Next Article in Journal
Applications of Graphene and Its Derivatives in the Upstream Oil and Gas Industry: A Systematic Review
Next Article in Special Issue
Measurement of Quantum Yields of Monolayer TMDs Using Dye-Dispersed PMMA Thin Films
Previous Article in Journal
Signal-On Fluorescent Imprinted Nanoparticles for Sensing of Phenols in Aqueous Olive Leaves Extracts
Previous Article in Special Issue
High-Yield Production of Few-Layer Graphene via New-fashioned Strategy Combining Resonance Ball Milling and Hydrothermal Exfoliation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transition Metal Dichalcogenides for the Application of Pollution Reduction: A Review

1
State Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
2
Central European Institute of Technology, Brno University of Technology, Purkynova 656/123, 612 00 Brno, Czech Republic
3
Institute of Process Engineering & NETME Centre, Brno University of Technology, Technicka 2896/2, 616 69 Brno, Czech Republic
4
Department of Chemical Engineering, Monash University, Clayton, Melbourne 3800, Australia
5
Research Centre for Sustainable Technologies, Faculty of Engineering, Computing and Science, Swinburne University of Technology, Jalan Simpang Tiga, Kuching 93350, Malaysia
6
Department of Chemical and Environmental Engineering, University of Nottingham, Semenyih 43500, Malaysia
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(6), 1012; https://doi.org/10.3390/nano10061012
Submission received: 5 May 2020 / Revised: 18 May 2020 / Accepted: 19 May 2020 / Published: 26 May 2020
(This article belongs to the Special Issue Characterization, Synthesis and Applications of 2D Nanomaterials)

Abstract

:
The material characteristics and properties of transition metal dichalcogenide (TMDCs) have gained research interest in various fields, such as electronics, catalytic, and energy storage. In particular, many researchers have been focusing on the applications of TMDCs in dealing with environmental pollution. TMDCs provide a unique opportunity to develop higher-value applications related to environmental matters. This work highlights the applications of TMDCs contributing to pollution reduction in (i) gas sensing technology, (ii) gas adsorption and removal, (iii) wastewater treatment, (iv) fuel cleaning, and (v) carbon dioxide valorization and conversion. Overall, the applications of TMDCs have successfully demonstrated the advantages of contributing to environmental conversation due to their special properties. The challenges and bottlenecks of implementing TMDCs in the actual industry are also highlighted. More efforts need to be devoted to overcoming the hurdles to maximize the potential of TMDCs implementation in the industry.

Graphical Abstract

1. Introduction

Transition metal dichalcogenides (TMDCs) are a large family of two-dimensional (2D) layered materials, which are scientifically interesting and industrially important. These materials have attracted tremendous attention because of the unique structural features and interesting properties, such as optoelectronics, electronics, mechanical, optical, catalytical, energy-storage, thermal, and superconductivity properties [1,2,3,4,5,6,7]. TMDCs are the compounds of the chemical formula MX2, where M is a transition metal element of groups IV-VII B (Mo, W, V, Nb, Ta, Ti, Zr, Hf, Tc, and Re) and X is a chalcogen element (S, Se, and Te). The X-M-X unit layer consists of three atomic layers, in which one centre atom layer (M) is sandwiched between two chalcogen atom layers (X). TMDCs occupy the layered structures, which resemble that of graphite. The interlayers are stacked by weak van der Waals force, leading to the formation of monolayers or nanolayers from the bulk materials via exfoliation [8]. Different stacking of the layers along c-axis determines polymorphic crystal structures in TMDCs, and the common phases are 1T, 2H, 3R, and Td phases (T—trigonal, H—hexagonal, R—rhombohedral, and Td—distorted octahedral) [9].
There are more than 40 different TMDC types, including metals (such as TiS2 and VSe2), superconductors (such as TaS2 and NbS2), semimetals (such as MoTe2 and WTe2), and semiconductors (MoS2, MoSe2, WS2, and WSe2). TMDCs exhibit interesting band structures with tunable bandgaps. The bandgap is one of the most important factors in 2D materials for determining the properties and applications. For instance, graphene is a semimetal with zero bandgap, which limits its applications in electronics and photo-electronics. TMDCs exhibit variable bandgaps from 0 to 3 eV, which can be tuned by thickness [10], defects [11], dopants [12], and mechanical deformations (by applying the tensile strain or compressive strain) [13,14]. The most studied semiconducting TMDCs (e.g., MoS2, MoSe2, WS2, and WSe2) have shown typical features in electronic structures. The bandgap increases with the decreasing thickness and it possesses the transition from indirect in the bulk crystals to direct in the monolayers [10,15]. For instance, the indirect bandgap of −1.29 eV will be changed to a direct bandgap of −1.8 eV when bulk MoS2 is down to a monolayer [16].
Benefiting from their unique crystal structures and electronic structures, TMDCs have shown great potential in various fields, including electronics/optoelectronics [1,17], catalysis [18], energy storage [19] and conversion [20], sensing [21,22], and so on. The application of TMDCs in pollution reduction is a compelling research topic. The increasing environmental pollution issue has been one of the most serious problems on Earth. Enormous efforts have been made to search the efficient and low-cost methods for addressing the environmental pollution issue. TMDCs may be a kind of promising materials for tackling these problems with several advantages. Firstly, TMDCs have a high surface-to-volume ratio. They offer more effective active sites on the surface, as well as abundant unsaturated surface sites. Thus, the layered TMDCs are excellent platforms for the anchor of semiconductor nanoparticles in various photocatalytic applications [23,24]. Due to their high surface-to-volume ratio, TMDCs are extremely sensitive to the surrounding atmosphere and can be utilized in toxic gas sensing and adsorption. Secondly, TMDCs have tunable bandgaps, which enhances the photocatalytic performance in nanocomposite by offering appropriate bandgap and band alignment [25]. Thirdly, defect engineering can be easy to implement in 2D materials, which have been confirmed to be an efficient method for intensifying the catalytic activities in TMDCs [26,27,28,29]. Lastly but importantly, there is a large variety for TMDCs (about 40 kinds) and they have an abundant amount in nature or can be synthesized [9]. So far, MoS2, WS2, MoSe2, and ReS2 have been naturally found [30,31,32]. Specifically, MoS2 exists as molybdenite in nature and is the main source of molybdenum with a large amount [33]. The main metals (W and Mo) in TMDCs are both abundant, cheap, and widely used in industry [34]. TMDCs can be prepared by using various techniques, such as chemical vapor deposition (CVD) [35,36], chemical vapor transport (CVT) [37,38], flux growth method [39], hydrothermal synthesis [40], Langmuir−Schaefer deposition [41], etc. In addition, the top-down exfoliated method can be also used to fabricate few-layer TMDCs from bulk crystals, e.g., mechanical exfoliations and liquid phase exfoliations [42,43,44]. With increasing interests in TMDCs for applications, we aim to prepare an overview of the recent progress of TMDCs in reducing the environmental pollution. We will summarize the representative efforts, including gas adsorption and removal, gas sensing, wastewater treatment, fuel cleaning, CO2 valorization, and conversion.

2. Gas Adsorption and Removal

In recent years, TMDCs have been used as gas removal (via adsorption) in pollution reduction. Figure 1 presents the number of research articles that involved the use of TMDCs in gas adsorption applications. The increasing trend (the total number of related articles published in the year 2019 is more than three-fold than that of the year 2015) indicates the potential of TMDCs in gas removal processes. Based on the bibliology search, MoS2 is still the most widely studied TMDCs among the TMDCs considered in this review. However, researchers started to realize that there are other TMDCs (e.g., MoSe2 and WS2) that offer better performances as compared to MoS2 (e.g., greater charge storage ability [45]). This further led to a gradual increase in research articles that studied the use of other potential TMDCs, since the year 2018. Figure 2 outlines the schematic diagrams of each gas adsorption process. In general, the mechanism of these adsorption processes is mainly driven by the charge transfer process between the adsorbates and the TMDCs-based adsorbent, where the charge movement is dependent on the nature of the gas adsorbate (i.e., oxidizing or reducing) [46]. For instance, carbon monoxide (CO) in Figure 2a is an example of reducing gas. Due to the existence of lone pair on the carbon atom, CO will donate electrons to the TMDC surface, which further cause the CO to be chemisorbed on the surface. Whereas, oxidizing gas such as nitrogen dioxide (NO2) in Figure 2c, will uptake the electron from the surface instead (mainly due to the existence of unpaired electron on the nitrogen atom). To note, such electron movement will lead to the deviation in electrical conductivity of the TMDC materials [47]. The following subsections discuss various gas adsorption processes.

2.1. Adsorption of Carbon Monoxide (CO)

Carbon monoxide (CO) is by far the most hazardous greenhouse gases, which is 210 times easier to bind with hemoglobin as compared to the oxygen [49]. To date, numerous studies have discovered the use of TMDCs as a catalyst to adsorb CO and further convert it into other products via catalytic reaction. For instance, Li et al. [50] proposed the use of aluminum oxide-doped MoS2 as the nanocatalysts to promote the CO methanation process (CO + 3H2 → CH4 + H2O) and enhance the stability of this catalytic reaction (MoS2 with 25.6% of Al2O3 provides the greatest methanation stability). The schematic diagram of this process is shown in Figure 2a. The incorporation of metal-based promoters, in this case, Al2O3 powder has effectively reduced the aggregation effect between MoS2 (lower tendency to pore blockage) [50]. Aside from that, researchers also adopted the density functional theory (DFT) calculation to explore the theoretical potential of each TMDC as the alternative catalyst for CO oxidation (CO + O2 → CO2 + O) and CO dissociation. For instances, studies showed that metal-doped MoS2 (e.g., Au29, Cu, Ag, Co, Rh, Ni, Ir, and Fe [51,52,53]) can significantly enhance the O2 dissociation (which is the essential step for CO oxidation). On the other hand, the latter five mentioned metals possess the greatest potential as they favor CO dissociation [52]. More recently, application of other TMDCs (e.g., Pt and Au nanoclusters with WSe2, where the carbon atom in CO will be strongly adsorbed on the Pt/Au decorated WSe2 monolayer [54]; Rh-doped MoSe2 [55]), in CO adsorption have also been carried out. Since CO dissociation is the first step of the Fischer–Tropsch process [56], the aforementioned TMDC-based nanocatalysts can be served as the substituent cheap and stable catalyst to convert CO into other valuable liquid hydrocarbons.

2.2. Adsorption of Water Vapor (H2O)

Water vapor (H2O) is the largest contributor to the current global warming issue (i.e., roughly accounted for 60% of the entire warming effect [57]). Numerous works have proposed the use of TMDC-based nanocatalysts to convert H2O into clean hydrogen (H2) via hydrogen evolution reaction (HER) (see Figure 2b). In general, hydrogen atoms in H2O will be adsorbed on the active site of the HER nanocatalysts. With the aid of the doped metal, the dissociation of the O–H bond in H2O is enhanced. The gas is then desorbed as H2 [58,59]. This can mitigate the warming effect attributed to the water vapor content in the atmosphere, and at the same time, serves as an alternative greenway for hydrogen production. The past two decades ago, MoS2 was referred to as one of the most prominent alternatives to substitute the conventional HER catalysts [60]. Nevertheless, the commercialization of TMDC-based HER nanocatalysts is still hindered due to various technical challenges (e.g., poor HER stability under acidic conditions [61,62] and weak intrinsic conductivity [63]). Numerous works have discovered ways to enhance the HER performance of MoS2. Just to name a few, these works proposed to improve the HER activity by (i) coating the MoS2 with metals (e.g., palladium [64], aluminum [65], gold [66], platinum [67], etc.); (ii) using a novel electrochemical approach to deposit the MoS2 nanosheet [68]; (iii) introducing the use promising support for the MoS2 nanoparticles (e.g., multi-wall carbon nanotubes [69]); (iv) varying the choice of dispersion media used [70]; and (v) coupling with other electrocatalysts (e.g., molybdenum carbide [71]). On the other hand, some researchers also proposed the use of other TMDCs in HER applications. For instance, Wang et al. [72] proposed to use a low-cost MoSe2 nanosheet, which offers decent active sites for HER activity and better HER performance as compared to the aforementioned MoS2-based catalysts. On the other hand, Seok et al. [73] have reported the intrinsic activity for HER on MoTe2 nanosheet via the first-principle calculation and scanning tunneling microscopy (STM) study. Besides, numerous researchers also successfully developed promising WS2-based and WSe2-based electrocatalysts recently (e.g., Te–WS2 nanosheet by Pan et al. [74]; and Ni-WSe2 nanosheet by Kadam et al. [75]).

2.3. Adsorption of Carbon Dioxide (CO2)

Aside from water vapor, carbon dioxide (CO2) is another key contributor to climate change issues (accounting for more than 75% of the total emitted greenhouse gases [76]). To address this issue, some researchers have attempted to apply MoS2 nanosheets as an effective membrane [77] and adsorbent [78,79] to separate CO2 from the gaseous mixture. Sun et al. [78] reported that the adsorption force between CO2 and MoS2 varies according to the strength of the applied electric field. Generally, CO2 will adsorb on the surface when an electric field is applied and desorb when the electric field is relieved (see Figure 2d). This unique feature makes it become a potential carbon capture media. In addition to the carbon capture process, TMDCs can be modified so that the captured CO2 can be catalytically converted into other CO2-reduction products. Shi et al. [80] were the first to discover the potential of copper modified MoS2 nanosheets in the CO2 reduction process. The incorporation of Cu nanoparticles has enhanced the adsorption capability as CO2 can now be adsorbed not only on the MoS2 nanosheets but also on the doped metal nanoparticles [80] (see Figure 2d). This work found that most of the CO2 was converted into CO (i.e., faradaic efficiency (FE) = 33–41%), followed by methane (FE = 7–17%) and with a trace amount of ethylene (FE = 2–3.5%) [80]. More recently, a research project in Korea has discovered the potential of a composite catalyst that encompasses MoS2 nanosheets and n-type Bi2S3, in the CO2-to-CO photoreduction process [81]. DFT calculations of various modified-TMDC nanosheets (e.g., SnO2-loaded MoS2 nanosheet [82]; TiO2-doped MoS2 nanosheets [83]) for the CO2 reduction process were also conducted recently. Nevertheless, the research on the potential of other TMDCs in CO2 adsorption is still considered scarce. To date, only a few have discovered their potential for the CO2 sensor [84,85].

2.4. Sulphur Content Removal

Based on the air quality data observed by the NASA Ozone Monitoring Instrument (OMI) satellite, about 60% of the total sulphur emissions in 2018 were anthropogenic emissions [86]. Despite the recent advancement in renewable energy and air treatment technologies have reduced the global sulphur emissions (sulphur emissions in 2015 is about 30% less as compared to that of 1990 [87]), the health impact of sulphur content in the atmosphere can still be severe. Thus, the sulphur emissions (including SO2, H2S, etc.) must still be monitored and controlled closely. Wei et al. [56], based on the DFT simulation, have explored the adsorption performance of SO2 and H2S on the Ni-doped MoS2 monolayer material. To note, due to the large number of free electrons offered by the doped metals, the overall adsorption capability of the TMDC-based adsorbents are, therefore, gradually enhanced [88]. This study is then extended by other researchers, by testing the effect of other doped metal atoms (e.g., palladium [89], platinum [90], gold [89], and copper [91]). All these studies show that the metal-doped MoS2 monolayer materials can offer promising adsorption properties to SO2 and H2S gases, where both gases are chemisorbed on the surface with strong interactive forces. Similarly, various DFT studies have also been conducted to study the potential of using metal-doped MoSe2 nanosheets (e.g., rhodium-doped MoSe2 monolayer material [55] and palladium-doped MoSe2 monolayer material [92]) and WSe2 nanosheets (e.g., N-doped TiO2/WSe2 nanocomposite [93] and silver-doped WSe2 monolayer material [94]) in sulphur content removal applications. Dan et al. [95], on the other hand, studied the photocatalytic H2S splitting (H2S → H2 + S) on the proposed novel composite catalysts, MnS/In2S3-MoS2 based on first-principles calculations. This is an attractive and cost-effective way to generate clean H2 that can substitute conventional coal-derived hydrogen production [96]. In other words, this also indirectly contributes to pollution mitigation. The schematic diagram for the aforementioned sulphur removal process is shown in Figure 2e, where the desorption process can be conducted by altering the operating conditions (e.g., temperature and pressure).

2.5. Nitrogen Oxide (NOx) Removal

Nitrogen oxide (NOx) has adverse effects on human health and ecosystems. To date, numerous works have revealed the potential of applying TMDCs for NOx removal purposes. Notably, a novel Cu2O-anchored MoS2 nanocomposite developed by Yuan et al. [97], can achieve up to 53% NOx removal. In another research conducted in China, a similar NO removal rate (about 51%) can also be obtained using the MoS2-g-C3N4 nanocomposite [98]. In the same year, Xiong et al. [99] managed to develop a novel (BiO)2CO3/MoS2 photocatalysts that could achieve 50% NO removal for five consecutive cycles. The study was extended by incorporating the use of carbon nanofibers into nanocomposites fabrication [100]. The developed Bi2O2CO3-MoS2-CNFs nanocomposites can attain almost 70% of NO removal rate. Under visible light irradiation, the oxidative power offered by the nanocomposites is sufficient to oxidize the NOx into NO3− (see Figure 2c). Besides, first-principle studies were also conducted to determine the metal-doping effects on the adsorption capabilities on MoS2 (e.g., (vanadium, niobium, tantalum)-doped MoS2 [101]), MoSe2 (e.g., palladium-doped MoSe2 [92] and rhodium-doped MoSe2 [55]), and WSe2 [102] monolayer materials to NOx gases.

3. Gas Sensing for Pollution Reduction

Gas pollution is commonly caused by direct greenhouse gases (CO, CO2, N2O, CH4, Fluorinated Gas, etc.) [103], indirect greenhouse gases (NH3, NOx, H2S, etc.) [104], and other traces of toxic gases. These gas pollutants may cause climate change, ozone pollution, and even threaten food security [105] Further adverse effects that are caused by these problems include reduced photosynthesis performances [106], increased chances of respiratory problems [107], and acid rain [108].
Gas sensing can be deployed in multiple locations to act as an environmental monitoring system [109]. More recent applications of gas sensing products include gas monitoring using drones [110], wearable gas sensing devices [111], and the internet of things (IoT) multi-gas sensing modules [112]. For gas sensing TMDC-material in the device, the most common working principle is for adsorption of gas particles towards the TMDCs for a charge transfer, giving a change of resistivity in material, then desorption from the TMDCs [113]. In this field, researchers are constantly searching for TMDC materials that can provide a very low limit of detection (LOD), high sensitivity, short response and recovery time [114,115,116].

3.1. NOx Detection

NOx detection by TMDCs has received much research interest due to its effectiveness and applicability. The measurement of NOx in environmental monitoring is a common requirement [117] while the design of a perfect NOx sensor poses some challenges. Traditional SnO2 sensors have problems with the dual response towards oxidizing gases and reducing gases, giving low sensitivity when performing measurements in a multi-gas environment [118]. The selectivity of the gas between NO2 and NO is also a crucial challenge for NOx sensors [119].
A popular TMDC material for NO2 detection is MoS2 [113,120,121]. Earlier works [121] used CVD-grown MoS2 to detect NO2 gas at room temperature with LOD down to parts per billion (ppb) ranges, however, the recovery time was long. Consequently, a higher temperature was used to accelerate the desorption kinetics of the NO2 gas molecule [113,120]. Other researchers also created nanocomposites of TMDCs with materials such as graphene aerogel [122] and reduced graphene oxide (rGO) [123] to improve sensitivity and lower LOD. Pham et al. [115] used a red-light illumination to provide photon energy, which matches the bandgap of the MoSe2 sensor. The technique resulted in low LOD (25 ppb) in room temperature conditions with high sensitivity. Moreover, Liu et al. [114] synthesized a flower-like porous SnS2 NSs with edge exposed MoS2 nanosphere, which resulted in a very fast response time of 2 seconds. Group-10 noble TMDCs such as PtSe2 has also received much research attention due to their widely tunable bandgap and excellent performance in gas sensing [124]. From a first-principle study, Sajjad et al. [125] discovered that monolayer PtSe2 exhibited lower adsorption energy than MoS2 and graphene. For this, Yim et al. [126] used thin films of PtSe2 stacked on Si to achieve NO2 detection down to 9 ppb. Other TMDC materials such as WS2 [122] and MoTe2 [127] also showed that they have excellent NO2 sensing properties, and more research work is to be expected for different transition metals. A collection of NO2 sensing TMDC materials is tabulated in Table 1.
Although NO and NO2 are sometimes measured as total NOx, many academic studies require differentiation between NO from NO2 [133,134]. A distinct property between the two gases is that NO is a colorless gas while NO2 is reddish-brown in color [135]. Due to its reduced electron valence in NO, first-principle studies demonstrated that the adsorption energy of NO to be slightly higher than that of NO2 [136]. Nevertheless, the binding of NO with MoS2 is still considered one of the strongest in many gas molecules including CO, CO2, NH3, CH4, H2O, N2, O2, and SO2 [136]. This computational finding was verified by a few experiments that successfully detected NO with MoS2 materials down to the ppm level [137,138,139]. Although there is not much research work that focuses on NO detection, it is demonstrated to give promising performances using TMDC material (see Table 2).

3.2. Ammonia Detection

Ammonia is an indirect greenhouse gas, as it is quickly soluble by water vapor or rain in the atmosphere and goes down to the land to be readily used as fertilizers by plants [141]. The lifespan of this alkaline and reactive gas in the atmosphere is short, around 1 day [142]. Nevertheless, direct exposure to ammonia gas from its emission source can still cause health issues to humans [143]. Gases such as NO2 are electron-accepting, while NH3 donates electrons to the TMDCs surface [113] to give a change of resistivity in the material (see Figure 3a,b). Due to their different adsorption mechanism, NO2 and NH3 are two flagship gases to study the generic effect of gas sensing [113,144,145].
Many works have also indicated that TMDC materials have excellent properties to act as excellent ammonia gas sensors [145]. For example, Cho et al. [146] used a 2D graphene/MoS2 heterogeneous structure to improve the sensitivity of the sensor towards gas molecules. Burman et al. [139] studied the effect of vacancy sites on MoS2 sensors and proposed a UV-treated sensor that can be synthesized from powder. Other TMDC materials such as MoTe2 [147] and WS2 [148] also showed performances of being able to detect ammonia gases down to ppm levels. Even artificial neural networks were used to post-process signals from TMDC ammonia gas sensors to improve gas concentration predictions [144]. A few of the significant works of applying TMDC materials as ammonia gas sensors are found in Table 3.

3.3. Volatile Organic Compound (VOC) Detection

Volatile organic compounds (VOC), commonly known as aromatic hydrocarbons fractions, are organic chemicals that have high vapor pressure in atmospheric conditions. VOC are studied to be emitted in landfills [153], newly renovated buildings [154], industrial refineries [155], and other leakage sources. Long-term exposure to VOC can lead to dysfunction of central nervous systems, memory loss, and cause congenital anomalies for reproduction [156]. VOC sensors are important for environmental detection [154], industrial emission control [157], and even for cancer diagnosis [158].
The challenge of designing sensors for VOC is that VOC is not a single gas molecule but can consist of multiple molecules (such as benzene, ethylene glycol, formaldehyde, etc.) with slightly different properties. Barzegar et al. [159] represented the VOC molecule behavior by using xylene isomers and methanol, showing high potential for adsorption between VOC molecules and a Ni-decorated MoS2 sensor. A thiolated ligand conjugate MoS2 sensor was also shown to exhibit excellent sensitivity down to a concentration of 1 ppm [160]. The work represented VOC by studying a combination of toluene, hexane, ethanol, propanal, and acetone gas molecules while demonstrating the conjugation of thiolated ligand improves the charge carrier density of the sensor (see Figure 4). A recent work from Tomer et al. [161] demonstrated that a cubic Ag(0)-MoS2 loaded g-CN 3D porous hybrid had multifunctional abilities towards sensing VOC by studying n-butanol, isopropanol, benzene, and xylene gas molecules. Short response time and recovery time of 7–15 s and 6–9.5 s respectively were obtained for all VOC molecules at 175 °C. Zhao et al. [162] presented an optical VOC microsensor using a photonic crystal cavity integrated with MoS2. VOC molecules of acetone, methanol, ester, dichloromethane, and methylbenzene were studied and the optical microsensor resulted in a LOD of 2.7 ppm, response time of 0.3 s, and recovery time of 100 s.

3.4. Detection of Sulphur Gases and Other Gases

The unique properties of TMDC materials also showed many preliminary potentials for the detection of sulphur gases and other emissions. Chen et al. [163] demonstrated from density functional theory (DFT) that MoS2 monolayers exhibit great adsorption energy towards SF6 decomposition (which includes SO2, SOF2, SO2F2, H2S, and HF). The adsorption of SF6 decomposition by monolayer PtSe2 was also studied by DFT methods and found to give excellent adsorption energies [164]. Park et al. [165] experimentally demonstrated that a Pt nanoparticle decorated MoS2 gas sensor can achieve high sensitivity detection of H2S down to 30 ppm. Recent work from Yang et al. [166] demonstrated that Ni- and Cu-embedded MoS2 monolayers can improve the adsorption energy of SOx and O3 molecules compared to that of the pristine MoS2.
TMDCs also demonstrate adsorption and sensing ability for CO gas. The work of Ma et al. [167] studied the potential of sensing CO and NO gas molecules by doping metal particles on MoS2 monolayer. The doping of Au, Pt, Pd, and Ni has shown to alter the transport property of MoS2, giving better adsorption energy for CO and NO gas detection. Recently, Yang et al. [91] presented a strategy of using Ti doping and the application of an electric field to improve adsorption energies in MoSe2 and MoS2 materials. The work concluded that Ti doping strategy improved the sensitivity of both materials towards CO and NO gas detection. Another recent work from Shen et al. [168] provided a study of using a borophene/MoS2 heterostructure to detect small gas molecules (CO, NO, NO2, and NH3), which exhibited different resistive properties towards these molecules when changing voltage direction (see Figure 5). Modulation of MoS2 by antisite doping and strain also showed improved gas detection sensitivity [169]. Ma et al. [170] studied the effects of defects on WSe2 monolayer sensor and found that Se vacancies can improve sensing ability of H2O and N2 molecules. Further interest gas detection also extends towards Cl2, PH3, AsH3, BBr3, and SF4 gas molecules on MoS2 [171]. More materials will be expected to be discovered for gas detection applications in the future.

4. TMDC Materials for Wastewater Treatment

The United Nations has highlighted that 80% of wastewater is released back to the environment without sufficient treatment [172]. The effect of industrial development and human activities has released many pollutants into the water. The major water pollution sources come from various source such as industry waste [173], sewage and wastewater [174], oil leakage [175], chemical fertilizer and pesticides [176], and mining activities [177]. These pollutants require high oxygen for oxidation decomposition, which reduces the dissolved oxygen level in the water that will damage the aquatic ecosystem [178]. Wang et al. [179] reviewed that many technologies including adsorption, ion exchange, membrane filtration, chemical precipitation, and electrochemistry have been widely used in water treatment. Despite the application of various technologies, quality water supply remains unsustainable [180]. Many researchers have reported promising outcomes with the application of nanomaterial such as TMDC-based material in water treatment technology [181].

4.1. Adsorption for Wastewater Treatment

Among wastewater treatment technologies, the adsorption is considered the most inexpensive, fast and simple operation method [182]. Recently, researchers are focusing on the application of TMDC-based material in adsorption technology. That MoS2 nanosheet is used in the adsorption process mainly due to its large surface area, excellent chemical and thermal stability and environmentally friendly [183]. Many application of TMDC-based material has been successfully demonstrated with an effective outcome including the removal of methyl orange [184], Rhodamine B (RhB) [185], and Congo Red [186]. Li et al. [187] stated that the application of MoS2 adsorbent can increase the adsorption capacity in removing organic dye. An alternative facile oxidation strategy was proposed by Li et al. [182] to synthesize tungsten disulphide/tungsten trioxide (WS2/WO3) heterostructures to remove RhB molecules from wastewater (see Figure 6a). This strategy has contributed to higher adsorption capacity through manipulating the surface property. Massey et al. [188] further added that the hierarchical microsphere of MoS2 nanosheets has demonstrated high adsorption capacity with 297, 216, and 183 mg/g of methylene blue, rhodamine 6G, and fuchsin acid dye respectively as compare to conventional absorbent such as activated carbon. Most importantly, the hierarchical microsphere of MoS2 nanosheets can be regenerated effectively without affecting the performance of adsorption capacity.
Kumar et al. [189] concluded that the synthesis of MoS2 with magnetic nanoparticles can be effective adsorption material especially in removing heavy metal such as chromium. The efficiency of chromium removal is highly dependent on the pH of the solution and the adsorption of Cr(VII) and Cr(III) can be selected by changing the pH (see Figure 6b). Besides that, the regeneration of absorbent does not reduce the efficiency of chromium uptake significantly.
On top of that, the CeO2-MoS2 hybrid magnetic biochar (CMMB) also exhibits a strong magnetic ability to remove lead(II) (Pb(II)) and humate from water treatment [191]. The CMMB can remove >99% of Pb(II) and humate within 6 h. The application of MoS2/thiol-functionalized multiwalled carbon nanotube (SH-MWCNT) has also displayed high adsorption capacity for heavy metal removal [177]. However, the spent adsorbent can be further used for photocatalytic and electrochemical applications.
Furthermore, the synthesized MoS2-coated melamine-formaldehyde (MF@MoS2) sponges that exhibit a superhydrophobic and superhydrophilic characteristic (see Figure 6c) demonstrated a highly selective adsorption capacity [190]. The MF@MoS2 sponges have high absorption performance for oil and organic solvent and water-soluble dye. It also exhibits high discoloration efficiency of 98% methyl orange within 10 min. Wan et al. [192] added that the MF@MoS2 can be modified from room temperature vulcanized silicon rubber.

4.2. Membrane Technology in Wastewater Treatment

In water treatment systems, the typical microporous membrane pore size is about 0.1–5 µm, which limits membrane technology on water purification [179]. The development of nanoporous membranes exhibited high filtering performance in dealing with most of the pollutants inducing microbes, organic molecules, heavy metal, and salts.
Heiranian et al. [193] highlighted that the single-layer MoS2 membrane with a full Mo pore (see Figure 7a) exhibited 88% of ion rejection. The work also demonstrated that the water flux was two to five orders of magnitude greater than other known nanoporous membranes. This technology is crucial to replacing the reverse osmosis (RO) membrane, especially in water desalination processes. Later, Kou et al. [194] discussed that a 0.74 nm nanopores in monolayer MoS2 membranes should be compatible with Debye screening length for the electrostatic interaction and lesser than the mean free path of molecules in water. Recent work from Kozubek et al. [195] proposed using irradiation with highly charged ions (HCIs) to create pores in MoS2. They found that pore creation efficiency has a linear relationship with potential energies for pore radius of 0.55–2.65 nm. Kozubek et al. [195] added that the HCI method has parallel writing capabilities, giving a high potential for mass production.
In the application of water desalination, Ma et al. [196] stated that the synthesis of MoS2/GO membrane resulted in an enhanced water flux (from 8.83 to 48.27 L·m−2∙h−1) with improved salts removal capacity (from 54.32% to 96.85%). The MoS2/GO membrane also exhibited different mechanism for dyes and ions (see Figure 7b) where dyes are physically sieved while ions have a Donnan effect due to existing trapped ions within interlayer spacing [196]. Gao et al. [197] demonstrated the performance of MoS2 membrane’s performance in rejecting 100% of methylene blue dye. The work highlighted that the increase in nanosheets to the membrane will reduce permeability and increase dye rejection rate.

4.3. Photocatalyst Technology in Wastewater Treatment

As the majority of water treatment facilities are located outdoors, the availability of solar energy can be useful for photocatalytic technology in water treatment. Chu et al. [198] reported that the development of hexagonal 2H-MoSe2 photocatalyst exhibits outstanding photo-absorption and photocatalytic reaction in reducing hexavalent chromium (Cr(VI)) under ultraviolet (UV) light. The 2H-MoSe2 yields 99% Cr(VI) reduction rate with the presence of UV light. Mittal and Khanuja [199] revealed that MoSe2 is a good photocatalyst not only for Cr(VI) but also for methylene blue (MB) and RhB. The MoSe2/strontium titanate (SrTiO3) heterostructure also showed great potential as a wastewater treatment photocatalyst with a degradation rate of methyl orange at 99.46% under the optimum loading weight of 0.1 wt.% under UV light [200]. It is also found that there is no significant loss in the performance of MoSe2/SrTiO3 after reuse for 6 times. The working principle of MoSe2/SrTiO3 photocatalyst is demonstrated in Figure 8.

5. Fuel Cleaning

Fuels are means of transfer for energy in our ecosystems. Most fuels have many imperfections related to conversion inefficiencies [201] and fuel impurity (including sulphur, nitrogen, aromatic, etc.) [202]. Direct usage of crude fuels with such impurity content will lead to an increase in gas emissions [203,204]. Most refineries for oil-based fuel carry out fuel cleaning processes such as desulphurization, dearomatization, denitrogenation, and deoxygenation [202,205,206] to remove such impurities before the fuel is being converted. The use of impurity removal technology generally depends on the techno-economics and fuel costs during application [207]). Fuels with a high impurity such as sulphur content can potentially cause an elevated particulate emission [208] while causing catalyst poisoning in subsequent systems [209].

5.1. Fuel Hydrodesulfurization

Hydrodesulfurization is the process of reduction of sulphur content from fuel such as diesel [202] by catalytically reacting them with hydrogen. Paul et al. [210] studied the mechanism of vacancy formation on the MoS2 catalyst in the hydrodesulfurization process, demonstrating the activation energy of 0.5 eV. TMDC materials such as Co/Ni promoted MoS2 were popular for the use of hydrodesulfurization due to its high activity, good selectivity, resistance to deactivation, and regeneration ability [211]. Commonly, TMDC-based catalysts are used for hydrodesulfurization in the oil and gas industry for high-efficiency sulphur removal from naphtha [212]. For the hydrodesulfurization of naphtha, a recent work from Mahmoudabadi et al. [213] achieved 100% conversion efficiency using MoS2 quantum dots nanocatalyst under the pressure of 15 bars, temperature of 280 °C and liquid hourly space velocity (LHSV) of 4 h−1. The direct use of unsupported Ni/MoS2 catalyst was reported [214] to remove furfurylamine (FA) and dibenzothiophene (DBT). Liu et al. [215] demonstrated that for a MoS2/NiMo catalyst, the higher the number of MoS2 slabs being stacked, the better the hydrodesulfurization selectivity for DBT. Additionally, Rangarajan et al. [216] studied the preferred active sites of a metal-promoted MoS2 and found that organosulfur and organonitrogen compounds bind weaker on sites with exposed metal on the corner and the sulphur edge of MoS2. Hydrodesulfurization of DBT by various morphologies of MoS2 catalyst was also studied by Tye and Smith [217]. This work showed that exfoliated MoS2 has the highest conversion, compared to the commercially crystalline MoS2 powder and MoS2 derived from soluble Mo precursors. While selectivity was correlated to the edge sites of MoS2. A 3D NiS-MoS2/Graphene nanohybrid was also shown to give a high DBT conversion rate of 82.6%, showing potential for 3D composites. Recent work from Abbasi et al. [218] demonstrated that a cobalt-promoted MoS2 achieved more than 80% conversion for the application of diesel, showing potential for the transfer of high-performance TMDC catalysts from oil and gas to diesel/biodiesel industries.
Besides hydrodesulfurization to remove DBT, the removal of thiophene was also demonstrated to be feasible with MoS2 catalysis by DFT methods [219]. Kaluza et al. [220] synthesized a MoS2 catalyst supported on mesoporous alumina that gave a thiophene conversion of more than 60% at around 400 °C. For the removal of carbonyl sulphide (COS), Liu et al. [221] used a microwave activated MoS2/graphene catalyst that achieved over 90% conversion for temperatures over 280 °C. Later, the use of a monolayer MoS2 anchored on reduced graphene oxides [222] has shown to give over 90% conversion and a reduced temperature at about 180 °C.

5.2. Fuel Hydrodeoxygenation

Hydrodeoxygenation is the removal of oxygenated components from the fuel by reacting it with hydrogen to form water [206]. Liu et al. [223] found that a MoS2 monolayer doped with an isolated Co atom can effectively reduce hydrodeoxygenation of 4-methylphenol to toluene from 300 °C to 180 °C while maintaining high conversion and selectivity of 97.6% and 98.4% respectively. From a DFT perspective, Li et al. [224] demonstrate a generic structure of Co atom vacancy in MoS2, claiming that the catalyst can activate hydrodeoxygenation reactions at lowered temperatures. Recent work from Wu et al. [225] adsorbed Co oxide at the edge of MoS2 with sulphur defects and formed a Co-MoS2−x catalyst that converts lignin-derived phenols (4-methylphenol) to arenes with 97.4% conversion and 99.6% toluene selectivity at a mild temperature of 120 °C (see Figure 9a). For direct industrial application, palm kernel oil was converted with a yield over 90% to jet fuel-like hydrocarbon (see Figure 9b) via a Ni-MoS2/γ-Al2O3 hydrodeoxygenation catalyst [206]. The work showed high potential for converting biomass-based fuel to high-value fuels using TMDC-based catalysts. Moreover, Co- and Ni- promoted MoS2 catalyst was also used to co-process diesel and vegetable oil via hydrodeoxygenation reaction [226]. Alvarez-Galvan et al. [227] demonstrated that transition metals phosphides were useful as a hydrodeoxygenation catalyst for waste cooking oil to green biodiesel. More research work is required to explore the possibility of TMDC materials in this field.
For catalytic deoxygenation of alkali lignin into bio-oil, Li et al. [228] proposed the use of a flower-like hierarchical MoS2-based composite catalyst that achieved a lignin conversion of 91.26% and bio-oil yield of 86.24%. Later, work from Zhou et al. [229] synthesized a series of MoSe2 catalysts for the conversion of alkali lignin into bio-oil. The work resulted in 96.46% conversion of lignin and 93.68% yield of bio-oil. A recent review by Porsin et al. [230] revealed that sulphide catalysts such as MoS2 can also convert fatty acid triglycerides to motor oil by hydrodeoxygenation. As for wider fuel cleaning applications, TMDC materials have also been reported to purify gas fuels such as hydrogen and methane [231] showing high potential in diverse fuel applications in the future.

6. CO2 Valorization and Conversion

Valorization of carbon dioxide (CO2) to raw chemical materials and clean fuels is an opportunity for the artificial carbon cycle, which contributes to the mitigation of global warming and alleviates the usage of fossil [232]. Over the decades, enormous efforts in searching of alternative technologies to mitigate the CO2 emissions through CO2 capture from concentrated industrial exhausts [233,234], sequestration of CO2 in the underground [235,236] and conversion of CO2 to energy-rich fuels powered by renewable energy resources [237] have been discovered. Throughout all these methods, CO2 molecules are not solely can be removed from the atmosphere but also can be converted into value-added chemicals such as methanol, formic acid, methane, and syngas [238,239,240]. Recently, electrochemical conversion of CO2 to value-added chemicals through the principal of CO2 electrochemical reduction reaction (CO2ERR) as shown in Figure 10a [241], has emerged as a comparative alternative to its counterparts such as biochemical and thermochemical technologies [242,243,244]. Alternatively, semiconductor-based photocatalysis of CO2 reduction has been another frontier in CO2 conversion [245]. When a light source with appropriate photoenergy is illuminated on this photocatalyst, electron-hole pairs are formed [246,247]. With the migration of the electron-hole pair to the surface, the release of energy from within can reduce surface adsorbed CO2 [246]. Another interesting type of photocatalyst is the Z-scheme photocatalyst, which performs charge transfer between a reduction semiconductor catalyst and an oxidation semiconductor catalyst, potentially giving higher performances, efficiency, and synthesis possibilities [248]. The general working principle of photocatalyst for CO2 conversion is illustrated in Figure 10b.

6.1. Conversion of CO2 to Syngas and Other Gases

To achieve this attractive blueprint, the key issue is to develop electrocatalysts with high activity, superior selectivity, highly durable, environmentally friendly as well as low cost [250,251,252]. In 2014, the first pioneering work of elucidating the TMDCs materials as advanced electrocatalysts for CO2 reduction was by Asadi and coauthors [253]. They have claimed that MoS2 material serves as a highly efficient electrocatalyst for the conversion of CO2 to syngas. The combination of the edge states of MoS2 in contact with ionic liquid solvent electrolytes has served as a new paradigm for CO2 reduction, providing the advantage of favorable electronic properties of MoS2 and an electrolyte that transfer the CO2 molecules to the active sites for reaction. This study is a breakthrough in this field, in which the performance of the catalytic activity for CO2 reduction reaction is far exceeding than other conventional catalysts such as carbon nanotubes, graphene, noble metal carbides, and transition metals.
Recent research work of TMDCs in converting CO2 to interesting gas products using photo- and electro-catalysis methods are shown in Table 4. Up to date, most reports of TMDCs in this field are focusing on the conversion of CO2 to CO, suggesting that there is still a huge potential of different electrochemical reaction pathways using TDMCs that can be explored in the near future such as the conversion of CO2 to CO and H2 to CH4. Wang et al. [254] synthesized a marigold-like SiC@MoS2 nanoflower for the conversion of CO2 and H2O to CH4 and O2 using no sacrificial agents while operating within the visible light spectrum (see Figure 11a). The work reported production rates of 323 and 621 µL∙g–1∙h–1 for CH4 and O2 while maintaining stable characteristics for 40 h. Asadi et al. [255] also have synthesized a series of TMDCs (e.g. WSe2, WS2, MoSe2, and MoS2) using the chemical vapor transport (CVT) growth technique followed by liquid exfoliation for electrochemical CO2 conversion using ionic liquid electrolyte (EMIM-BF4). By benchmarking with the bulk Ag and Ag NPs, the current density of TMDCs are more than tenfold (130–330 mA cm−2) of the current density of Ag (3.3 mA cm−2) and Ag NPs (10 mA cm−2) with 90% CO. Overall, WSe2 NFs gave the best performance and exhibited a current density of 18.95 mA cm−2 with a high turn-over frequency of CO of 0.28 s−1 at an overpotential of 54 mV. This can be related to the intrinsic properties of WSe2 (very low work function and high-volume surface area).
In addition, a lab-scale synthetic custom-builds wireless setup for photochemical studies of WSe2/IL is also being developed by this group [255]. The build-up of the artificial leaf that performed the photosynthesis process is investigated to further confirm the CO2 conversion rate efficiency (see Figure 11b). The system comprises of three main components: (i) harvest light using two amorphous silicon triple-junction photovoltaic cells in series, (ii) CO2 reduction using WSe2/IL on the cathode and lastly, and (iii) evolution of oxygen using cobalt (CoII) oxide/hydroxide in potassium phosphate electrolyte [265,266]. Surprisingly, the cell managed to function continuously for more than 4 h before corrosion happens at the transparent indium tin oxide layer on the anode, proving that this theoretical experiment is workable and insightful, which could be extended to the technology development scale in near future. Further studies on the defect engineering for CO2 reduction catalysts were also discussed by Wang et al. [267] to improve functionality by defect engineering (such as holes, doping, vacancies, phase change, edge, grain boundary, lattices distortion, and other methods). For this, MoS2 with substitutional defects on the edge was studied [253] and showed to give better current density than Au, Ag, Cu nanopowder, and the polycrystalline catalyst. Ji et al. [268] suggested that group 10 TMDCs have a high potential for CO2 reduction as they do not suffer from OH poisoning. The work proposed that the reaction energies for CO2 reduction could be tuned by vacancy densities.

6.2. Direct Conversion of CO2 to High-Value Chemicals

Direct conversion of CO2 to high-value chemicals (such as intermediate chemicals, fuels, and alcohols) has a high potential for commercial value [269,270]. TMDCs have demonstrated properties for catalyzing CO2 into important alcohol, especially methanol [85,271]. Bolivar et al. [272] studied the use of a Pt/MoOx/MoSe2 electrocatalyst and proposed that oxygen-containing surface species lowers the electric potential for methanol production. Moreover, Tu et al. [273] have demonstrated the methanol production of TiO2 and 0.5 wt.% MoS2 nanocomposites was higher than Pt/TiO2, Au/TiO2, and Ag/TiO2 for the photocatalytic reduction of CO2. For this TiO2/MoS2 composites, Xu et al. [274] showed that electrons are transferred from MoS2 to TiO2 upon contact, therefore promoting the separation of charge carriers upon photoexcitation. In 2018, Francis et al. [275] have reported on the activity of single crystals and thin films of MoS2 for the reduction of CO2 dissolved in an aqueous electrolyte to yield 1-propanol, ethylene glycol, and t-butanol and hydrogen. At an applied potential of −0.59 V, the Faradaic efficiencies for the reduction of CO2 to 1-propanol are ca. 3.5% for MoS2 single crystals and 1% for thin films with low edge-site densities. WS2 quantum dots doped Bi2S3 nanotubes also demonstrated high (38.2 μmol∙g−1∙h−1) production of methanol while co-producing ethanol [85]. Many researchers have also performed the conversion of CO2 to high-value products such as methanol, acetaldehyde [249] and ethanol [276] by using TMDC catalysts (see Table 5).

7. Challenges and Future Prospects

The application of TMDCs in pollution reduction is a breakthrough as has high availability [30,31,32], high economic potential [280,281,282], and excellent efficiency [254,276] for the application of pollution reduction. For TMDCs availability and efficiency, we have already discussed them in Section 1 and Section 2, Section 3, Section 4, Section 5 and Section 6 (depending on application) respectively. In terms of commercial techno-economic analysis, a few works have also pointed out the advantage of using TMDC catalysts. Valle et al. [280] studied the economics of converting biomass-derived syngas to ethanol using an indirect circulating fluidized bed gasification (iCFBG) system via the KCoMoS2 catalyst. The work highlighted that the iCFBG system could reach a minimum ethanol selling price (MESP) of 0.74 $/L (USD) while technologies such as entrained-flow gasification have that at 0.98$/L. This MoS2-based catalyst was also found to be more commercially viable than the Rh-Mn/SiO2 catalyst due to their cheaper economical costs, although the latter has better energy efficiency [281]. Phillips [282] also concluded that the use of Co/MoS2 has potential to lower down ethanol MESP down to 0.267 $/L (USD) highlighting that Ru or Rh-based catalysts are economically expensive even at low concentrations. Nevertheless, authors acknowledge that the techno-economic analysis of cutting-edge TMDCs technologies is still generally unknown. Thus, more TMDCs research studies should include consideration of economic analysis. Furthermore, despite the unique physiochemical properties offered by the TMDCs (e.g., the large surface area, which acts as a two-edged sword [283], tunable absorption heterostructure [200,284], flexible, diversiform, and functional properties [146]), more exploration on the operating conditions, kinetic, and thermodynamic parameters should be carried out to further confirm its ability [285]. This is because most of the reviewed works are conducted under a controlled environment with a specially designed flow. Under this context, whether obtained simulation and experimental results reflect the actual situation remains uncertain. The main limitations of commercializing of TMDCs-based technology are:
(i)
Most research works are focusing on the development of the materials as an effective pollutant degradation catalyst during oxidation, neglecting that in reality, the number of organic pollutants is in a mixture that may affect the performance of the engineered material [286,287].
(ii)
The rate of product formation for TMDC-based technology, such as syngas and methanol from the conversion of CO2 is still far from the industrial scale, which is not sufficient to accommodate the global demand [286,287]. More research effort is required to improve efficiencies.
(iii)
High preparation/fabrication cost and poor uniformity of the materials in large-scale production [288]. More effort is required in sustainable synthesis and fabrication of TMDC-based technologies.
(iv)
The larger surface area offered by the TMDCs can enhance the adsorption capability, but also cause higher effects of environmental disturbance [283]. Precise defect engineering to improve material performances can also be carried out [225,267]. Optimization and data analytics on the design dilemma and defect engineering should be highlighted and properly studied.
As a future prospect, commercial validation on the performance on a larger scale needs to be carried out. The results obtained from the experiment can be validated by conducting techno-economic analysis, such as Monte Carlo analysis for economic and risk feasibility [289,290]. Furthermore, advanced machine learning methods can also be used to accelerate experimental designs [291,292], obtain optimal performances [293,294], and for the discovery of new TMDCs composite structures [295,296].

8. Conclusions

The unique and exclusive features of TMDCs (e.g., layered structure, tunable bandgap, unique optical, thermal and electrical properties, etc.) have been the main driving force that drives the researchers’ attention on exploring the potential of the 2D materials in pollution reduction applications. This work summarized the state-of-the-art applications of various TMDCs under the context of pollution mitigation (including (i) gas adsorption and removal, (ii) gas sensing, (iii) wastewater treatment, (iv) flue cleaning, and (v) CO2 valorization and conversion). In addition to the up-to-date progress of TMDCs research, this article also discussed some of the key challenges for the future commercialization of TMDC materials. Apparently, many of the reviewed research works have authenticated their substantial potential to substitute the existing pollution mitigation media. Nevertheless, the current applications are still restricted to a lab basis, where the deviation of the actual performance of TMDCs under larger-scale production remains as the research gap. To usher TMDCs into the next level of utilization (i.e., from the lab scale to the industrial scale), the following three research directions should be followed up, (i) techno-economic analysis (TEA) study, (ii) experimentation under more rigorous and realistic conditions, and (iii) experimental optimization for application purpose. The authors sincerely hope this review can serve as an insightful guideline that inspires more researchers to venture into this new and exciting cutting-edge field.

Author Contributions

Conceptualization, X.Z.; writing—original draft preparation, X.Z., S.Y.T., A.C.M.L., B.S.H., W.D.L.; writing—X.Z., S.Y.T., A.C.M.L., B.S.H., W.D.L., X.T.; visualization, S.Y.T., A.C.M.L., B.S.H.; supervision, X.Z., X.T.; project administration, X.Z.; funding acquisition, X.Z., X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grant Nos. 61975098, and 51972194) and IMPROVE II CZ.02.2.69/0.0/0.0/18_070/0009469 under OP VVV program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011. [Google Scholar] [CrossRef]
  2. Yin, Z.; Li, H.; Li, H.; Jiang, L.; Shi, Y.; Sun, Y.; Lu, G.; Zhang, Q.; Chen, X.; Zhang, H. Single-layer MoS2 phototransistors. ACS Nano 2012. [Google Scholar] [CrossRef] [Green Version]
  3. Mak, K.F.; McGill, K.L.; Park, J.; McEuen, P.L. The valley hall effect in MoS2 transistors. Science 2014. [Google Scholar] [CrossRef] [Green Version]
  4. Li, H.; Jia, X.; Zhang, Q.; Wang, X. Metallic transition-metal dichalcogenide nanocatalysts for energy conversion. Chem 2018, 4, 1510–1537. [Google Scholar] [CrossRef] [Green Version]
  5. Gao, Y.-P.; Wu, X.; Huang, K.-J.; Xing, L.-L.; Zhang, Y.-Y.; Liu, L. Two-dimensional transition metal diseleniums for energy storage application: A review of recent developments. CrystEngComm 2017, 19, 404–418. [Google Scholar] [CrossRef]
  6. Lu, J.M.; Zheliuk, O.; Leermakers, I.; Yuan, N.F.Q.; Zeitler, U.; Law, K.T.; Ye, J.T. Evidence for two-dimensional Ising superconductivity in gated MoS2. Science 2015. [Google Scholar] [CrossRef] [Green Version]
  7. Liu, X.; Zhang, Y.W. Thermal properties of transition-metal dichalcogenide. Chin. Phys. B 2018, 27, 034402. [Google Scholar] [CrossRef]
  8. Kappera, R.; Voiry, D.; Yalcin, S.E.; Branch, B.; Gupta, G.; Mohite, A.D.; Chhowalla, M. Phase-engineered low-resistance contacts for ultrathin MoS2 transistors. Nat. Mater. 2014. [Google Scholar] [CrossRef]
  9. Kolobov, A.V.; Tominaga, J. Two-Dimensional Transition-Metal Dichalcogenides; Springer International Publishing: Cham, Switzerland, 2016; ISBN 978-3-319-31449-5. [Google Scholar]
  10. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef]
  11. Santosh, K.C.; Longo, R.C.; Addou, R.; Wallace, R.M.; Cho, K. Impact of intrinsic atomic defects on the electronic structure of MoS2 monolayers. Nanotechnology 2014, 25, 375703. [Google Scholar] [CrossRef]
  12. Feng, S.; Lin, Z.; Gan, X.; Lv, R.; Terrones, M. Doping two-dimensional materials: Ultra-sensitive sensors, band gap tuning and ferromagnetic monolayers. Nanoscale Horizons 2017. [Google Scholar] [CrossRef]
  13. Ghorbani-Asl, M.; Borini, S.; Kuc, A.; Heine, T. Strain-dependent modulation of conductivity in single-layer transition-metal dichalcogenides. Phys. Rev. B Condens. Matter Mater. Phys. 2013. [Google Scholar] [CrossRef] [Green Version]
  14. Su, X.; Ju, W.; Zhang, R.; Guo, C.; Zheng, J.; Yong, Y.; Li, X. Bandgap engineering of MoS2/MX2 (MX2 = WS2, MoSe2 and WSe2) heterobilayers subjected to biaxial strain and normal compressive strain. RSC Adv. 2016. [Google Scholar] [CrossRef]
  15. Gusakova, J.; Wang, X.; Shiau, L.L.; Krivosheeva, A.; Shaposhnikov, V.; Borisenko, V.; Gusakov, V.; Tay, B.K. Electronic properties of bulk and monolayer TMDs: Theoretical study within DFT framework (GVJ-2e method). Phys. Status Solidi Appl. Mater. Sci. 2017. [Google Scholar] [CrossRef]
  16. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010. [Google Scholar] [CrossRef] [Green Version]
  17. Acerce, M.; Voiry, D.; Chhowalla, M. Metallic 1T phase MoS2 nanosheets as supercapacitor electrode materials. Nat. Nanotechnol. 2015. [Google Scholar] [CrossRef]
  18. Li, H.; Tan, Y.; Liu, P.; Guo, C.; Luo, M.; Han, J.; Lin, T.; Huang, F.; Chen, M. Atomic-sized pores enhanced electrocatalysis of TaS2 nanosheets for hydrogen evolution. Adv. Mater. 2016. [Google Scholar] [CrossRef]
  19. Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. Lithium ion battery applications of molybdenum disulfide (MoS2) nanocomposites. Energy Environ. Sci. 2014. [Google Scholar] [CrossRef]
  20. Chang, Y.H.; Lin, C.T.; Chen, T.Y.; Hsu, C.L.; Lee, Y.H.; Zhang, W.; Wei, K.H.; Li, L.J. Highly efficient electrocatalytic hydrogen production by MoSx grown on graphene-protected 3D Ni foams. Adv. Mater. 2013. [Google Scholar] [CrossRef]
  21. Kannan, P.K.; Late, D.J.; Morgan, H.; Rout, C.S. Recent developments in 2D layered inorganic nanomaterials for sensing. Nanoscale 2015. [Google Scholar] [CrossRef]
  22. Sun, Y.; Gao, S.; Lei, F.; Xie, Y. Atomically-thin two-dimensional sheets for understanding active sites in catalysis. Chem. Soc. Rev. 2015, 44, 623–636. [Google Scholar] [CrossRef]
  23. Sang, Y.; Zhao, Z.; Zhao, M.; Hao, P.; Leng, Y.; Liu, H. From UV to near-infrared, WS2 nanosheet: A novel photocatalyst for full solar light spectrum photodegradation. Adv. Mater. 2015. [Google Scholar] [CrossRef]
  24. Atkin, P.; Daeneke, T.; Wang, Y.; Carey, B.J.; Berean, K.J.; Clark, R.M.; Ou, J.Z.; Trinchi, A.; Cole, I.S.; Kalantar-Zadeh, K. 2D WS2/carbon dot hybrids with enhanced photocatalytic activity. J. Mater. Chem. A 2016. [Google Scholar] [CrossRef]
  25. Di, J.; Xia, J.; Ge, Y.; Xu, L.; Xu, H.; Chen, J.; He, M.; Li, H. Facile fabrication and enhanced visible light photocatalytic activity of few-layer MoS2 coupled BiOBr microspheres. Dalt. Trans. 2014. [Google Scholar] [CrossRef]
  26. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 2007. [Google Scholar] [CrossRef] [Green Version]
  27. Mohanty, B.; Ghorbani-Asl, M.; Kretschmer, S.; Ghosh, A.; Guha, P.; Panda, S.K.; Jena, B.; Krasheninnikov, A.V.; Jena, B.K. MoS2 quantum dots as efficient catalyst materials for the oxygen evolution reaction. ACS Catal. 2018. [Google Scholar] [CrossRef]
  28. Ye, G.; Gong, Y.; Lin, J.; Li, B.; He, Y.; Pantelides, S.T.; Zhou, W.; Vajtai, R.; Ajayan, P.M. Defects engineered monolayer MoS2 for improved hydrogen evolution reaction. Nano Lett. 2016. [Google Scholar] [CrossRef]
  29. Chan, K.; Tsai, C.; Hansen, H.A.; Nørskov, J.K. Molybdenum sulfides and selenides as possible electrocatalysts for CO2 reduction. ChemCatChem 2014. [Google Scholar] [CrossRef]
  30. Mohajerin, T.J.; Helz, G.R.; Johannesson, K.H. Tungsten-molybdenum fractionation in estuarine environments. Geochim. Cosmochim. Acta 2016. [Google Scholar] [CrossRef] [Green Version]
  31. Ren, Z.; Zhou, T.; Hollings, P.; White, N.C.; Wang, F.; Yuan, F. Trace element geochemistry of molybdenite from the Shapinggou super-large porphyry Mo deposit, China. Ore Geol. Rev. 2018. [Google Scholar] [CrossRef]
  32. Golden, J.; McMillan, M.; Downs, R.T.; Hystad, G.; Goldstein, I.; Stein, H.J.; Zimmerman, A.; Sverjensky, D.A.; Armstrong, J.T.; Hazen, R.M. Rhenium variations in molybdenite (MoS2): Evidence for progressive subsurface oxidation. Earth Planet. Sci. Lett. 2013. [Google Scholar] [CrossRef]
  33. Pumera, M.; Sofer, Z.; Ambrosi, A. Layered transition metal dichalcogenides for electrochemical energy generation and storage. J. Mater. Chem. A 2014. [Google Scholar] [CrossRef]
  34. Eftekhari, A. Tungsten dichalcogenides (WS2, WSe2, and WTe2): Materials chemistry and applications. J. Mater. Chem. A 2017, 5, 18299–18325. [Google Scholar] [CrossRef]
  35. Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P.M.; Lou, J. Large-area vapor-phase growth and characterization of MoS2 atomic layers on a SiO2 substrate. Small 2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ji, Q.; Li, C.; Wang, J.; Niu, J.; Gong, Y.; Zhang, Z.; Fang, Q.; Zhang, Y.; Shi, J.; Liao, L.; et al. Metallic vanadium disulfide nanosheets as a platform material for multifunctional electrode applications. Nano Lett. 2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wilson, J.A.; Yoffe, A.D. The transition metal dichalcogenides discussion and interpretation of the observed optical, electrical and structural properties. Adv. Phys. 1969. [Google Scholar] [CrossRef]
  38. Ubaldini, A.; Jacimovic, J.; Ubrig, N.; Giannini, E. Chloride-driven chemical vapor transport method for crystal growth of transition metal dichalcogenides. Cryst. Growth Des. 2013. [Google Scholar] [CrossRef]
  39. Zhang, X.; Lou, F.; Li, C.; Zhang, X.; Jia, N.; Yu, T.; He, J.; Zhang, B.; Xia, H.; Wang, S.; et al. Flux method growth of bulk MoS2 single crystals and their application as a saturable absorber. CrystEngComm 2015, 17, 4026–4032. [Google Scholar] [CrossRef]
  40. Ye, L.; Xu, H.; Zhang, D.; Chen, S. Synthesis of bilayer MoS2 nanosheets by a facile hydrothermal method and their methyl orange adsorption capacity. Mater. Res. Bull. 2014. [Google Scholar] [CrossRef]
  41. Kalosi, A.; Demydenko, M.; Bodik, M.; Hagara, J.; Kotlar, M.; Kostiuk, D.; Halahovets, Y.; Vegso, K.; Marin Roldan, A.; Maurya, G.S.; et al. Tailored langmuir–schaefer deposition of few-layer MoS2 nanosheet films for electronic applications. Langmuir 2019, 35, 9802–9808. [Google Scholar] [CrossRef]
  42. Huang, X.; Zeng, Z.; Zhang, H. Metal dichalcogenide nanosheets: Preparation, properties and applications. Chem. Soc. Rev. 2013, 42, 1934–1946. [Google Scholar] [CrossRef] [PubMed]
  43. Li, H.; Wu, J.; Yin, Z.; Zhang, H. Preparation and applications of mechanically exfoliated single-layer and multilayer MoS2 and WSe2 nanosheets. Acc. Chem. Res. 2014, 47, 1067–1075. [Google Scholar] [CrossRef] [PubMed]
  44. Cunningham, G.; Lotya, M.; Cucinotta, C.S.; Sanvito, S.; Bergin, S.D.; Menzel, R.; Shaffer, M.S.P.; Coleman, J.N. Solvent exfoliation of transition metal dichalcogenides: Dispersibility of exfoliated nanosheets varies only weakly between compounds. ACS Nano 2012. [Google Scholar] [CrossRef] [PubMed]
  45. Bissett, M.A.; Worrall, S.D.; Kinloch, I.A.; Dryfe, R.A.W. Comparison of two-dimensional transition metal dichalcogenides for electrochemical supercapacitors. Electrochim. Acta 2016. [Google Scholar] [CrossRef]
  46. Yue, Q.; Shao, Z.; Chang, S.; Li, J. Adsorption of gas molecules on monolayer MoS2 and effect of applied electric field. Nanoscale Res. Lett. 2013. [Google Scholar] [CrossRef] [Green Version]
  47. Kumar, R.; Goel, N.; Hojamberdiev, M.; Kumar, M. Transition metal dichalcogenides-based flexible gas sensors. Sensors Actuators A Phys. 2020, 303, 111875. [Google Scholar] [CrossRef]
  48. Elsevier. Welcome to Scopus Preview. Available online: https://www.scopus.com/home.uri (accessed on 4 May 2020).
  49. Blumenthal, I. Carbon monoxide poisoning. J. R. Soc. Med. 2001, 94, 270–272. [Google Scholar] [CrossRef]
  50. Li, Z.; He, J.; Wang, H.; Wang, B.; Ma, X. Enhanced methanation stability of nano-sized MoS2 catalysts by adding Al2O3. Front. Chem. Sci. Eng. 2015. [Google Scholar] [CrossRef]
  51. Almeida, K.; Peña, P.; Rawal, T.B.; Coley, W.C.; Akhavi, A.A.; Wurch, M.; Yamaguchi, K.; Le, D.; Rahman, T.S.; Bartels, L. A single layer of MoS2 activates gold for room temperature CO oxidation on an inert silica substrate. J. Phys. Chem. C 2019. [Google Scholar] [CrossRef]
  52. Lolla, S.; Luo, X. Tuning the catalytic properties of monolayer MoS2 through doping and sulfur vacancies. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef] [Green Version]
  53. Ma, D.; Tang, Y.; Yang, G.; Zeng, J.; He, C.; Lu, Z. CO catalytic oxidation on iron-embedded monolayer MoS2. Appl. Surf. Sci. 2015. [Google Scholar] [CrossRef]
  54. Lu, X.; Guo, L.; Wang, P.; Cui, M.; Kanghong, D.; Peng, W. Theoretical investigation of the adsorption of gas molecules on WSe2 monolayers decorated with Pt, Au nanoclusters. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  55. Cui, H.; Zhang, G.; Zhang, X.; Tang, J. Rh-doped MoSe2 as a toxic gas scavenger: A first-principles study. Nanoscale Adv. 2019. [Google Scholar] [CrossRef] [Green Version]
  56. Chen, W.; Zijlstra, B.; Filot, I.A.W.; Pestman, R.; Hensen, E.J.M. Mechanism of carbon monoxide dissociation on a cobalt Fischer–Tropsch catalyst. ChemCatChem 2018. [Google Scholar] [CrossRef] [Green Version]
  57. American Chemical Society (ACS). It’s Water Vapor, Not the CO2. Available online: https://www.acs.org/content/acs/en/climatescience/climatesciencenarratives/its-water-vapor-not-the-co2.html (accessed on 2 May 2020).
  58. 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. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, H.; Kong, D.; Johanes, P.; Cha, J.J.; Zheng, G.; Yan, K.; Liu, N.; Cui, Y. MoSe2 and WSe2 nanofilms with vertically aligned molecular layers on curved and rough surfaces. Nano Lett. 2013. [Google Scholar] [CrossRef]
  60. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jørgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Nørskov, J.K. Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005. [Google Scholar] [CrossRef] [PubMed]
  61. Yu, K.; Groom, D.J.; Wang, X.; Yang, Z.; Gummalla, M.; Ball, S.C.; Myers, D.J.; Ferreira, P.J. Degradation mechanisms of platinum nanoparticle catalysts in proton exchange membrane fuel cells: The role of particle size. Chem. Mater. 2014. [Google Scholar] [CrossRef]
  62. Zhao, Y.; Hwang, J.; Tang, M.T.; Chun, H.; Wang, X.; Zhao, H.; Chan, K.; Han, B.; Gao, P.; Li, H. Ultrastable molybdenum disulfide-based electrocatalyst for hydrogen evolution in acidic media. J. Power Sources 2020. [Google Scholar] [CrossRef]
  63. Rong, H.; Zhang, H.; Xiao, S.; Li, C.; Hu, C. Optimizing energy consumption for data centers. Renew. Sustain. Energy Rev. 2016, 58, 674–691. [Google Scholar] [CrossRef]
  64. Li, B.B.; Qiao, S.Z.; Zheng, X.R.; Yang, X.J.; Cui, Z.D.; Zhu, S.L.; Li, Z.Y.; Liang, Y.Q. Pd coated MoS2 nanoflowers for highly efficient hydrogen evolution reaction under irradiation. J. Power Sources 2015. [Google Scholar] [CrossRef]
  65. Jian, J.; Li, Y.; Bi, H.; Wang, X.; Wu, X.; Qin, W. Aluminum decoration on MoS2 ultrathin nanosheets for highly efficient hydrogen evolution. ACS Sustain. Chem. Eng. 2020. [Google Scholar] [CrossRef]
  66. Bar-Ziv, R.; Ranjan, P.; Lavie, A.; Jain, A.; Garai, S.; Bar Hen, A.; Popovitz-Biro, R.; Tenne, R.; Arenal, R.; Ramasubramaniam, A.; et al. Au-MoS2 hybrids as hydrogen evolution electrocatalysts. ACS Appl. Energy Mater. 2019. [Google Scholar] [CrossRef] [Green Version]
  67. Chia, X.; Sutrisnoh, N.A.A.; Pumera, M. Tunable Pt-MoSx hybrid catalysts for hydrogen evolution. ACS Appl. Mater. Interfaces 2018. [Google Scholar] [CrossRef] [PubMed]
  68. Rastogi, P.K.; Sarkar, S.; Mandler, D. Ionic strength induced electrodeposition of two-dimensional layered MoS2 nanosheets. Appl. Mater. Today 2017. [Google Scholar] [CrossRef]
  69. Cao, J.; Zhou, J.; Zhang, Y.; Liu, X. A Clean and facile synthesis strategy of MoS2 nanosheets grown on multi-wall CNTs for enhanced hydrogen evolution reaction performance. Sci. Rep. 2017. [Google Scholar] [CrossRef] [Green Version]
  70. Chua, X.J.; Pumera, M. The effect of varying solvents for MoS2 treatment on its catalytic efficiencies for HER and ORR. Phys. Chem. Chem. Phys. 2017. [Google Scholar] [CrossRef]
  71. Ren, J.; Zong, H.; Sun, Y.; Gong, S.; Feng, Y.; Wang, Z.; Hu, L.; Yu, K.; Yu, K. 2D organ-like molybdenum carbide (MXene) coupled with MoS2 nanoflowers enhances the catalytic activity in the hydrogen evolution reaction. CrystEngComm 2020. [Google Scholar] [CrossRef]
  72. Wang, Y.; Jian, C.; He, X.; Liu, W. Self-supported molybdenum selenide nanosheets grown on urchin-like cobalt selenide nanowires array for efficient hydrogen evolution. Int. J. Hydrogen Energy 2020. [Google Scholar] [CrossRef]
  73. Seok, J.; Lee, J.H.; Cho, S.; Ji, B.; Kim, H.W.; Kwon, M.; Kim, D.; Kim, Y.M.; Oh, S.H.; Kim, S.W.; et al. Active hydrogen evolution through lattice distortion in metallic MoTe2. 2D Mater. 2017. [Google Scholar] [CrossRef] [Green Version]
  74. Pan, Y.; Zheng, F.; Wang, X.; Qin, H.; Liu, E.; Sha, J.; Zhao, N.; Zhang, P.; Ma, L. Enhanced electrochemical hydrogen evolution performance of WS2 nanosheets by Te doping. J. Catal. 2020. [Google Scholar] [CrossRef]
  75. Kadam, S.R.; Enyashin, A.N.; Houben, L.; Bar-Ziv, R.; Bar-Sadan, M. Ni-WSe2 nanostructures as efficient catalysts for electrochemical hydrogen evolution reaction (HER) in acidic and alkaline media. J. Mater. Chem. A 2020. [Google Scholar] [CrossRef]
  76. Rahman, F.A.; Aziz, M.M.A.; Saidur, R.; Bakar, W.A.W.A.; Hainin, M.; Putrajaya, R.; Hassan, N.A. Pollution to solution: Capture and sequestration of carbon dioxide (CO2) and its utilization as a renewable energy source for a sustainable future. Renew. Sustain. Energy Rev. 2017, 71, 112–126. [Google Scholar] [CrossRef]
  77. Shen, Y.; Wang, H.; Zhang, X.; Zhang, Y. MoS2 nanosheets functionalized composite mixed matrix membrane for enhanced CO2 capture via surface drop-coating method. ACS Appl. Mater. Interfaces 2016. [Google Scholar] [CrossRef] [PubMed]
  78. Sun, Q.; Qin, G.; Ma, Y.; Wang, W.; Li, P.; Du, A.; Li, Z. Electric field controlled CO2 capture and CO2/N2 separation on MoS2 monolayers. Nanoscale 2017. [Google Scholar] [CrossRef]
  79. Aguilar, N.; Aparicio, S. Theoretical insights into CO2 adsorption by MoS2 nanomaterials. J. Phys. Chem. C 2019. [Google Scholar] [CrossRef]
  80. Shi, G.; Yu, L.; Ba, X.; Zhang, X.; Zhou, J.; Yu, Y. Copper nanoparticle interspersed MoS2 nanoflowers with enhanced efficiency for CO2 electrochemical reduction to fuel. Dalt. Trans. 2017. [Google Scholar] [CrossRef]
  81. Kim, R.; Kim, J.; Do, J.Y.; Seo, M.W.; Kang, M. Carbon dioxide photoreduction on the Bi2S3/MoS2 catalyst. Catalysts 2019, 998. [Google Scholar] [CrossRef] [Green Version]
  82. Hu, X.; Yang, H.; Gao, M.; Tian, H.; Li, Y.; Liang, Z.; Jian, X. Insights into the photoassisted electrocatalytic reduction of CO2 over a two-dimensional MoS2 nanostructure loaded on SnO2 nanoparticles. ChemElectroChem 2019. [Google Scholar] [CrossRef]
  83. Yu, L.; Xie, Y.; Zhou, J.; Li, Y.; Yu, Y.; Ren, Z. Robust and selective electrochemical reduction of CO2: The case of integrated 3D TiO2@MoS2 architectures and Ti-S bonding effects. J. Mater. Chem. A 2018. [Google Scholar] [CrossRef]
  84. Ai, W.; Kou, L.; Hu, X.; Wang, Y.; Krasheninnikov, A.V.; Sun, L.; Shen, X. Enhanced sensitivity of MoSe2 monolayer for gas adsorption induced by electric field. J. Phys. Condens. Matter 2019. [Google Scholar] [CrossRef] [PubMed]
  85. Rathi, K.; Pal, K. Ruthenium-decorated tungsten disulfide quantum dots for a CO2 gas sensor. Nanotechnology 2020. [Google Scholar] [CrossRef] [PubMed]
  86. Myllyvirta, L. Global Air Pollution Map: Ranking the World’s Worst SO2 and NO2 Emission Hotspots. Available online: https://storage.googleapis.com/planet4-africa-stateless/2019/03/625c2655-ranking-so2-and-no2-hotspots_19-march-2019.pdf (accessed on 2 May 2020).
  87. Aas, W.; Mortier, A.; Bowersox, V.; Cherian, R.; Faluvegi, G.; Fagerli, H.; Hand, J.; Klimont, Z.; Galy-Lacaux, C.; Lehmann, C.M.B.; et al. Global and regional trends of atmospheric sulfur. Sci. Rep. 2019. [Google Scholar] [CrossRef] [PubMed]
  88. Ye, X.; Ma, S.; Jiang, X.; Yang, Z.; Jiang, W.; Wang, H. The adsorption of acidic gaseous pollutants on metal and nonmetallic surface studied by first-principles calculation: A review. Chinese Chem. Lett. 2019. [Google Scholar] [CrossRef]
  89. Gui, Y.; Chen, J.; Wang, W.; Zhu, Y.; Tang, C.; Xu, L. Adsorption mechanism of hydrogen sulfide and sulfur dioxide on Au–MoS2 monolayer. Superlattices Microstruct. 2019. [Google Scholar] [CrossRef]
  90. Qian, H.; Lu, W.; Wei, X.; Chen, W.; Deng, J. H2S and SO2 adsorption on Pt-MoS2 adsorbent for partial discharge elimination: A DFT study. Results Phys. 2019. [Google Scholar] [CrossRef]
  91. Yang, Y.; Ashraf, M.A.; Jermsittiparsert, K.; Jiang, L.; Zhang, D. Enhancing the adsorption performance and sensing capability of Ti-doped MoSe2 and MoS2 monolayers by applying electric field. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  92. Ma, S.; Su, L.; Jin, L.; Su, J.; Jin, Y. A first-principles insight into Pd-doped MoSe2 monolayer: A toxic gas scavenger. Phys. Lett. A 2019. [Google Scholar] [CrossRef]
  93. Abbasi, A.; Sardroodi, J.J. A novel strategy for SOx removal by N-doped TiO2/WSe2 nanocomposite as a highly efficient molecule sensor investigated by van der Waals corrected DFT. Comput. Theor. Chem. 2017. [Google Scholar] [CrossRef]
  94. Ni, J.; Wang, W.; Quintana, M.; Jia, F.; Song, S. Adsorption of small gas molecules on strained monolayer WSe2 doped with Pd, Ag, Au, and Pt: A computational investigation. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  95. Dan, M.; Xiang, J.; Wu, F.; Yu, S.; Cai, Q.; Ye, L.; Ye, Y.; Zhou, Y. Rich active-edge-site MoS2 anchored on reduction sites in metal sulfide heterostructure: Toward robust visible light photocatalytic hydrogen sulphide splitting. Appl. Catal. B Environ. 2019. [Google Scholar] [CrossRef]
  96. Li, Y.; Yu, S.; Doronkin, D.E.; Wei, S.; Dan, M.; Wu, F.; Ye, L.; Grunwaldt, J.D.; Zhou, Y. Highly dispersed PdS preferably anchored on In2S3 of MnS/In2S3 composite for effective and stable hydrogen production from H2S. J. Catal. 2019. [Google Scholar] [CrossRef] [Green Version]
  97. Yuan, Y.; Wang, W.; Shi, Y.; Song, L.; Ma, C.; Hu, Y. The influence of highly dispersed Cu2O-anchored MoS2 hybrids on reducing smoke toxicity and fire hazards for rigid polyurethane foam. J. Hazard. Mater. 2020. [Google Scholar] [CrossRef] [PubMed]
  98. Wen, M.Q.; Xiong, T.; Zang, Z.G.; Wei, W.; Tang, X.S.; Dong, F. Synthesis of MoS2/g-C3N4 nanocomposites with enhanced visible-light photocatalytic activity for the removal of nitric oxide (NO). Opt. Express 2016. [Google Scholar] [CrossRef] [PubMed]
  99. Xiong, T.; Wen, M.; Dong, F.; Yu, J.; Han, L.; Lei, B.; Zhang, Y.; Tang, X.; Zang, Z. Three dimensional Z-scheme (BiO)2CO3/MoS2 with enhanced visible light photocatalytic NO removal. Appl. Catal. B Environ. 2016. [Google Scholar] [CrossRef]
  100. Hu, J.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. In situ fabrication of Bi2O2CO3/MoS2 on carbon nanofibers for efficient photocatalytic removal of NO under visible-light irradiation. Appl. Catal. B Environ. 2017. [Google Scholar] [CrossRef]
  101. Zhu, J.; Zhang, H.; Tong, Y.; Zhao, L.; Zhang, Y.; Qiu, Y.; Lin, X. First-principles investigations of metal (V, Nb, Ta)-doped monolayer MoS2: Structural stability, electronic properties and adsorption of gas molecules. Appl. Surf. Sci. 2017. [Google Scholar] [CrossRef]
  102. Ovcharenko, R.; Dedkov, Y.; Voloshina, E. Adsorption of NO2 on WSe2: DFT and photoelectron spectroscopy studies. J. Phys. Condens. Matter 2016. [Google Scholar] [CrossRef] [Green Version]
  103. Ala-Mantila, S.; Heinonen, J.; Junnila, S. Relationship between urbanization, direct and indirect greenhouse gas emissions, and expenditures: A multivariate analysis. Ecol. Econ. 2014. [Google Scholar] [CrossRef]
  104. Denmead, O.T.; Chen, D.; Griffith, D.W.T.; Loh, Z.M.; Bai, M.; Naylor, T. Emissions of the indirect greenhouse gases NH3 and NOx from Australian beef cattle feedlots. Aust. J. Exp. Agric. 2008, 48, 213. [Google Scholar] [CrossRef]
  105. Tai, A.P.K.; Martin, M.V.; Heald, C.L. Threat to future global food security from climate change and ozone air pollution. Nat. Clim. Chang. 2014. [Google Scholar] [CrossRef] [Green Version]
  106. Tjoelker, M.G.; Volin, J.C.; Oleksyn, J.; Reich, P.B. Interaction of ozone pollution and light effects on photosynthesis in a forest canopy experiment. Plant. Cell Environ. 1995. [Google Scholar] [CrossRef]
  107. Hulin, M.; Simoni, M.; Viegi, G.; Annesi-Maesano, I. Respiratory health and indoor air pollutants based on quantitative exposure assessments. Eur. Respir. J. 2012, 40, 1033–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Ajugwo, A.O. Negative effects of gas flaring: The Nigerian experience. J. Environ. Pollut. Hum. Health 2013. [Google Scholar] [CrossRef]
  109. Collins, F.; Orpen, D.; Fay, C.; Foley, C.; Smeaton, A.F.; Diamond, D. Web-based monitoring of year-length deployments of autonomous gas sensing platforms on landfill sites. In Proceedings of the 2011 IEEE Sensors, Limerick, Ireland, 28–31 October 2011; pp. 1620–1623. [Google Scholar]
  110. Rossi, M.; Brunelli, D.; Adami, A.; Lorenzelli, L.; Menna, F.; Remondino, F. Gas-drone: Portable gas sensing system on UAVs for gas leakage localization. In Proceedings of the 2014 IEEE Sensors, Valencia, Spain, 2–5 November 2014; pp. 1431–1434. [Google Scholar]
  111. Li, L.; Fu, C.; Lou, Z.; Chen, S.; Han, W.; Jiang, K.; Chen, D.; Shen, G. Flexible planar concentric circular micro-supercapacitor arrays for wearable gas sensing application. Nano Energy 2017. [Google Scholar] [CrossRef]
  112. Suh, J.H.; Cho, I.; Kang, K.; Kweon, S.J.; Lee, M.; Yoo, H.J.; Park, I. Fully integrated and portable semiconductor-type multi-gas sensing module for IoT applications. Sensors Actuators B Chem. 2018. [Google Scholar] [CrossRef]
  113. Cho, B.; Hahm, M.G.; Choi, M.; Yoon, J.; Kim, A.R.; Lee, Y.J.; Park, S.G.; Kwon, J.D.; Kim, C.S.; Song, M.; et al. Charge-transfer-based gas sensing using atomic-layer MoS2. Sci. Rep. 2015. [Google Scholar] [CrossRef]
  114. Liu, L.; Ikram, M.; Ma, L.; Zhang, X.; Lv, H.; Ullah, M.; Khan, M.; Yu, H.; Shi, K. Edge-exposed MoS2 nanospheres assembled with SnS2 nanosheet to boost NO2 gas sensing at room temperature. J. Hazard. Mater. 2020. [Google Scholar] [CrossRef]
  115. Pham, T.; Li, G.; Bekyarova, E.; Itkis, M.E.; Mulchandani, A. MoS2-based optoelectronic gas sensor with sub-parts-per-billion limit of NO2 gas detection. ACS Nano 2019. [Google Scholar] [CrossRef]
  116. Guo, S.; Yang, D.; Zhang, S.; Dong, Q.; Li, B.; Tran, N.; Li, Z.; Xiong, Y.; Zaghloul, M.E. Development of a cloud-based epidermal MoSe2 device for hazardous gas sensing. Adv. Funct. Mater. 2019. [Google Scholar] [CrossRef]
  117. Zhang, Q.; Shen, Z.; Cao, J.; Zhang, R.; Zhang, L.; Huang, R.J.; Zheng, C.; Wang, L.; Liu, S.; Xu, H.; et al. Variations in PM2.5, TSP, BC, and trace gases (NO2, SO2, and O3) between haze and non-haze episodes in winter over Xi’an, China. Atmos. Environ. 2015. [Google Scholar] [CrossRef]
  118. Pijolat, C.; Pupier, C.; Sauvan, M.; Tournier, G.; Lalauze, R. Gas detection for automotive pollution control. Sensors Actuators B Chem. 1999. [Google Scholar] [CrossRef]
  119. Yamazoe, N.; Miura, N. Environmental gas sensing. Sensors Actuators B Chem. 1994. [Google Scholar] [CrossRef]
  120. Long, H.; Harley-Trochimczyk, A.; Pham, T.; Tang, Z.; Shi, T.; Zettl, A.; Carraro, C.; Worsley, M.A.; Maboudian, R. High surface area MoS2/graphene hybrid aerogel for ultrasensitive NO2 detection. Adv. Funct. Mater. 2016. [Google Scholar] [CrossRef] [Green Version]
  121. Liu, B.; Chen, L.; Liu, G.; Abbas, A.N.; Fathi, M.; Zhou, C. High-performance chemical sensing using Schottky-contacted chemical vapor deposition grown monolayer MoS2 transistors. ACS Nano 2014. [Google Scholar] [CrossRef]
  122. Yan, W.; Worsley, M.A.; Pham, T.; Zettl, A.; Carraro, C.; Maboudian, R. Effects of ambient humidity and temperature on the NO2 sensing characteristics of WS2/graphene aerogel. Appl. Surf. Sci. 2018. [Google Scholar] [CrossRef]
  123. Zhou, Y.; Liu, G.; Zhu, X.; Guo, Y. Ultrasensitive NO2 gas sensing based on rGO/MoS2 nanocomposite film at low temperature. Sensors Actuators B Chem. 2017. [Google Scholar] [CrossRef]
  124. Mirõ, P.; Ghorbani-Asl, M.; Heine, T. Two dimensional materials beyond MoS2: Noble-transition-metal dichalcogenides. Angew. Chemie Int. Ed. 2014. [Google Scholar] [CrossRef]
  125. Sajjad, M.; Montes, E.; Singh, N.; Schwingenschlögl, U. Superior gas sensing properties of monolayer PtSe2. Adv. Mater. Interfaces 2017. [Google Scholar] [CrossRef]
  126. Yim, C.; Lee, K.; McEvoy, N.; O’Brien, M.; Riazimehr, S.; Berner, N.C.; Cullen, C.P.; Kotakoski, J.; Meyer, J.C.; Lemme, M.C.; et al. High-performance hybrid electronic devices from layered PtSe2 films grown at low temperature. ACS Nano 2016. [Google Scholar] [CrossRef] [Green Version]
  127. Wu, E.; Xie, Y.; Yuan, B.; Zhang, H.; Hu, X.; Liu, J.; Zhang, D. Ultrasensitive and fully reversible NO2 gas sensing based on p-type MoTe2 under ultraviolet illumination. ACS Sensors 2018. [Google Scholar] [CrossRef] [PubMed]
  128. Ko, K.Y.; Song, J.G.; Kim, Y.; Choi, T.; Shin, S.; Lee, C.W.; Lee, K.; Koo, J.; Lee, H.; Kim, J.; et al. Improvement of gas-sensing performance of large-area tungsten disulfide nanosheets by surface functionalization. ACS Nano 2016. [Google Scholar] [CrossRef] [PubMed]
  129. Cho, B.; Kim, A.R.; Park, Y.; Yoon, J.; Lee, Y.J.; Lee, S.; Yoo, T.J.; Kang, C.G.; Lee, B.H.; Ko, H.C.; et al. Bifunctional sensing characteristics of chemical vapor deposition synthesized atomic-layered MoS2. ACS Appl. Mater. Interfaces 2015. [Google Scholar] [CrossRef] [PubMed]
  130. Wang, J.; Deng, J.; Li, Y.; Yuan, H.; Xu, M. ZnO nanocrystal-coated MoS2 nanosheets with enhanced ultraviolet light gas sensitive activity studied by surface photovoltage technique. Ceram. Int. 2020. [Google Scholar] [CrossRef]
  131. Baek, J.; Yin, D.; Liu, N.; Omkaram, I.; Jung, C.; Im, H.; Hong, S.; Kim, S.M.; Hong, Y.K.; Hur, J.; et al. A highly sensitive chemical gas detecting transistor based on highly crystalline CVD-grown MoSe2 films. Nano Res. 2017. [Google Scholar] [CrossRef]
  132. Perrozzi, F.; Emamjomeh, S.M.; Paolucci, V.; Taglieri, G.; Ottaviano, L.; Cantalini, C. Thermal stability of WS2 flakes and gas sensing properties of WS2/WO3 composite to H2, NH3 and NO2. Sensors Actuators B Chem. 2017. [Google Scholar] [CrossRef]
  133. Han, S.; Bian, H.; Feng, Y.; Liu, A.; Li, X.; Zeng, F.; Zhang, X. Analysis of the relationship between O3, NO and NO2 in Tianjin, China. Aerosol Air Qual. Res. 2011. [Google Scholar] [CrossRef] [Green Version]
  134. Jones, A.E.; Weller, R.; Wolff, E.W.; Jacobi, H.W. Speciation and rate of photochemical NO and NO2 production in Antarctic snow. Geophys. Res. Lett. 2000. [Google Scholar] [CrossRef] [Green Version]
  135. Environmental Protection Agency (EPA). Nitrogen Oxides (NOx), Why and How They Are Controlled; Technical Bulletin No. EPA-456/F-99-006R; EPA: Research Triangle Park, NC, USA, 1999.
  136. Zhao, S.; Xue, J.; Kang, W. Gas adsorption on MoS2 monolayer from first-principles calculations. Chem. Phys. Lett. 2014. [Google Scholar] [CrossRef] [Green Version]
  137. Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.H.; Tok, A.I.Y.; Zhang, Q.; Zhang, H. Fabrication of single- and multilayer MoS2 film-based field-effect transistors for sensing NO at room temperature. Small 2012. [Google Scholar] [CrossRef]
  138. Ramu, S.; Chandrakalavathi, T.; Murali, G.; Kumar, K.S.; Sudharani, A.; Ramanadha, M.; Peta, K.R.; Jeyalakshmi, R.; Vijayalakshmi, R.P. UV enhanced NO gas sensing properties of the MoS2 monolayer gas sensor. Mater. Res. Express 2019. [Google Scholar] [CrossRef]
  139. Burman, D.; Ghosh, R.; Santra, S.; Kumar Ray, S.; Kumar Guha, P. Role of vacancy sites and UV-ozone treatment on few layered MoS2 nanoflakes for toxic gas detection. Nanotechnology 2017. [Google Scholar] [CrossRef]
  140. Medina, H.; Li, J.G.; Su, T.Y.; Lan, Y.W.; Lee, S.H.; Chen, C.W.; Chen, Y.Z.; Manikandan, A.; Tsai, S.H.; Navabi, A.; et al. Wafer-scale growth of WSe2 monolayers toward phase-engineered hybrid WOx/WSe2 films with sub-ppb NOx gas sensing by a low-temperature plasma-assisted selenization process. Chem. Mater. 2017. [Google Scholar] [CrossRef]
  141. Committee on the Environment and Natural Resources Air Quality Research Subcommittee. Atmospheric Ammonia: Sources and Fate. A review of Ongoing Federal Research and Future Needs; NOAA Aeronomy Laboratory: Boulder, CO, USA, 2000. Available online: https://www.esrl.noaa.gov/csl/aqrsd/reports/ammonia.pdf (accessed on 2 May 2020).
  142. Pinder, R.W.; Gilliland, A.B.; Dennis, R.L. Environmental impact of atmospheric NH3 emissions under present and future conditions in the eastern United States. Geophys. Res. Lett. 2008. [Google Scholar] [CrossRef]
  143. Donham, K.J.; Cumro, D.; Reynolds, S. Synergistic effects of dust and ammonia on the occupational health effects of poultry production workers. J. Agromedicine 2002. [Google Scholar] [CrossRef] [PubMed]
  144. Guo, S.; Yang, D.; Li, B.; Dong, Q.; Li, Z.; Zaghloul, M.E. An artificial intelligent flexible gas sensor based on ultra-large area MoSe2 nanosheet. In Proceedings of the 2019 IEEE 62nd International Midwest Symposium on Circuits and Systems (MWSCAS), Dallas, TX, USA, 4–7 August 2019; pp. 884–887. [Google Scholar]
  145. Luo, H.; Cao, Y.; Zhou, J.; Feng, J.; Cao, J.; Guo, H. Adsorption of NO2, NH3 on monolayer MoS2 doped with Al, Si, and P: A first-principles study. Chem. Phys. Lett. 2016. [Google Scholar] [CrossRef]
  146. Cho, B.; Yoon, J.; Lim, S.K.; Kim, A.R.; Kim, D.H.; Park, S.G.; Kwon, J.D.; Lee, Y.J.; Lee, K.H.; Lee, B.H.; et al. Chemical sensing of 2D graphene/MoS2 heterostructure device. ACS Appl. Mater. Interfaces 2015. [Google Scholar] [CrossRef] [PubMed]
  147. Feng, Z.; Xie, Y.; Chen, J.; Yu, Y.; Zheng, S.; Zhang, R.; Li, Q.; Chen, X.; Sun, C.; Zhang, H.; et al. Highly sensitive MoTe2 chemical sensor with fast recovery rate through gate biasing. 2D Mater. 2017. [Google Scholar] [CrossRef]
  148. Qin, Z.; Zeng, D.; Zhang, J.; Wu, C.; Wen, Y.; Shan, B.; Xie, C. Effect of layer number on recovery rate of WS2 nanosheets for ammonia detection at room temperature. Appl. Surf. Sci. 2017. [Google Scholar] [CrossRef]
  149. Late, D.J.; Doneux, T.; Bougouma, M. Single-layer MoSe2 based NH3 gas sensor. Appl. Phys. Lett. 2014. [Google Scholar] [CrossRef]
  150. Zhang, D.; Jiang, C.; Li, P.; Sun, Y. Layer-by-layer self-assembly of Co3O4 nanorod-decorated MoS2 nanosheet-based nanocomposite toward high-performance ammonia detection. ACS Appl. Mater. Interfaces 2017. [Google Scholar] [CrossRef] [PubMed]
  151. Abun, A.; Huang, B.R.; Saravanan, A.; Kathiravan, D.; Hong, P. Da Effect of PMMA on the surface of exfoliated MoS2 nanosheets and their highly enhanced ammonia gas sensing properties at room temperature. J. Alloys Compd. 2020. [Google Scholar] [CrossRef]
  152. Zhang, S.; Wang, J.; Torad, N.L.; Xia, W.; Aslam, M.A.; Kaneti, Y.V.; Hou, Z.; Ding, Z.; Da, B.; Fatehmulla, A.; et al. Rational design of nanoporous MoS2/VS2 heteroarchitecture for ultrahigh performance ammonia sensors. Small 2020. [Google Scholar] [CrossRef]
  153. Durmusoglu, E.; Taspinar, F.; Karademir, A. Health risk assessment of BTEX emissions in the landfill environment. J. Hazard. Mater. 2010. [Google Scholar] [CrossRef] [PubMed]
  154. Dai, H.; Jing, S.; Wang, H.; Ma, Y.; Li, L.; Song, W.; Kan, H. VOC characteristics and inhalation health risks in newly renovated residences in Shanghai, China. Sci. Total Environ. 2017. [Google Scholar] [CrossRef]
  155. Cetin, E.; Odabasi, M.; Seyfioglu, R. Ambient volatile organic compound (VOC) concentrations around a petrochemical complex and a petroleum refinery. Sci. Total Environ. 2003. [Google Scholar] [CrossRef]
  156. Malherbe, L.; Mandin, C. VOC emissions during outdoor ship painting and health-risk assessment. Atmos. Environ. 2007. [Google Scholar] [CrossRef]
  157. Chen, C.; Tsow, F.; Campbell, K.D.; Iglesias, R.; Forzani, E.; Tao, N. A wireless hybrid chemical sensor for detection of environmental volatile organic compounds. IEEE Sens. J. 2013. [Google Scholar] [CrossRef] [Green Version]
  158. Sun, X.; Shao, K.; Wang, T. Detection of volatile organic compounds (VOCs) from exhaled breath as noninvasive methods for cancer diagnosis. Anal. Bioanal. Chem. 2016. [Google Scholar] [CrossRef]
  159. Barzegar, M.; Berahman, M.; Asgari, R. First-principles study of molecule adsorption on Ni-decorated monolayer MoS2. J. Comput. Electron. 2019. [Google Scholar] [CrossRef]
  160. Kim, J.S.; Yoo, H.W.; Choi, H.O.; Jung, H.T. Tunable volatile organic compounds sensor by using thiolated ligand conjugation on MoS2. Nano Lett. 2014. [Google Scholar] [CrossRef] [PubMed]
  161. Tomer, V.K.; Malik, R.; Chaudhary, V.; Mishra, Y.K.; Kienle, L.; Ahuja, R.; Lin, L. Superior visible light photocatalysis and low-operating temperature VOCs sensor using cubic Ag(0)-MoS2 loaded g-CN 3D porous hybrid. Appl. Mater. Today 2019. [Google Scholar] [CrossRef]
  162. Zhao, C.; Gan, X.; Yuan, Q.; Hu, S.; Fang, L.; Zhao, J. High-performance volatile organic compounds microsensor based on few-layer MoS2-coated photonic crystal cavity. Adv. Opt. Mater. 2018. [Google Scholar] [CrossRef]
  163. Chen, D.; Tang, J.; Zhang, X.; Li, Y.; Liu, H. Detecting decompositions of sulfur hexafluoride using MoS2 monolayer as gas sensor. IEEE Sens. J. 2019. [Google Scholar] [CrossRef]
  164. Chen, D.; Zhang, X.; Tang, J.; Cui, Z.; Cui, H.; Pi, S. Theoretical study of monolayer PtSe2 as outstanding gas sensor to detect SF6 decompositions. IEEE Electron Device Lett. 2018. [Google Scholar] [CrossRef]
  165. Park, J.; Mun, J.; Shin, J.S.; Kang, S.W. Highly sensitive two dimensional MoS2 gas sensor decorated with Pt nanoparticles. R. Soc. Open Sci. 2018. [Google Scholar] [CrossRef] [Green Version]
  166. Yang, Y.; Jermsittiparsert, K.; Gao, W.; Zhang, D. Structural, electronic and magnetic properties of the Ni/Cu-embedded S-vacancy defective MoS2 monolayers and their effects on the adsorption of SOx and O3 molecules. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  167. Ma, D.; Ju, W.; Li, T.; Zhang, X.; He, C.; Ma, B.; Lu, Z.; Yang, Z. The adsorption of CO and NO on the MoS2 monolayer doped with Au, Pt, Pd, or Ni: A first-principles study. Appl. Surf. Sci. 2016. [Google Scholar] [CrossRef]
  168. Shen, J.; Yang, Z.; Wang, Y.; Xu, L.C.; Liu, R.; Liu, X. The gas sensing performance of borophene/MoS2 heterostructure. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  169. Sahoo, M.P.K.; Wang, J.; Zhang, Y.; Shimada, T.; Kitamura, T. Modulation of gas adsorption and magnetic properties of monolayer-MoS2 by antisite defect and strain. J. Phys. Chem. C 2016. [Google Scholar] [CrossRef]
  170. Ma, D.; Ma, B.; Lu, Z.; He, C.; Tang, Y.; Lu, Z.; Yang, Z. Interaction between H2O, N2, CO, NO, NO2 and N2O molecules and a defective WSe2 monolayer. Phys. Chem. Chem. Phys. 2017. [Google Scholar] [CrossRef]
  171. Jasmine, J.M.; Aadhityan, A.; Preferencial kala, C.; Thiruvadigal, D.J. A first-principles study of Cl2, PH3, AsH3, BBr3 and SF4 gas adsorption on MoS2 monolayer with S and Mo vacancy. Appl. Surf. Sci. 2019. [Google Scholar] [CrossRef]
  172. UNEP. Tackling Global Water Pollution. Available online: https://www.unenvironment.org/explore-topics/water/what-we-do/tackling-global-water-pollution (accessed on 2 May 2020).
  173. Malik, S.N.; Khan, S.M.; Ghosh, P.C.; Vaidya, A.N.; Kanade, G.; Mudliar, S.N. Treatment of pharmaceutical industrial wastewater by nano-catalyzed ozonation in a semi-batch reactor for improved biodegradability. Sci. Total Environ. 2019, 678, 114–122. [Google Scholar] [CrossRef] [PubMed]
  174. Khan, N.A.; Khan, K.A.; Islam, M. Water and wastewater treatment using nano-technology. In Chemistry of Phytopotentials: Health, Energy and Environmental Perspectives; Springer: Berlin/Heidelberg, Germany, 2012; pp. 315–318. ISBN 978-3-642-23394-4. [Google Scholar]
  175. Prince, R.C. Oil spill dispersants: Boon or bane? Environ. Sci. Technol. 2015, 49, 6376–6384. [Google Scholar] [CrossRef] [Green Version]
  176. Abiodun, O.I.; Jantan, A.; Omolara, A.E.; Dada, K.V.; Mohamed, N.A.; Arshad, H. State-of-the-art in artificial neural network applications: A survey. Heliyon 2018, 4, e00938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Gusain, R.; Kumar, N.; Fosso-Kankeu, E.; Ray, S.S. Efficient removal of Pb(II) and Cd(II) from industrial mine water by a hierarchical MoS2/SH-MWCNT nanocomposite. ACS Omega 2019, 4, 13922–13935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Göktaş, R.K.; MacLeod, M. Remoteness from sources of persistent organic pollutants in the multi-media global environment. Environ. Pollut. 2016, 217, 33–41. [Google Scholar] [CrossRef] [Green Version]
  179. Wang, Z.; Wu, A.; Ciacchi, L.C.; Wei, G. Recent advances in nanoporous membranes for water purification. Nanomaterials 2018, 8, 65. [Google Scholar] [CrossRef] [Green Version]
  180. SDG Knowledge Platform. Sustainable Development Goals. Available online: https://sustainabledevelopment.un.org/?menu=1300 (accessed on 2 May 2020).
  181. Lu, H.; Wang, J.; Stoller, M.; Wang, T.; Bao, Y.; Hao, H. An overview of nanomaterials for water and wastewater treatment. Adv. Mater. Sci. Eng. 2016, 2016, 4964828. [Google Scholar] [CrossRef] [Green Version]
  182. Li, G.; Wang, Y.; Bi, J.; Huang, X.; Mao, Y.; Luo, L.; Hao, H. Partial oxidation strategy to synthesize WS2/WO3 heterostructure with enhanced adsorption performance for organic dyes: Synthesis, modelling, and mechanism. Nanomaterials 2020, 10, 278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Wang, Z.; Mi, B. Environmental applications of 2D molybdenum disulfide (MoS2) nanosheets. Environ. Sci. Technol. 2017, 51, 8229–8244. [Google Scholar] [CrossRef] [PubMed]
  184. Tang, H.; Huang, H.; Wang, X.; Wu, K.; Tang, G.; Li, C. Hydrothermal synthesis of 3D hierarchical flower-like MoSe2 microspheres and their adsorption performances for methyl orange. Appl. Surf. Sci. 2016, 379, 296–303. [Google Scholar] [CrossRef]
  185. Fang, Y.; Huang, Q.; Liu, P.; Shi, J.; Xu, G. Easy-separative MoS2-glue sponges with high-efficient dye adsorption and excellent reusability for convenient water treatment. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 540, 112–122. [Google Scholar] [CrossRef]
  186. Song, H.J.; You, S.; Jia, X.H.; Yang, J. MoS2 nanosheets decorated with magnetic Fe3O4 nanoparticles and their ultrafast adsorption for wastewater treatment. Ceram. Int. 2015, 41, 13896–13902. [Google Scholar] [CrossRef]
  187. Li, H.; Xie, F.; Li, W.; Fahlman, B.D.; Chen, M.; Li, W. Preparation and adsorption capacity of porous MoS2 nanosheets. RSC Adv. 2016, 6, 105222–105230. [Google Scholar] [CrossRef]
  188. Massey, A.T.; Gusain, R.; Kumari, S.; Khatri, O.P. Hierarchical microspheres of MoS2 nanosheets: Efficient and regenerative adsorbent for removal of water-soluble dyes. Ind. Eng. Chem. Res. 2016, 55, 7124–7131. [Google Scholar] [CrossRef]
  189. Krishna Kumar, A.S.; Jiang, S.J.; Warchoł, J.K. Synthesis and characterization of two-dimensional transition metal dichalcogenide magnetic MoS2@Fe3O4 nanoparticles for adsorption of Cr(VI)/Cr(III). ACS Omega 2017, 2, 6187–6200. [Google Scholar] [CrossRef] [PubMed]
  190. Wan, Z.; Li, D.; Jiao, Y.; Ouyang, X.; Chang, L.; Wang, X. Bifunctional MoS2 coated melamine-formaldehyde sponges for efficient oil–water separation and water-soluble dye removal. Appl. Mater. Today 2017, 9, 551–559. [Google Scholar] [CrossRef]
  191. Li, R.; Deng, H.; Zhang, X.; Wang, J.J.; Awasthi, M.K.; Wang, Q.; Xiao, R.; Zhou, B.; Du, J.; Zhang, Z. High-efficiency removal of Pb(II) and humate by a CeO2–MoS2 hybrid magnetic biochar. Bioresour. Technol. 2019, 273, 335–340. [Google Scholar] [CrossRef]
  192. Wan, Z.; Liu, Y.; Chen, S.; Song, K.; Peng, Y.; Zhao, N.; Ouyang, X.; Wang, X. Facile fabrication of a highly durable and flexible MoS2@RTV sponge for efficient oil-water separation. Colloids Surfaces A Physicochem. Eng. Asp. 2018, 546, 237–243. [Google Scholar] [CrossRef]
  193. Heiranian, M.; Farimani, A.B.; Aluru, N.R. Water desalination with a single-layer MoS2 nanopore. Nat. Commun. 2015, 6, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Kou, J.; Yao, J.; Wu, L.; Zhou, X.; Lu, H.; Wu, F.; Fan, J. Nanoporous two-dimensional MoS2 membranes for fast saline solution purification. Phys. Chem. Chem. Phys. 2016. [Google Scholar] [CrossRef] [PubMed]
  195. Kozubek, R.; Tripathi, M.; Ghorbani-Asl, M.; Kretschmer, S.; Madauß, L.; Pollmann, E.; O’Brien, M.; McEvoy, N.; Ludacka, U.; Susi, T.; et al. Perforating freestanding molybdenum disulfide monolayers with highly charged ions. J. Phys. Chem. Lett. 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Ma, J.; Tang, X.; He, Y.; Fan, Y.; Chen, J.; Hao, Y. Robust stable MoS2/GO filtration membrane for effective removal of dyes and salts from water with enhanced permeability. Desalination 2020, 480, 114328. [Google Scholar] [CrossRef]
  197. Gao, J.; Zhang, M.; Wang, J.; Liu, G.; Liu, H.; Jiang, Y. Bioinspired modification of layer-stacked molybdenum disulfide (MoS2) membranes for enhanced nanofiltration performance. ACS Omega 2019, 4, 4012–4022. [Google Scholar] [CrossRef] [Green Version]
  198. Chu, H.; Liu, X.; Liu, B.; Zhu, G.; Lei, W.; Du, H.; Liu, J.; Li, J.; Li, C.; Sun, C. Hexagonal 2H-MoSe2 broad spectrum active photocatalyst for Cr(VI) reduction. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  199. Mittal, H.; Khanuja, M. Nanosheets- and nanourchins-like nanostructures of MoSe2 for photocatalytic water purification: Kinetics and reusability study. Environ. Sci. Pollut. Res. 2019, 1–13. [Google Scholar] [CrossRef]
  200. Zhou, X.; Yao, J.; Yang, M.; Ma, J.; Zhou, Q.; Ou, E.; Zhang, Z.; Sun, X. Synthesis of MoSe2/SrTiO3 heterostructures with enhanced ultraviolet-light-driven and visible-light-driven photocatalytic properties. Nano 2018, 13. [Google Scholar] [CrossRef] [Green Version]
  201. Kridiotis, A.C.; Longwell, J.P.; Sarofim, A.F.; Bar-ziv, E. Application of a stochastic model of imperfect mixing to the combustion of fuel-lean COH2 mixtures in air. Chem. Eng. Sci. 1989. [Google Scholar] [CrossRef]
  202. Kim, J.H.; Ma, X.; Zhou, A.; Song, C. Ultra-deep desulfurization and denitrogenation of diesel fuel by selective adsorption over three different adsorbents: A study on adsorptive selectivity and mechanism. Catal. Today 2006, 111, 74–83. [Google Scholar] [CrossRef]
  203. Kalam, M.A.; Masjuki, H.H. Emissions and deposit characteristics of a small diesel engine when operated on preheated crude palm oil. Biomass Bioenergy 2004. [Google Scholar] [CrossRef]
  204. Ashraful, A.M.; Masjuki, H.H.; Kalam, M.A.; Rizwanul Fattah, I.M.; Imtenan, S.; Shahir, S.A.; Mobarak, H.M. Production and comparison of fuel properties, engine performance, and emission characteristics of biodiesel from various non-edible vegetable oils: A review. Energy Convers. Manag. 2014, 80, 202–228. [Google Scholar] [CrossRef]
  205. Song, C.; Ma, X. New design approaches to ultra-clean diesel fuels by deep desulfurization and deep dearomatization. Appl. Catal. B Environ. 2003, 41, 207–238. [Google Scholar] [CrossRef]
  206. Itthibenchapong, V.; Srifa, A.; Kaewmeesri, R.; Kidkhunthod, P.; Faungnawakij, K. Deoxygenation of palm kernel oil to jet fuel-like hydrocarbons using Ni-MoS2/γ-Al2O3 catalysts. Energy Convers. Manag. 2017. [Google Scholar] [CrossRef]
  207. Jiang, L.; Kronbak, J.; Christensen, L.P. The costs and benefits of sulphur reduction measures: Sulphur scrubbers versus marine gas oil. Transp. Res. Part D Transp. Environ. 2014. [Google Scholar] [CrossRef]
  208. Ushakov, S.; Valland, H.; Nielsen, J.B.; Hennie, E. Effects of high sulphur content in marine fuels on particulate matter emission characteristics. J. Mar. Eng. Technol. 2013. [Google Scholar] [CrossRef]
  209. Jones, J.; Dupont, V.; Brydson, R.; Fullerton, D.; Nasri, N.; Ross, A.; Westwood, A.V. Sulphur poisoning and regeneration of precious metal catalysed methane combustion. Catal. Today 2003, 81, 589–601. [Google Scholar] [CrossRef]
  210. Paul, J.F.; Payen, E. Vacancy formation on MoS2 hydrodesulfurization catalyst: DFT study of the mechanism. J. Phys. Chem. B 2003. [Google Scholar] [CrossRef]
  211. Grimblot, J. Genesis, architecture and nature of sites of Co(Ni)-MoS2 supported hydroprocessing catalysts. Catal. Today 1998. [Google Scholar] [CrossRef]
  212. Liaw, S.J.; Raje, A.; Chary, K.V.R.; Davis, B.H. Catalytic hydrotreatment of Illinois No. 6 coal-derived naphtha: The removal of individual nitrogen and sulfur compounds over MoS2 and Mo2N. Appl. Catal. A Gen. 1995. [Google Scholar] [CrossRef]
  213. Mahmoudabadi, Z.S.; Rashidi, A.; Tavasoli, A. Synthesis of MoS2 quantum dots as a nanocatalyst for hydrodesulfurization of naphtha: Experimental and DFT study. J. Environ. Chem. Eng. 2020. [Google Scholar] [CrossRef]
  214. Olivas, A.; Zepeda, T.A.; Villalpando, I.; Fuentes, S. Performance of unsupported Ni(Co,Fe)/MoS2 catalysts in hydrotreating reactions. Catal. Commun. 2008. [Google Scholar] [CrossRef]
  215. Liu, L.H.; Liu, D.; Liu, B.; Li, G.C.; Liu, Y.Q.; Liu, C.G. Relation between the morphology of MoS2 in NiMo catalyst and its selectivity for dibenzothiophene hydrodesulfurization. Ranliao Huaxue Xuebao J. Fuel Chem. Technol. 2011. [Google Scholar] [CrossRef]
  216. Rangarajan, S.; Mavrikakis, M. On the preferred active sites of promoted MoS2 for hydrodesulfurization with minimal organonitrogen inhibition. ACS Catal. 2017. [Google Scholar] [CrossRef]
  217. Tye, C.T.; Smith, K.J. Hydrodesulfurization of dibenzothiophene over exfoliated MoS2 catalyst. Catal. Today 2006. [Google Scholar] [CrossRef]
  218. Abbasi, A.; Karimi, A.; Aghabozorg, H.; Sadighi, S.; Aval, M.A. Cobalt-promoted MoS2 nanosheets: A promising novel diesel hydrodesulfurization catalyst. Int. J. Chem. Kinet. 2020. [Google Scholar] [CrossRef]
  219. Moses, P.G.; Hinnemann, B.; Topsøe, H.; Nørskov, J.K. The hydrogenation and direct desulfurization reaction pathway in thiophene hydrodesulfurization over MoS2 catalysts at realistic conditions: A density functional study. J. Catal. 2007. [Google Scholar] [CrossRef]
  220. Kaluža, L.; Zdražil, M.; Žilková, N.; Čejka, J. High activity of highly loaded MoS2 hydrodesulfurization catalysts supported on organised mesoporous alumina. Catal. Commun. 2002. [Google Scholar] [CrossRef]
  221. Liu, N.; Wang, X.; Xu, W.; Hu, H.; Liang, J.; Qiu, J. Microwave-assisted synthesis of MoS2/graphene nanocomposites for efficient hydrodesulfurization. Fuel 2014. [Google Scholar] [CrossRef]
  222. Yang, L.; Wang, X.z.; Liu, Y.; Yu, Z.f.; Liang, J.j.; Chen, B.b.; Shi, C.; Tian, S.; Li, X.; Qiu, J.-S. Monolayer MoS2 anchored on reduced graphene oxide nanosheets for efficient hydrodesulfurization. Appl. Catal. B Environ. 2017. [Google Scholar] [CrossRef]
  223. Liu, G.; Robertson, A.W.; Li, M.M.J.; Kuo, W.C.H.; Darby, M.T.; Muhieddine, M.H.; Lin, Y.C.; Suenaga, K.; Stamatakis, M.; Warner, J.H.; et al. MoS2 monolayer catalyst doped with isolated Co atoms for the hydrodeoxygenation reaction. Nat. Chem. 2017. [Google Scholar] [CrossRef] [PubMed]
  224. Li, Q.; Bai, X.; Ling, C.; Zhou, Q.; Yuan, S.; Chen, Q.; Wang, J. Forming atom–vacancy interface on the MoS2 catalyst for efficient hydrodeoxygenation reactions. Small Methods 2019. [Google Scholar] [CrossRef]
  225. Wu, K.; Wang, W.; Guo, H.; Yang, Y.; Huang, Y.; Li, W.; Li, C. Engineering co nanoparticles supported on defect MoS2–x for mild deoxygenation of lignin-derived phenols to arenes. ACS Energy Lett. 2020. [Google Scholar] [CrossRef]
  226. Varakin, A.N.; Mozhaev, A.V.; Pimerzin, A.A.; Nikulshin, P.A. Toward HYD/DEC selectivity control in hydrodeoxygenation over supported and unsupported Co(Ni)-MoS2 catalysts. A key to effective dual-bed catalyst reactor for co-hydroprocessing of diesel and vegetable oil. Catal. Today 2019. [Google Scholar] [CrossRef]
  227. Alvarez-Galvan, M.; Campos-Martin, J.; Fierro, J. Transition metal phosphides for the catalytic hydrodeoxygenation of waste oils into green diesel. Catalysts 2019, 9, 293. [Google Scholar] [CrossRef] [Green Version]
  228. Li, N.; Wei, L.; Bibi, R.; Chen, L.; Liu, J.; Zhang, L.; Zheng, Y.; Zhou, J. Catalytic hydrogenation of alkali lignin into bio-oil using flower-like hierarchical MoS2-based composite catalysts. Fuel 2016. [Google Scholar] [CrossRef]
  229. Zhou, X.; Zhou, J.; Yu, Y.; Ma, J.; Sun, X.; Hu, L. Catalytic hydrogenation of alkali lignin into bio-oil via nano-lamellar MoSe2-based composite catalysts. Nano 2017. [Google Scholar] [CrossRef]
  230. Porsin, A.A.; Vlasova, E.N.; Bukhtiyarova, G.A.; Nuzhdin, A.L.; Bukhtiyarov, V.I. Sulfide catalysts for production of motor fuels from fatty acid triglycerides. Russ. J. Appl. Chem. 2018, 91, 1905–1911. [Google Scholar] [CrossRef]
  231. Zhang, Y.; Meng, Z.; Shi, Q.; Gao, H.; Liu, Y.; Wang, Y.; Rao, D.; Deng, K.; Lu, R. Nanoporous MoS2 monolayer as a promising membrane for purifying hydrogen and enriching methane. J. Phys. Condens. Matter 2017. [Google Scholar] [CrossRef]
  232. Yu, L.; Mishra, I.K.; Xie, Y.; Zhou, H.; Sun, J.; Zhou, J.; Ni, Y.; Luo, D.; Yu, F.; Yu, Y.; et al. Ternary Ni2(1−x)Mo2xP nanowire arrays toward efficient and stable hydrogen evolution electrocatalysis under large-current-density. Nano Energy 2018. [Google Scholar] [CrossRef]
  233. Mohamed, M.; Yusup, S.; Loy, A.C.M. Effect of empty fruit bunch in calcium oxide for cyclic CO2 capture. Chem. Eng. Technol. 2019. [Google Scholar] [CrossRef]
  234. Yamamoto, A.; Shinkai, T.; Loy, A.C.M.; Mohamed, M.; Balean, F.H.; Yusup, S.; Quitain, A.T.; Kida, T. Application of a solid electrolyte CO2 sensor to the performance evaluation of CO2 capture materials. Sensors Actuators B Chem. 2020. [Google Scholar] [CrossRef]
  235. Holloway, S. Underground sequestration of carbon dioxide—A viable greenhouse gas mitigation option. Energy 2005, 30, 2318–2333. [Google Scholar] [CrossRef]
  236. Song, Y.; Peng, R.; Hensley, D.K.; Bonnesen, P.V.; Liang, L.; Wu, Z.; Meyer, H.M.; Chi, M.; Ma, C.; Sumpter, B.G.; et al. High-selectivity electrochemical conversion of CO2 to ethanol using a copper nanoparticle/N-doped graphene electrode. ChemistrySelect 2016. [Google Scholar] [CrossRef] [Green Version]
  237. Yang, X.; Zhang, R.; Fu, J.; Geng, S.; Cheng, J.J.; Sun, Y. Pyrolysis kinetic and product analysis of different microalgal biomass by distributed activation energy model and pyrolysis-gas chromatography-mass spectrometry. Bioresour. Technol. 2014. [Google Scholar] [CrossRef] [PubMed]
  238. Manenti, F. syngas. In Computer Aided Chemical Engineering; Elsevier BV: Amsterdam, The Netherlands, 2015. [Google Scholar]
  239. Marlin, D.S.; Sarron, E.; Sigurbjörnsson, Ó. Process advantages of direct CO2 to methanol synthesis. Front. Chem. 2018. [Google Scholar] [CrossRef] [PubMed]
  240. Melo Bravo, P.; Debecker, D.P. Combining CO2 capture and catalytic conversion to methane. Waste Dispos. Sustain. Energy 2019. [Google Scholar] [CrossRef] [Green Version]
  241. Nguyen, D.L.T.; Kim, Y.; Hwang, Y.J.; Won, D.H. Progress in development of electrocatalyst for CO2 conversion to selective CO production. Carbon Energy 2020. [Google Scholar] [CrossRef] [Green Version]
  242. Ali, N.; Bilal, M.; Nazir, M.S.; Khan, A.; Ali, F.; Iqbal, H.M.N. Thermochemical and electrochemical aspects of carbon dioxide methanation: A sustainable approach to generate fuel via waste to energy theme. Sci. Total Environ. 2020. [Google Scholar] [CrossRef]
  243. Wang, Z.; Roberts, R.R.; Naterer, G.F.; Gabriel, K.S. Comparison of thermochemical, electrolytic, photoelectrolytic and photochemical solar-to-hydrogen production technologies. Int. J. Hydrogen Energy 2012, 37, 16287–16301. [Google Scholar] [CrossRef]
  244. Zhang, J.-H.; Zhou, Y.-G. Nano-impact electrochemistry: Analysis of single bioentities. TrAC Trends Anal. Chem. 2020, 123, 115768. [Google Scholar] [CrossRef]
  245. Abdullah, H.; Khan, M.M.R.; Ong, H.R.; Yaakob, Z. Modified TiO2 photocatalyst for CO2 photocatalytic reduction: An overview. J. CO2 Util. 2017, 22, 15–32. [Google Scholar] [CrossRef]
  246. Xie, S.; Zhang, Q.; Liu, G.; Wang, Y. Photocatalytic and photoelectrocatalytic reduction of CO2 using heterogeneous catalysts with controlled nanostructures. Chem. Commun. 2016. [Google Scholar] [CrossRef] [PubMed]
  247. Kisch, H. Semiconductor photocatalysis-mechanistic and synthetic aspects. Angew. Chemie Int. Ed. 2013, 52, 812–847. [Google Scholar] [CrossRef] [PubMed]
  248. Xu, Q.; Zhang, L.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Jaroniec, M. Direct Z-scheme photocatalysts: Principles, synthesis, and applications. Mater. Today 2018, 21, 1042–1063. [Google Scholar] [CrossRef]
  249. Geioushy, R.A.; El-Sheikh, S.M.; Hegazy, I.M.; Shawky, A.; El-Sherbiny, S.; Kandil, A.H.T. Insights into two-dimensional MoS2 sheets for enhanced CO2 photoreduction to C1 and C2 hydrocarbon products. Mater. Res. Bull. 2019. [Google Scholar] [CrossRef]
  250. Beheshti, M.; Kakooei, S.; Ismail, M.C.; Shahrestani, S. Investigation of CO2 electrochemical reduction to syngas on Zn/Ni-based electrocatalysts using the cyclic voltammetry method. Electrochim. Acta 2020. [Google Scholar] [CrossRef]
  251. Debe, M.K. Electrocatalyst approaches and challenges for automotive fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  252. Zinola, C.F.; Martins, M.E.; Tejera, E.P.; Neves, N.P. Electrocatalysis: Fundamentals and applications. Int. J. Electrochem. 2012. [Google Scholar] [CrossRef] [Green Version]
  253. Asadi, M.; Kumar, B.; Behranginia, A.; Rosen, B.A.; Baskin, A.; Repnin, N.; Pisasale, D.; Phillips, P.; Zhu, W.; Haasch, R.; et al. Robust carbon dioxide reduction on molybdenum disulphide edges. Nat. Commun. 2014. [Google Scholar] [CrossRef] [Green Version]
  254. Wang, Y.; Zhang, Z.; Zhang, L.; Luo, Z.; Shen, J.; Lin, H.; Long, J.; Wu, J.C.S.; Fu, X.; Wang, X.; et al. Visible-light driven overall conversion of CO2 and H2O to CH4 and O2 on 3D-SiC@2D-MoS2 Heterostructure. J. Am. Chem. Soc. 2018. [Google Scholar] [CrossRef] [PubMed]
  255. Asadi, M.; Kim, K.; Liu, C.; Addepalli, A.V.; Abbasi, P.; Yasaei, P.; Phillips, P.; Behranginia, A.; Cerrato, J.M.; Haasch, R.; et al. Nanostructured transition metal dichalcogenide electrocatalysts for CO2 reduction in ionic liquid. Science 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Lee, H.I.; Yu, H.; Rhee, C.K.; Sohn, Y. Electrochemical hydrogen evolution and CO2 reduction over hierarchical MoSxSe2−x hybrid nanostructures. Appl. Surf. Sci. 2019. [Google Scholar] [CrossRef]
  257. Xie, Y.; Li, X.; Wang, Y.; Li, B.; Yang, L.; Zhao, N.; Liu, M.; Wang, X.; Yu, Y.; Liu, J.M. Reaction mechanisms for reduction of CO2 to CO on monolayer MoS2. Appl. Surf. Sci. 2020. [Google Scholar] [CrossRef]
  258. Abbasi, P.; Asadi, M.; Liu, C.; Sharifi-Asl, S.; Sayahpour, B.; Behranginia, A.; Zapol, P.; Shahbazian-Yassar, R.; Curtiss, L.A.; Salehi-Khojin, A. Tailoring the edge structure of molybdenum disulfide toward electrocatalytic reduction of carbon dioxide. ACS Nano 2017. [Google Scholar] [CrossRef] [PubMed]
  259. Asadi, M.; Motevaselian, M.H.; Moradzadeh, A.; Majidi, L.; Esmaeilirad, M.; Sun, T.V.; Liu, C.; Bose, R.; Abbasi, P.; Zapol, P.; et al. Highly efficient solar-driven carbon dioxide reduction on molybdenum disulfide catalyst using choline chloride-based electrolyte. Adv. Energy Mater. 2019. [Google Scholar] [CrossRef]
  260. Liu, X.; Yang, H.; He, J.; Liu, H.; Song, L.; Li, L.; Luo, J. Highly active, durable ultrathin MoTe2 layers for the electroreduction of CO2 to CH4. Small 2018. [Google Scholar] [CrossRef]
  261. Primo, A.; He, J.; Jurca, B.; Cojocaru, B.; Bucur, C.; Parvulescu, V.I.; Garcia, H. CO2 methanation catalyzed by oriented MoS2 nanoplatelets supported on few layers graphene. Appl. Catal. B Environ. 2019. [Google Scholar] [CrossRef]
  262. Jung, H.; Cho, K.M.; Kim, K.H.; Yoo, H.W.; Al-Saggaf, A.; Gereige, I.; Jung, H.T. Highly efficient and stable CO2 reduction photocatalyst with a hierarchical structure of mesoporous TiO2 on 3D graphene with few-layered MoS2. ACS Sustain. Chem. Eng. 2018. [Google Scholar] [CrossRef]
  263. Qin, H.; Guo, R.T.; Liu, X.Y.; Pan, W.G.; Wang, Z.Y.; Shi, X.; Tang, J.Y.; Huang, C.Y. Z-scheme MoS2/g-C3N4 heterojunction for efficient visible light photocatalytic CO2 reduction. Dalt. Trans. 2018. [Google Scholar] [CrossRef]
  264. Jia, P.y.; Guo, R.t.; Pan, W.g.; Huang, C.y.; Tang, J.y.; Liu, X.y.; Qin, H.; Xu, Q.Y. The MoS2/TiO2 heterojunction composites with enhanced activity for CO2 photocatalytic reduction under visible light irradiation. Colloids Surfaces A Physicochem. Eng. Asp. 2019. [Google Scholar] [CrossRef]
  265. Kanan, M.W.; Nocera, D.G. In situ formation of an oxygen-evolving catalyst in neutral water containing phosphate and Co2+. Science 2008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Reece, S.Y.; Hamel, J.A.; Sung, K.; Jarvi, T.D.; Esswein, A.J.; Pijpers, J.J.H.; Nocera, D.G. Wireless solar water splitting using silicon-based semiconductors and earth-abundant catalysts. Science 2011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Wang, Q.; Lei, Y.; Wang, D.; Li, Y. Defect engineering in earth-abundant electrocatalysts for CO2 and N2 reduction. Energy Environ. Sci. 2019, 12, 1730–1750. [Google Scholar] [CrossRef]
  268. Ji, Y.; Nørskov, J.K.; Chan, K. Scaling relations on basal plane vacancies of transition metal dichalcogenides for CO2 reduction. J. Phys. Chem. C 2019. [Google Scholar] [CrossRef] [Green Version]
  269. Pérez-Fortes, M.; Schöneberger, J.C.; Boulamanti, A.; Tzimas, E. Methanol synthesis using captured CO2 as raw material: Techno-economic and environmental assessment. Appl. Energy 2016. [Google Scholar] [CrossRef]
  270. Pérez-Fortes, M.; Bocin-Dumitriu, A.; Tzimas, E. CO2 utilization pathways: Techno-economic assessment and market opportunities. Energy Procedia 2014, 63, 7968–7975. [Google Scholar] [CrossRef]
  271. Zheng, Y.; Yin, X.; Jiang, Y.; Bai, J.; Tang, Y.; Shen, Y.; Zhang, M. Nano Ag-decorated MoS2 nanosheets from 1T to 2H phase conversion for photocatalytically reducing CO2 to methanol. Energy Technol. 2019. [Google Scholar] [CrossRef]
  272. Bolivar, H.; Izquierdo, S.; Tremont, R.; Cabrera, C.R. Methanol oxidation at Pt/MoOx/MoSe2 thin film electrodes prepared with exfoliated MoSe2. J. Appl. Electrochem. 2003. [Google Scholar] [CrossRef]
  273. Tu, W.; Li, Y.; Kuai, L.; Zhou, Y.; Xu, Q.; Li, H.; Wang, X.; Xiao, M.; Zou, Z. Construction of unique two-dimensional MoS2-TiO2 hybrid nanojunctions: MoS2 as a promising cost-effective cocatalyst toward improved photocatalytic reduction of CO2 to methanol. Nanoscale 2017. [Google Scholar] [CrossRef]
  274. Xu, F.; Zhu, B.; Cheng, B.; Yu, J.; Xu, J. 1D/2D TiO2/MoS2 hybrid nanostructures for enhanced photocatalytic CO2 reduction. Adv. Opt. Mater. 2018. [Google Scholar] [CrossRef] [Green Version]
  275. Francis, S.A.; Velazquez, J.M.; Ferrer, I.M.; Torelli, D.A.; Guevarra, D.; McDowell, M.T.; Sun, K.; Zhou, X.; Saadi, F.H.; John, J.; et al. Reduction of aqueous CO2 to 1-propanol at MoS2 electrodes. Chem. Mater. 2018. [Google Scholar] [CrossRef] [Green Version]
  276. Dai, W.; Yu, J.; Luo, S.; Hu, X.; Yang, L.; Zhang, S.; Li, B.; Luo, X.; Zou, J. WS2 quantum dots seeding in Bi2S3 nanotubes: A novel Vis-NIR light sensitive photocatalyst with low-resistance junction interface for CO2 reduction. Chem. Eng. J. 2020. [Google Scholar] [CrossRef]
  277. Dai, W.; Yu, J.; Deng, Y.; Hu, X.; Wang, T.; Luo, X. Facile synthesis of MoS2/Bi2 WO6 nanocomposites for enhanced CO2 photoreduction activity under visible light irradiation. Appl. Surf. Sci. 2017. [Google Scholar] [CrossRef]
  278. Peng, H.; Lu, J.; Wu, C.; Yang, Z.; Chen, H.; Song, W.; Li, P.; Yin, H. Co-doped MoS2 NPs with matched energy band and low overpotential high efficiently convert CO2 to methanol. Appl. Surf. Sci. 2015. [Google Scholar] [CrossRef]
  279. Biswas, M.R.U.D.; Ali, A.; Cho, K.Y.; Oh, W.C. Novel synthesis of WSe2-Graphene-TiO2 ternary nanocomposite via ultrasonic technics for high photocatalytic reduction of CO2 into CH3OH. Ultrason. Sonochem. 2018. [Google Scholar] [CrossRef] [PubMed]
  280. Reyes Valle, C.; Villanueva Perales, A.L.; Vidal-Barrero, F.; Gómez-Barea, A. Techno-economic assessment of biomass-to-ethanol by indirect fluidized bed gasification: Impact of reforming technologies and comparison with entrained flow gasification. Appl. Energy 2013. [Google Scholar] [CrossRef]
  281. Villanueva Perales, A.L.; Reyes Valle, C.; Ollero, P.; Gómez-Barea, A. Technoeconomic assessment of ethanol production via thermochemical conversion of biomass by entrained flow gasification. Energy 2011. [Google Scholar] [CrossRef]
  282. Phillips, S.D. Technoeconomic analysis of a lignocellulosic biomass indirect gasification process to make ethanol via mixed alcohols synthesis. Ind. Eng. Chem. Res. 2007, 46, 8887–8897. [Google Scholar] [CrossRef]
  283. Kwok, K.M.; Ong, S.W.D.; Chen, L.; Zeng, H.C. Constrained growth of MoS2 nanosheets within a mesoporous silica shell and its effects on defect sites and catalyst stability for H2S decomposition. ACS Catal. 2018. [Google Scholar] [CrossRef]
  284. Xi, J.; Zhao, T.; Wang, D.; Shuai, Z. Tunable electronic properties of two-dimensional transition metal dichalcogenide alloys: A first-principles prediction. J. Phys. Chem. Lett. 2014. [Google Scholar] [CrossRef]
  285. Ahmed, S.; Yi, J. Two-dimensional transition metal dichalcogenides and their charge carrier mobilities in field-effect transistors. Nano Micro Lett. 2017. [Google Scholar] [CrossRef]
  286. Cates, E.L. Photocatalytic water treatment: So where are we going with this? Environ. Sci. Technol. 2017, 51, 757–758. [Google Scholar] [CrossRef]
  287. Hodges, B.C.; Cates, E.L.; Kim, J.H. Challenges and prospects of advanced oxidation water treatment processes using catalytic nanomaterials. Nat. Nanotechnol. 2018. [Google Scholar] [CrossRef]
  288. Chen, X.; Liu, C.; Mao, S. Environmental analysis with 2D transition-metal dichalcogenide-based field-effect transistors. Nano Micro Lett. 2020. [Google Scholar] [CrossRef] [Green Version]
  289. Ngan, S.L.; How, B.S.; Teng, S.Y.; Leong, W.D.; Loy, A.C.M.; Yatim, P.; Promentilla, M.A.B.; Lam, H.L. A hybrid approach to prioritize risk mitigation strategies for biomass polygeneration systems. Renew. Sustain. Energy Rev. 2020. [Google Scholar] [CrossRef]
  290. Vondra, M.; Touš, M.; Teng, S.Y. Digestate evaporation treatment in biogas plants: A Techno-economic assessment by Monte Carlo, neural networks and decision trees. J. Clean. Prod. 2019, 117870. [Google Scholar] [CrossRef] [Green Version]
  291. Teng, S.Y.; How, B.S.; Leong, W.D.; Teoh, J.H.; Siang Cheah, A.C.; Motavasel, Z.; Lam, H.L. Principal component analysis-aided statistical process optimisation (PASPO) for process improvement in industrial refineries. J. Clean. Prod. 2019, 225, 359–375. [Google Scholar] [CrossRef] [Green Version]
  292. Xu, C.; Gao, S.; Li, M. A novel PCA-based microstructure descriptor for heterogeneous material design. Comput. Mater. Sci. 2017. [Google Scholar] [CrossRef]
  293. Teng, S.Y.; Loy, A.C.M.; Leong, W.D.; How, B.S.; Chin, B.L.F.; Máša, V. Catalytic thermal degradation of Chlorella vulgaris: Evolving deep neural networks for optimization. Bioresour. Technol. 2019, 292, 121971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Ge, L.; Yuan, H.; Min, Y.; Li, L.; Chen, S.; Xu, L.; Goddard, W.A. Predicted optimal bifunctional electrocatalysts for the hydrogen evolution reaction and the oxygen evolution reaction using chalcogenide heterostructures based on machine learning analysis of in silico quantum mechanics based high throughput screening. J. Phys. Chem. Lett. 2020. [Google Scholar] [CrossRef] [PubMed]
  295. Zhao, H.; Hu, R.; Xie, X.; Wang, Z.; Jiang, P.; Zheng, C.; Gao, X.; Wu, T. Adopting big data to accelerate discovery of 2D TMDCs materials via CVR method for the potential application in urban airborne Hg0 sensor. Energy Procedia 2018, 152, 847–852. [Google Scholar] [CrossRef]
  296. Bassman, L.; Rajak, P.; Kalia, R.K.; Nakano, A.; Sha, F.; Aykol, M.; Huck, P.; Persson, K.; Sun, J.; Singh, D.J.; et al. Efficient discovery of optimal N-layered TMDC hetero-structures. MRS Adv. 2018, 3, 397–402. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The number of research articles that applied transition metal dichalcogenides (TMDCs) in gas adsorption application between the year 2015 to April 2020, where pie charts represent the utilization percentage of various types of TMDC. Queried from Scopus Database [48]. Copyright Elsevier B.V., 2020.
Figure 1. The number of research articles that applied transition metal dichalcogenides (TMDCs) in gas adsorption application between the year 2015 to April 2020, where pie charts represent the utilization percentage of various types of TMDC. Queried from Scopus Database [48]. Copyright Elsevier B.V., 2020.
Nanomaterials 10 01012 g001
Figure 2. Schematic diagrams for the adsorption processes of (a) CO, (b) H2O, (c) CO2, (d) sulphur content, and (e) NOx with the use of TMDCs.
Figure 2. Schematic diagrams for the adsorption processes of (a) CO, (b) H2O, (c) CO2, (d) sulphur content, and (e) NOx with the use of TMDCs.
Nanomaterials 10 01012 g002
Figure 3. (a) Mechanism of electron-accepting NO2 gas molecule on MoS2 surface and (b) mechanism of electron-donating NH3 gas molecule on MoS2 surface (Reproduced with permission from [113]. Copyright 2015 Springer Nature Limited CC BY-NC-ND 4.0).
Figure 3. (a) Mechanism of electron-accepting NO2 gas molecule on MoS2 surface and (b) mechanism of electron-donating NH3 gas molecule on MoS2 surface (Reproduced with permission from [113]. Copyright 2015 Springer Nature Limited CC BY-NC-ND 4.0).
Nanomaterials 10 01012 g003
Figure 4. (a) Gas sensing mechanism for volatile organic compounds (VOC); toluene, hexane, ethanol, propanal, and acetone) gas molecule on primitive MoS2 surface and (b) improved VOC sensing performance by Mercaptoundecanoic acid (MUA)-conjugated MoS2 due to increased charge carrier density (Reproduced with permission from [160]. Copyright 2014 American Chemical Society).
Figure 4. (a) Gas sensing mechanism for volatile organic compounds (VOC); toluene, hexane, ethanol, propanal, and acetone) gas molecule on primitive MoS2 surface and (b) improved VOC sensing performance by Mercaptoundecanoic acid (MUA)-conjugated MoS2 due to increased charge carrier density (Reproduced with permission from [160]. Copyright 2014 American Chemical Society).
Nanomaterials 10 01012 g004
Figure 5. (a) Direction-sensitive borophene/MoS2 heterostructure for multiple climate gas detection. Sensitivity and directional bias voltage of (b) CO, (c) NO, (d) NO2, and (e) NH3 in the horizontal and vertical direction (Reproduced with permission from [168]. Copyright 2020 Elsevier B.V.).
Figure 5. (a) Direction-sensitive borophene/MoS2 heterostructure for multiple climate gas detection. Sensitivity and directional bias voltage of (b) CO, (c) NO, (d) NO2, and (e) NH3 in the horizontal and vertical direction (Reproduced with permission from [168]. Copyright 2020 Elsevier B.V.).
Nanomaterials 10 01012 g005
Figure 6. (a) Facile oxidation strategy to prepare a flower-like WS2/WO3 heterostructure for adsorbing RhB from wastewater (Reproduced with permission from [182]. Copyright 2020 MDPI AG CC-BY 4.0). (b) Bifunctional cetyltrimethylammonium bromide (CTABr)-coated MoS2-decorated Fe3O4 nanoparticles using pH to alter adsorption properties between Cr(VII) and Cr(III) in wastewater (Reproduced with permission from [189]. Copyright 2017 American Chemical Society CC-BY-NC-ND Under ACS AuthorChoice). (c) A superhydrophobic/superhydrophilic MF@MoS2 sponge for water/oil separation (Reproduced with permission from [190]. Copyright 2017 Elsevier B.V).
Figure 6. (a) Facile oxidation strategy to prepare a flower-like WS2/WO3 heterostructure for adsorbing RhB from wastewater (Reproduced with permission from [182]. Copyright 2020 MDPI AG CC-BY 4.0). (b) Bifunctional cetyltrimethylammonium bromide (CTABr)-coated MoS2-decorated Fe3O4 nanoparticles using pH to alter adsorption properties between Cr(VII) and Cr(III) in wastewater (Reproduced with permission from [189]. Copyright 2017 American Chemical Society CC-BY-NC-ND Under ACS AuthorChoice). (c) A superhydrophobic/superhydrophilic MF@MoS2 sponge for water/oil separation (Reproduced with permission from [190]. Copyright 2017 Elsevier B.V).
Nanomaterials 10 01012 g006
Figure 7. (a) Single-layer MoS2 water desalination membrane with Mo only pores (Reproduced with permission from [193]. Copyright 2015 Springer Nature Limited CC BY 4.0). (b) MoS2/GO nanosheets intercalated membrane for removal of dyes and salts (Reproduced with permission from [196]. Copyright 2020 Elsevier B.V.).
Figure 7. (a) Single-layer MoS2 water desalination membrane with Mo only pores (Reproduced with permission from [193]. Copyright 2015 Springer Nature Limited CC BY 4.0). (b) MoS2/GO nanosheets intercalated membrane for removal of dyes and salts (Reproduced with permission from [196]. Copyright 2020 Elsevier B.V.).
Nanomaterials 10 01012 g007
Figure 8. Mechanism of a MoSe2/SrTiO3 photocatalyst to oxidize and degrade methyl orange dyes in wastewater. UV irradiation causes electron transfer from the valence band (VB) to the conduction band (CB), which activates the separation process (Reproduced with permission from [200]. Copyright 2018 World Scientific Publishing Co Pte Ltd CC BY Under Open Access).
Figure 8. Mechanism of a MoSe2/SrTiO3 photocatalyst to oxidize and degrade methyl orange dyes in wastewater. UV irradiation causes electron transfer from the valence band (VB) to the conduction band (CB), which activates the separation process (Reproduced with permission from [200]. Copyright 2018 World Scientific Publishing Co Pte Ltd CC BY Under Open Access).
Nanomaterials 10 01012 g008
Figure 9. (a) Hydrodeoxygenation (HDO) of lignin-derived phenols into arenes using Co-MoS2−x (Reproduced with permission from [225]. Copyright 2020 American Chemical Society) and (b) deoxygenation of palm kernel oil to jet fuel-like hydrocarbon using catalysts (Reproduced with permission [206]. Copyright 2017 Elsevier B.V.).
Figure 9. (a) Hydrodeoxygenation (HDO) of lignin-derived phenols into arenes using Co-MoS2−x (Reproduced with permission from [225]. Copyright 2020 American Chemical Society) and (b) deoxygenation of palm kernel oil to jet fuel-like hydrocarbon using catalysts (Reproduced with permission [206]. Copyright 2017 Elsevier B.V.).
Nanomaterials 10 01012 g009
Figure 10. (a) Scheme of the electrochemical conversion of CO2 into value-added chemicals (Reproduced with permission from [241]. Copyright 2019 CC BY-NC 4.0 John Wiley and Son) and (b) general working principle of photocatalyst for CO2 conversion (Reproduced with permission from [249]. Copyright Elsevier 2019).
Figure 10. (a) Scheme of the electrochemical conversion of CO2 into value-added chemicals (Reproduced with permission from [241]. Copyright 2019 CC BY-NC 4.0 John Wiley and Son) and (b) general working principle of photocatalyst for CO2 conversion (Reproduced with permission from [249]. Copyright Elsevier 2019).
Nanomaterials 10 01012 g010
Figure 11. (a) Photochemical reaction pathway for the conversion of CO2 and H2O to CH4 and O2 under visible light using SiC@MoS2 nanoflower (Reproduced with permission from [254]. Copyright 2018 American Chemical Society) and (b) schematic of the WSe2-based artificial leaf cell design for the conversion of CO2 and H2O to CO and H2 (Reproduced with permission from [255]. Copyright 2016 The American Association for the Advancement of Science).
Figure 11. (a) Photochemical reaction pathway for the conversion of CO2 and H2O to CH4 and O2 under visible light using SiC@MoS2 nanoflower (Reproduced with permission from [254]. Copyright 2018 American Chemical Society) and (b) schematic of the WSe2-based artificial leaf cell design for the conversion of CO2 and H2O to CO and H2 (Reproduced with permission from [255]. Copyright 2016 The American Association for the Advancement of Science).
Nanomaterials 10 01012 g011
Table 1. NO2 sensing TMDC materials and their performances.
Table 1. NO2 sensing TMDC materials and their performances.
MaterialTarget GasLODConditionRemarks
MoS2/graphene hybrid aerogel [120]NO250 ppbRT to 200 °CResponse and recovery time < 1 min.
WS2 with AgNW functionalization [128]NO225 ppm100 °C667% response compared to pristine 4L WS2 sensor.
MoS2 [121]NO220 ppbRTSensitivity of 194%/ppm
MoS2 [129]NO2120 ppbRT, 100 °CPeak to valley sensitivity > 50%.
MoS2 [115]NO225 ppbRT (with red light)Sensitivity of 3300%/ppm.
rGO/MoS2 nanocomposite [123]NO25.7 ppb60 °COver 55% sensing response for NO2 at 8 ppm
ZnO/MoS2 [130]NO2200 ppbRT (with monochromatic light)Sensitivity at 29.3%/ppm, response time of 4.3 min, recovery time of 1.2 min.
SnS2/MoS2 [114]NO225.9 ppmRTResponse time of 2 s and recovery time of 28.2 s.
MoSe2 [131]NO2300 ppmRTResponse time of 20 min and recovery time of 30 min.
MoSe2 [116]NO210 ppmRT (with UV light)Response time <200 s.
WS2/WO3 composite film [132]NO2100 ppb150 °CResponse time of 70 s and recover time of 120 s.
WS2/graphene aerogel [122]NO210-15 ppbRTResponse time of 70 s, recovery time of 300 s. (2 ppm)
MoTe2 [127]NO220 ppbRT (with UV of 2.5 mW/cm2)Response time of 120 s with response of 18%.
PtSe2 [126]NO20.9 ppbRTResponse time of 1 s and recovery time of 4 s
Table 2. NO sensing TMDC materials and their performances.
Table 2. NO sensing TMDC materials and their performances.
MaterialTarget GasLODConditionRemarks
MoS2 deposited onto Si/SiO2 [137]NO0.8 ppmRT80% decreased response at 2 ppm.
MoS2 [138]NO100 ppmRT, 50 °C, 100 °C (with UV light)Response and recovery time below 600 s. Response at 25.63%.
UV-ozone treated MoS2 [139]NO20 ppm125 °CStability issue over 120 ppm. Poor recovery.
WOx/WSe2 hybrid [140]NO0.3 ppbRTResponse time of 250 s, S/N ratio > 10, sensitivity of 520%/ppm
Table 3. Ammonia gas sensing TMDC materials and their performances.
Table 3. Ammonia gas sensing TMDC materials and their performances.
MaterialTarget GasLODConditionRemarks
MoSe2 [149]NH350 ppmRTResponse time of 2.5 min, recovery time of 9 min.
Graphene/MoS2 [146]NH35 ppm150 °CResponse time < 10 min, Recovery time < 30 min.
WS2 [148]NH350 ppmRTResponse time of 200 s, recovery time of 232.3 s.
MoS2 [121]NH31 ppmRTResponse time 5–9 min, recovery time < 15 min.
UV-treated MoS2 [139]NH3100 ppmRTResponse time of 7 min,
Recovery time of 12 min.
MoTe2 [147]NH32 ppmRTOver 95% recovery using gate biases of 0 V and 20 V. Response time 10 min, recovery time 20 min.
MoS2/Co3O4 [150]NH30.1 ppmRTResponse time is 98 s, recovery time is 100 s.
PMMA-MoS2 [151]NH31 ppmRTSensitivity of 54%, response time of 10 s, recovery time of 14 s.
MoS2/VS2 [152]NH35 ppm40 °CRecovery and response time both < 5 min.
Table 4. Application of TMDCs in CO2 conversion to syngas and other gases.
Table 4. Application of TMDCs in CO2 conversion to syngas and other gases.
MechanismCatalyst/MaterialDetailsConditionProduct
Electro-catalyst2D nanoflake WSe2 [255]50 vol% EMIM-BF4 in water−0.764 VCO
Electro-catalystHierarchical MoSxSe(2 − x) hybrid nanostructures [256]0.1 M H2SO4 solution−0.70VH2
Electro-catalyst1 Monolayer MoS2 [257]--CO
Electro-catalyst5% niobium (Nb)-doped
vertically aligned MoS2 [258]
50 vol% EMIM-BF4 in water−0.8 VCO
Electro-catalystmolybdenum disulfide nanoflakes (MoS2 NFs) [259]2.0 M C5H14ClNO−2.0 VCO
Electro-catalystUltrathin MoTe2 [260]0.1 M KHCO3–1.0 VCH4
Photo-catalystMoS2 nanoplatelet supported on few layer graphene [261]Up to 60% conversion, 90% CH4 selectivityReact with H2, 250–500 °CCH4, CO
Photo-catalystMarigold-like Si@MoS2[254]Produced 323 μL∙g−1∙h−1 CH4, 23 μL∙g−1∙h−1 O2React with H2OCH4, O2
Photo-catalystMesoporous TiO2 on 3D Graphene with Few-layered MoS2[262]CO selectivity of 97% and yield of 93.22 µmol/g h-CO
Photo-catalystZ-scheme MoS2/g-C3N4 heterojunction [263]Produced 58.59 μmol∙g−1 in 7 h.25 °C, 100 kPaCO
Photo-catalystMoS2/TiO2 heterojunction [264]Produced 268.97 μmol∙g−1 CO and 49.93 μmol∙g−1 CH425 °C, 100 kPaCO, CH4
1 Computational simulation using density functional theory (DFT).
Table 5. Application of TMDCs in converting CO2 to high-value chemicals.
Table 5. Application of TMDCs in converting CO2 to high-value chemicals.
MechanismCatalyst/MaterialDetailsConditionProduct
Photo-catalystUndecorated 2D-MoS2 [249]Produced 27.4 μmol∙g−1∙h−1 methanol and 2.2 μmol∙g−1∙h−1 acetaldehyde0.5 M NaHCO3Methanol, acetaldehyde
Photo-catalystMoS2/Bi2WO6 nanocomposites [277]Produced 36.7 μmol∙g−1 methanol and 36.6 μmol∙g−1 ethanol (in 4 h)Deionized waterMethanol, ethanol
Photoelectro-catalystCo-doped MoS2 NPs [278]Produced 35 mmol∙L−1 methanol (in 350 min)−0.64 VMethanol
Photo-catalystNano Ag decorated MoS2 nanosheets [271]Produced 365.08 μmol∙g−1∙h−120 mL isopropanol,Methanol
Electro-catalysisMoS2 electrodes [275]0.1 M potassium phosphate buffer−0.59 V1-propanol
Photo-CatalystWSe2/Graphene/TiO2 nanocomposite [279]Produced 6.326 μmol∙g−1∙h−1 methanolDistilled water with Na2SO3 reagentMethanol
Photo-catalystMoS2/TiO2[273]Produced 10.6 μmol∙g−1∙h−1 methanolDistilled Water and methanolMethanol
Photo-catalystWS2 quantum dots doped Bi2S3 nanotubes [276]Produced 38.2 μmol∙g−1∙h−1 methanol and 27.8 μmol∙g−1∙h−1 ethanolUltrapure waterMethanol, ethanol

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/nano10061012

AMA Style

Zhang X, Teng SY, Loy ACM, How BS, Leong WD, Tao X. Transition Metal Dichalcogenides for the Application of Pollution Reduction: A Review. Nanomaterials. 2020; 10(6):1012. https://doi.org/10.3390/nano10061012

Chicago/Turabian Style

Zhang, Xixia, Sin Yong Teng, Adrian Chun Minh Loy, Bing Shen How, Wei Dong Leong, and Xutang Tao. 2020. "Transition Metal Dichalcogenides for the Application of Pollution Reduction: A Review" Nanomaterials 10, no. 6: 1012. https://doi.org/10.3390/nano10061012

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