Figure 13. The optimized structure of TiO2/g-C3N4/MoS2 nanocomposite. (Blue-N, Red-O, White-Ti, Grey-C, Yellow-Mo, Cyan-S) Reprinted with permission from ref. [111]. Copyright 2018-Elsevier.
Photochem 02 00042 g013
Figure 14. The mechanism of the MoS2-based photocatalyst on the dye (a) traditional mechanism in binary system MoS2/Cu2O [144], (b) Z-scheme mechanism in binary system MoS2/g-C3N4 [145], (c) Traditional mechanism in ternary system MoS2/CdS/TiO2 [116], (d) Z-scheme in ternary system MoS2/g-C3N4/TiO2 [59], and (e) s-scheme in ternary system MoS2/g-C3N4/TiO2 [112]. Reprinted with permission from ref. [59,112,116,144,145]. Copyright 2021-Springer, Copyright 2020-Elsevier, Copyright 2019-Elsevier, Copyright 2020-Springer, Copyright 2017-Elsevier.
Figure 14. The mechanism of the MoS2-based photocatalyst on the dye (a) traditional mechanism in binary system MoS2/Cu2O [144], (b) Z-scheme mechanism in binary system MoS2/g-C3N4 [145], (c) Traditional mechanism in ternary system MoS2/CdS/TiO2 [116], (d) Z-scheme in ternary system MoS2/g-C3N4/TiO2 [59], and (e) s-scheme in ternary system MoS2/g-C3N4/TiO2 [112]. Reprinted with permission from ref. [59,112,116,144,145]. Copyright 2021-Springer, Copyright 2020-Elsevier, Copyright 2019-Elsevier, Copyright 2020-Springer, Copyright 2017-Elsevier.
Photochem 02 00042 g014
Figure 15. PL spectra and their mechanism (a) Traditional mechanism of MoS2@ TiO2 [86], (b) Z-scheme mechanism of MoS2/CoFe2O4 [134]. Reprinted with permission from ref. [86,134]. Copyright 2020-Elsevier, Copyright 2019-Elsevier.
Figure 15. PL spectra and their mechanism (a) Traditional mechanism of MoS2@ TiO2 [86], (b) Z-scheme mechanism of MoS2/CoFe2O4 [134]. Reprinted with permission from ref. [86,134]. Copyright 2020-Elsevier, Copyright 2019-Elsevier.
Photochem 02 00042 g015
Table 1. Comparison of the photocatalytic degradation of various dye using MoS2-based catalysts.
Table 1. Comparison of the photocatalytic degradation of various dye using MoS2-based catalysts.
SystemDyeReaction ConditionsDegradation (%)Ref.
Time (Min)Light SourceCatalyst Amount (mg)Conc of Dye (ppm)
MoS2/SnO2MB90Visible20 95.0[30]
MoS2 nanosheetsMB120Fluorescent lamp1010049.3[121]
CeO2-ZrO2@MoS2NPX40Visible light 5096.0[142]
CeO2/MoS2MO90Visible light252096.1[106]
3D/2D CeO2/MoS2MV30Visible light202096.25[107]
Cerium-doped MoS2Cr(IV)30Visible light62040[108]
MoS2/BiOBrRhB30Xenon arc201096.0[70]
MoS2-x nanosheet arraysRhB60Visible 597.2[122]
10% g-C3N4/TiO2/MoS2(0.2)MB60Visible301099.5[110]
Layered MoS2MB90Visible501071.0[133]
TiO2/SnS2/MoS2MB90Visible 581.8[135]
TiO2/g-C3N4/MoS2MO60Visible1002090.6[111]
g-C3N4/MoS2/TiO2MO60Visible500.0388.0[112]
MoS2/g-C3N4/TiO2MG60Visible501086.0[59]
TiO2/MoS2RhB
MB
MO
60Visible501099.6
96.4
87.5
[100]
RGO-MoS2 supported NiCo2O4RhB90Visible501095.0[91]
SrZrO3/Flower-likeMoS2RhB80solar152599.7[104]
g-C3N4/Ag/MoS2RhB90Xenon arc1002095.8[114]
MoS2/NiFe LDHRhB120Solar202090.0[92]
MoS2/COFRhB30Solar102098.0[136]
MoS2–GOMB60solar101099.0[103]
BiPO4, MoS2 and grapheneRhB90Mercury1005-[137]
g-C3N4/WO3/MoS2RhB
MO
MB
Xenon arc10050
20
20
99.9
83.4
91.8
[115]
MoS2/CdS/TiO2MB15Visible2510100[116]
g-C3N4-based MoS2 and Bi2O3MB90Visible502098.5[117]
MoS2/ZnSMB32Visible501099.9[105]
MoS2/CdIn2S4RhB30Visible 10-[138]
CoFe2O4/MoS2CRMBMO60Visible302094.9
67.8
[139]
ZnO/MoS2MB300Visible10175.0[99]
Type II MoS2/ZnOMB120UV 1099.0[101]
BiOIO3/MoS2 (BM-x) 2D/2DRhB90500 W Xenon lamp501098.7[140]
ZnO-MoS2MR60Solar101089.0[141]
BiFeO3/MoS2RhB200Visible501089.0[102]
ZnO nanorods/MoS2RhB UV-4-[123]
Ultrathin layered MoS2MB
RhB
36Visible100595.3
41.1
[124]
MoS2/TiO2MB12UV-Vis50599.3[125]
TiO2/MoS2MB30Visible101094.2[126]
NMS incorporateda-Fe2O3/ZnORhB240Visible40 91.0[127]
Ag3PO4/MoS2MB1560 W CFL0.20 g/L2097.6[128]
MoS2/g-C3N4RhB/MO60Visible55092.0[98]
PbS/MoS2MB180Visible1%3083.0[129]
MoS2QDs@ZnIn2S4@RGO)RhB/MB30300 W Xenon lamp1008098.8
98.5
[130]
SnO2-MoS2MR/MB120Visible110058.5
94.0
[131]
MoS2-SrFe12O19RhB120300 W Xenon
lamp
1096.5[132]
Au-CoFe2O4/MoS2MO60300 W iodine tungsten lamp705096.0[134]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jaleel UC, J.R.; R, M.; Devi K R, S.; Pinheiro, D.; Mohan, M.K. Structural, Morphological and Optical Properties of MoS2-Based Materials for Photocatalytic Degradation of Organic Dye. Photochem 2022, 2, 628-650. https://doi.org/10.3390/photochem2030042

AMA Style

Jaleel UC JR, R M, Devi K R S, Pinheiro D, Mohan MK. Structural, Morphological and Optical Properties of MoS2-Based Materials for Photocatalytic Degradation of Organic Dye. Photochem. 2022; 2(3):628-650. https://doi.org/10.3390/photochem2030042

Chicago/Turabian Style

Jaleel UC, Jadan Resnik, Madhushree R, Sunaja Devi K R, Dephan Pinheiro, and Mothi Krishna Mohan. 2022. "Structural, Morphological and Optical Properties of MoS2-Based Materials for Photocatalytic Degradation of Organic Dye" Photochem 2, no. 3: 628-650. https://doi.org/10.3390/photochem2030042

Article Metrics

Back to TopTop