Praveen P. Singh
a,
Surabhi Sinha
a,
Geetika Pandey
b and
Vishal Srivastava
*c
aDepartment of Chemistry, United College of Engineering & Research, Prayagraj 211002, Uttar Pradesh, India. E-mail: ppsingh23@gmail.com
bDepartment of Physics, United University, Prayagraj 211012, Uttar Pradesh, India
cDepartment of Chemistry, CMP Degree College, University of Allahabad, Prayagraj-211002, Uttar Pradesh, India. E-mail: vishalgreenchem@gmail.com
First published on 18th October 2022
Photoredox catalysis has been explored for chemical reactions by irradiation of photoactive catalysts with visible light, under mild and environmentally benign conditions. Furthermore, this methodology permits the activation of abundant chemicals into valuable products through novel mechanisms that are otherwise inaccessible. In this context, MoS2 has drawn attention due to its excellent solar spectral response and its notable electrical, optical, mechanical and magnetic properties. MoS2 has a number of characteristic properties like tunable band gap, enhanced absorption of visible light, a layered structure, efficient photon electron conversion, good photostability, non-toxic nature and quantum confinement effects that make it an ideal photocatalyst and co-catalyst for chemical transformations. Recently, MoS2 has gained synthetic utility in chemical transformations. In this review, we will discuss MoS2 properties, structure, synthesis techniques, and photochemistry along with modifications of MoS2 to enhance its photocatalytic activity with a focus on its applications and future challenges.
In recent years, MoS2 has received much attention due to its unique electrical and optical properties.8a MoS2 is a two-dimensional (2D) transition metal dichalcogenide, a narrow band semiconductor with a band gap of 1.9 eV which can be altered by adjusting the number of layers. Among all 2D TMDs, MoS2 is one of the few with a natural layered structure, indicating that MoS2 can be stripped to obtain high-quality 2D MoS2 without complicated chemical synthesis.8b Therefore, the cost of preparing 2D MoS2 is much lower than other 2D TMDs due to the simple synthesis conditions. More importantly, 2D MoS2 is a 2D semiconductor with a direct band-gap, which has the best electric performance among 2D TMDs.8c These make 2D MoS2 get more attention among 2D TMDs.
Several characteristics of MoS2 such as high absorbance in the visible region,9 high carrier mobility,10 stability,11 preferable catalytic activity,12 good photostability,13 outstanding electrochemical properties,14 strong surface adsorption capacity and higher specific surface area15 favours its application as a photocatalyst as well as a co-catalyst and also, it is capable of addressing the limitations faced by typical photocatalysts. Additionally, MoS2 is found to be abundant and cheap.7 Currently, semiconductor photocatalysts with controllable morphologies are gaining significant attraction in applications associated with environmental remediation process.16 Phase engineering, defect engineering, doping of external materials, interlayer engineering, surface modification, and heterojunction17a,b construction make MoS2 more flexible in terms of tuning its properties during synthesis.
Molybdenum disulfide (MoS2)17c has emerged as one of the most important two-dimensional functional materials, popular new co-catalysts due to their excellent photocatalytic activity, substantial adsorptivity, great value and its non-toxicity. The visible region of electromagnetic spectrum exhibits the absorption spectrum of 2D MoS2, providing a plethora of opportunities for widespread applications.17d Applications for MoS2-based materials are said to include energy storage,17e hydrogen production,17f pollutant degradation,17g disinfection,17h etc. Numerous studies have recently provided a summary of photocatalytic properties of MoS2 for numerous application areas.17i–n
In continuation of our work on photocatalysed organic synthesis18,19 this review aims to provide a comprehensive report on the current research, especially the role of MoS2 in the chemical transformation.
No. | Synthesis method | Metal precursor used | Source of sulfur used | Morphology and particle size | Applications | Ref. |
---|---|---|---|---|---|---|
1 | Colloidal | Mo(CO)6 | Sulfur powder | Nanosheets | Electrochemical studies | 1 |
2 | Hydrothermal | (NH4)Mo7O24·4H2O | Thiourea (H2CSNH2) | Layered MoS2 nanoflowers with ∼0.1 μm particle size | Photocatalytic degradation of methylene blue and crystal violet dyes | 2 |
3 | Hydrothermal | Sodium molybdate dihydrate (Na2MoO4·2H2O) | Thioacetamide (C2H5NS) | Nano flowers with average size ∼100 nm | Photocatalytic degradation of rhodamine B | 3 |
4 | Hydrothermal | Sodium molybdate dihydrate (Na2MoO4·2H2O) | L-cysteine | Quantum dots with ∼2.5 nm particle size | Detection of methyl parathion (pesticide) | 4 |
5 | Hydrothermal | (NH4)MoS4 | (NH4)MoS4 | Fluorescent probe for hyaluronidase detection | 5 | |
6 | Chemical Exfoliation | Commercially available MoS2 powder | Nanosheets | Photocatalytic oxidation of benzyl halides | 6 | |
30 | ||||||
7 | Chemical vapor Deposition | Ammonium heptamolybdate | Sulfur | Not mentioned | 7 | |
31 | ||||||
8 | Colloidal | Ammonium tetrathiomolybdate [(NH4)2MoS4] | Ammonium tetrathiomolybdate [(NH4)2MoS4] | Spherical quantum dots with average size ∼5 nm | Bioimaging | 8 |
32 | ||||||
9 | Heating | Mo(acac)2 | 1-Dodecanethiol | Nanosheet | Electrical bistability performance | 9 |
33 | ||||||
10 | Hot injection | Molybdenum(V) chloride (MoCl5) | N,N′-diphenylthiourea | Nanosheets | 10 | |
34 | ||||||
11 | Hydrothermal | Sodium molybdate dihydrate (Na2MoO4·2H2O) | Thioacetamide (CH3CSNH2) | Coral-like | Lubrication additive; photocatalytic degradation of liquid paraffin | 11 |
35 | ||||||
12 | Hydrothermal | Na2MoO4·2H2O | Cysteine | Quantum dots with 2 ∼3.5 nm particle size | Fluorescent probe for sensing of hydroquinone and bioimaging | 12 |
36 | ||||||
13 | Solid state | (NH4)6Mo7O24·4H2O | Sulfur | Nanosheets, thinner than 5 nm | Photocatalytic degradation of rhodamine B | 24 |
14 | Hydrothermal | Na2MoO4·2H2O | Thiourea (H2CSNH2) | Irregular with average size in the range 12–25 nm | Electrochemical studies | 25 |
15 | Hydrothermal | Ammonium hepta molybdate tetrahydrate [(NH4)6Mo7O24·4H2O] | Ammonium polysulfide | Ammonium polysulfide as the sulfur source: Uniform MoS2 nanospheres with average size of ∼100 nm | Lubrication additive | 27 |
16 | Hydrothermal | MoO3 | Potassium thiocyanate | Flowerlike MoS2 spheres with average diameter of 1–2 nm | Photocatalytic degradation of methylene blue | 37 |
17 | Hydrothermal | Na2MoO4·2H2O | L-cysteine | Microspheres comprising crossed-linked nanorods ∼100 nm in length | Photocatalytic degradation of thiocarbamate herbicides | 38 |
18 | Hydrothermal | Ammonium tetrathiomolybdate [(NH4)2MoS4] | Thiourea (H2CSNH2) | Flowerlike microsphere | Photocatalytic degradation of rhodamine B and methylene blue | 39 |
19 | Hydrothermal | (NH4)6Mo7O24·4H2O | Thiourea | Nanosheets | Detection of dopamine | 40 |
20 | Hydrothermal | (NH4)6Mo7O24·4H2O | Na2S | Hierarchical porous with the thickness of ∼20–40 nm | Detection of phenol sulphite oxidase, nicotinamide adenine dinucleotide oxidase and superoxide dismutase mimicking activities | 41 |
21 | Microwave | Ammonium tetrathiomolybdate [(NH4)2MoS4] | Ammonium tetrathiomolybdate [(NH4)2MoS4] | Quantum dots with average diameter of ∼1.72 nm | Determination of terramycin | 42 |
22 | Solid state | (NH4)2MoS4 | Thiourea | Sheetlike structure and ultrathin layers | Electrochemical measurement | 43 |
The benefits and drawbacks of synthesis techniques44 such as thermal annealing, plasma-assisted synthesis, magnetron sputtering-based synthesis, atomic layer deposition (ALD), and wet chemical processes are summarised in Fig. 1.
Fig. 1 MoS2 synthesis techniques with their advantages and disadvantages.44 |
Fig. 2 (a). Visualization of the layered structure of MoS2 and the forces responsible for its layered structure. (b). Pictorial illustration of the increased interlayer distance due to the presence of impurities. Adapted with permission from ref. 47, Copyrights 2022 Elsevier. (c). Graphical illustration of the tunable band gap exhibited by MoS2. (d). Crystal structures of MoS2. (d1). 2H, (d2). 3R and (d3).1T. Adapted with permission from ref. 11, Copyrights 2022 Elsevier. (e). Different morphologies portrayed by MoS2 (through synthesis) (e1). Nanoflowers at the resolution of 2 μm (e2). Nanoflowers at the resolution of 500 nm Adapted with permission from ref. 48, Copyrights 2022 Elsevier (e3). TEM image of the nanosheets at 10 nm (e4). TEM image of the nanosheets at 200 nm Adapted with permission from ref. 49, Copyrights 2022 Elsevier (e5). Nanotubes Adapted with permission from ref. 53, Copyrights 2022 Elsevier (e6). Nanoworms. Adapted with permission from ref. 54, Copyrights 2022 Elsevier (e7). Coral-like structure. Adapted with permission from ref. 49 Copyrights 2022 Elsevier (e8). Quantum dots. Reproduced from ref. 55 with permission from [Elsevier], Copyright [2022]. |
Fig. 3 Schematic illustration of the photocatalytic mechanism of MoS2. Reproduced from ref. 29 with permission from [American Chemical Society], copyright [2022]. |
Crystal lattices of MoS2 exists in four polymorphic forms namely 1H, 2H, 1T and 3R (H- Hexagonal, T-Tetragonal and R-Rhombohedral) which are classified based on the stacking arrangement and the co- ordination between Mo and S atoms (Fig. 2d).11 MoS2 is synthesized in various morphologies (Fig. 2e) which includes nanoflowers (Fig. 2e1, Fig. 2e2),48 nanosheets (Fig. 2e3, Fig. 2e4),49 nanoflakes,50 nanocrystals, nanospheres,51 nanoribbons,52 nanotubes (Fig. 2e5),53 nanoworms (Fig. 2e6),54 coral-like structures (Fig. 2e7),49 nanodots and quantum dots (Fig. 2e8).55
Due to the 2D nature of MoS2, it's intrinsic properties are greatly influenced by the surrounding environment on the nanoscale. The suspended structures in these systems are hybrid systems with transport performance,56a–g tunable optoelectronic behaviour,56h–l and nanomechanical properties,56m–p all of which are fascinating on their own. The size, surface area, and surface energy of the MoS2 particles are determined by its morphology, which has a significant impact on how it executes in photocatalytic applications. Due to their large surface area and narrow band gap, MoS2 nanoflowers are more effective for absorbtion of visible light, which improves their photocatalytic efficacy. The shape of MoS2 nanoflowers determines their surface area, which in turn determines how well they operate as photocatalysts. MoS2 is anisotropic, piezocatalytic,56 hydrophobic and low-toxic, which has high specific surface area with high surface energy,57 high surface to volume ratio, high e− mobility, high optical absorption, high catalytic activity, good tribological properties,58 good thermal stability and free sulfur groups.59 Due to its low price and a wealth of earth reserves, it is commercially viable. As a result, MoS2's characteristics are largely in favour of its application as a photocatalyst.
Unfortunately, the catalytic activity of MoS2 is greatly inhibited due to its low dispersion ability which can be attributed to its hydrophobicity and low electrical conductivity.60 The formation of composites using MoS2 may be a viable and efficient way to get around this problem and maximise the benefits of MoS2 as a photocatalyst in aqueous bodies because the properties of other compounds in the composite may enhance the electrical conductivity and dispersion of MoS2 and thereby facilitate the exhibition of its full potential and surplus advantages.
The photocatalytic redox processes begin with the production of photoexcited charge carrier pairs (e−/h+). As illustrated in Fig. 3, upon irradiation the MoS2 simultaneously generates e− and h+, where the photogenerated e− is promoted to the CB, leaving behind the hole (h+) in the VB. Subsequently, these photogenerated carriers migrate to the surface of the MoS2 to participate in reduction (e−) and oxidation (h+) processes. Photoexcited e− may react with adsorbed O2 to produce ˙O2− radicals and, simultaneously, the h+ remaining in the VB can abstract an e− from hydroxyl ions or adsorbed H2O molecules to produce ˙OH radicals, which are powerful oxidizing agents that can react with harmful organic, inorganic, and biological compounds.
Fig. 4 Modifications of MoS2 to improve its photocatalytic activity; (a) metal doping of MoS2, (b) nonmetal doping of MoS2, (c) metal deposited on MoS2, and (d) formation of a heterojunction with a second semiconductor. Reproduced from ref. 29 with permission from [American Chemical Society], copyright [2022]. |
A hydrothermal and calcination technique is used to successfully create a Z-scheme MoS2/CuO photocatalyst that degrades 2-mercaptobenzothiazole (MBT) with excellent activity (96 percent) when exposed to visible light68 (Fig. 5a–e).
Fig. 5 (a) SEM images of CuO, (b) MoS2, (c) MoS2/CuO, (d) TEM image of MoS2/CuO, (e) elemental mapping of MoS2/CuO, (f) photocatalytic degradation mechanism of MBT over the Z-scheme MoS2/CuO heterojunction under visible light irradiation. Reproduced from ref. 68 with permission from [Royal Society of Chemistry], copyright [2020]. |
MoS2 nanosheets and flower-shaped CuO combine to form a Z-scheme heterostructure, which significantly boosts the separation effectiveness of photogenerated carriers. CuO and MoS2's oxidation and reduction characteristics are enhanced by the Z-scheme electron transfer process that gives significant accumulation of photogenerated electrons and holes.
The authors proposed a possible Z-scheme charge transfer process and the reaction mechanism of MBT degradation by MoS2/CuO heterojunction under visible light irradiation and the results are shown in Fig. 5f. Electron–hole pairs are generated by the excitation of both MoS2 and CuO. The electrons on the CB of CuO migrate to the VB of MoS2 and combine with holes on the VB of MoS2. The photogenerated holes assembled by CuO can degrade MBT molecules in large quantities. In addition, due to the CB position of MoS2 is much more negative than the potential of O2/O2, the electrons on the CB of MoS2 will further undergo a reduction reaction to generate abundant O2, which will be further involved in the degradation process. Although CuO has a higher negative VB potential than OH, it is theoretically impossible to form OH; yet, some OH has been found using the ESR technique, which may be the result of further reducing O2. The Z-scheme charge transfer mechanism for MoS2/CuO significantly increases the separation efficiency of photogenerated carriers and positively influences the oxidizability and reducibility of the photocatalysts. Finally, the detailed photocatalytic reaction process of the Z-scheme MoS2/CuO heterojunction is in Fig. 5f.
A S-scheme MoS2/g-C3N4 photocatalyst is developed69 a one-pot solid-state reaction of thiourea and sodium molybdate as precursors at different temperatures under N2 gas was applied. The variation in component contents (MoS2 and gC3N4) is determined by the physicochemical characterization of the final products via an increase in synthesis temperature. The morphology and elemental composition of representatives were investigated using field emission scanning electron microscopy (FE-SEM), high resolution transmission electron microscopy (HR-TEM), along with energy dispersive X-ray spectroscopy (EDS) mapping (Fig. 6a–g).
Fig. 6 (a and b) FE-SEM; and (c and d) HR-TEM images of MCN-500; FE-SEM images of (e) MCN-600; and (f) CN-500; EDS mapping images of (g) mapping area, (g1) carbon, (g2) nitrogen, (g3) sulfur, and (g4) molybdenum, (h) schematic diagram of band alignment of heterojunction and S-scheme charge transfer on interface of MoS2 and g-C3N4. Reproduced from ref. 69 with permission from [Springer Nature], copyright [2021]. |
The degradation of Rhodamine B in an aqueous solution under visible light was used to evaluate the enhanced photocatalytic activity of the MoS2/gC3N4 composites. The best photocatalytic performance there was demonstrated by composites created at 500 °C, with a degradation efficiency of 90%, which was significantly better than that of a single g-C3N4. The improvement in light harvesting and the extension in the lifetime of photoinduced charge carriers of composites, which are the result of the synergistic interaction between the components, are credited with the significant improvement in photocatalytic performance. Besides, the photocatalytic mechanism is demonstrated to well-fit into the S-scheme pathway with apparent evidences (Fig. 6h).
S. N. | Application | Photocatalyst used | Morphology | Particle size | Light source | Performance | Ref. |
---|---|---|---|---|---|---|---|
1 | Photocatalytic oxidation of benzyl alcohol to benzaldehyde | Co-doped MoS2/g-C3N4 | 2D nanosheet morphology with curly stripes | 80 W LED lamp | Benzaldehyde production rate of 0.48 mmol g−1 h−1 | 66 | |
2 | Photocatalytic oxidative coupling of thiols | Pd@Cu/MoS2 | Spherical | 64.5 nm | 300 W Xe lamp | ∼99% conversion under 400–800 nm irradiation | 70 |
3 | Photocatalytic reduction of 4-nitrophenol | TiO2 hollow spheres/crumpled MoS2 nanosheet | Hollow sphere | ∼200 nm | 500 W Xe lamp | 99.35% photocatalytic reduction of 4-nitro- phenol | 71 |
4 | Photocatalytic conversion of CO2 to methane | MoS2/Cu | MoS2 nanosheets are coated on the surface of Cu nanorods | About 50–700 nm | 300 W Xe lamp | Maximum yield of methane ∼23 mmol g−1 h−1 | 72 |
5 | Photocatalytic reduction of CO2 to methanol | MoS2 grown on hexagonal boron nitride nanoplatelets | MoS2 nanosheets are uniformly grown over the hexagonal boron nitride nanoplatelets | Each MoS2 nanosheet is composed of 2–6 molecular lamellae | 20 W white LED lamp | Maximum yield of methanol 5994 μmol g−1 | 73 |
6 | Photocatalytic reduction of CO2 to methane and CO | In2S3/MoO3@MoS2 | Distorted hexagonal nanorods | 300 W Xe lamp | Yield ∼29.4 and ∼35.6 μmol g−1 h−1 for CH4 and CO, respectively | 74 | |
7 | Photocatalytic selective oxidation of benzyl alcohols to benzaldehyde | Ag3PO4 nanoparticle@MoS2 quantum dot/few-layered MoS2 nanosheet | Nanosheet | 300 W Xe lamp | ≤92% conversion of benzyl alcohol and ∼87% yield of benzaldehyde after 3 h of irradiation | 75 | |
8 | Photocatalytic reduction of 4- nitrophenol to 4-aminophenol | CdS-MoS2/rGO composite | Flowerlike morphology | 500 W Xe lamp | ≤70% reduction of 4-nitrophenol after 60 min of irradiation | 76 |
Fig. 7 (a) Photomediated oxidation and reduction cycle of MoS2 QDs. (b) Proposed mechanism for the catalytic activity of MoS2 blue QDs in CDC reaction synchronized with HER. Reproduced from ref. 77a with permission from [American Chemical Society], copyright [2022]. |
The authors proposed a mechanistic pathway for the CDC reaction parallel to the photochemical HER mediated by MoS2 QDs (Fig. 7). An electron is excited from the valence band (VB) to the conduction band (CB) by radiation. In presence of O2, 2-Aryl-1,2,3,4-tetrahydroisoquinoline 1 is oxidized to radical cation 1+˙ via transferring an electron to the VB to regenerate the ground state of the QD. There is no need of any external photosensitizer for the oxidation of 1. The radical cation of 1 results the formation of iminium intermediate 1+ by losing a H˙. The nucleophile 2 attacks the iminium intermediate 1+ to give the targeted cross-coupled product 3 with simultaneous evolution of H2. Thus, QDs act as a photosensitizer as well as a catalyst that also transfer electron for the water-splitting process and further regenerated in the process. In the proposed mechanism, they have stated that the source of H can be from H2O as well as the nucleophile NuH. As their target was to design a methodology for the CDC reaction, they did not carry out any deuterated experiments to confirm the source of H2.
De et al. also demonstrated77b 1T (metallic)-2H (semiconducting) phase boundaries intrinsic to individual sheets of chemically exfoliated 2D-MoS2 that can serve as heterojunctions for enhanced photocatalysis compare to only semiconducting phase. Due to the abundance of heterojunctions in these multiphasic materials, chemically exfoliated 2D-MoS2 provides improved stability as well as transfer of photogenerated charges to the reactants, giving better yield. Significantly, they proved that this simple synthesize material works well as a photocatalyst for the aerobic oxidative coupling of amines to imines when subjected to visible light. Given its broad applications, they believe mixed phase 2D-MoS2 can be of interest to several industrial synthetic applications related to semiconductor-based photocatalysis. An additional benefit is that this heterogeneous photocatalyst can be regenerated up to five times without significant loss in the activity (Fig. 8).
Fig. 8 Plausible mechanism for the oxidative coupling of benzylamine to imine photocatalyzed by mixed-phase 2D-MoS2 nanosheets. Reproduced from ref. 77b with permission from [John Wiley & Sons, Inc.], copyright [2018]. |
Further De et al.77c have developed an efficient method for Cross Dehydrogenative Coupling (CDC) reaction by using mixed phase 2D-MoS2 nanosheets and Eosin Y alongside the photochemical Hydrogen Evolution Reaction (HER). The reaction has been carried out at room temperature under visible light irradiation and results in good to excellent yields. It is an example of a CDC reaction using mixed phase 2D-MoS2 nanosheets as a catalyst regenerator and Eosin Y as photosensitizer. Due to incomplete recovery of the catalyst, the material can be reused with slight loss in the yields. Additionally, the scope of this protocol can be used to create various C–C coupled products. This could be an illustration of how to completely utilise the oxidation and reduction cycles of several semiconductor-based photocatalysts. (Fig. 9).
Fig. 9 Schematic illustration of the 2D-MoS2 photocatalyzed oxidation and reduction cycles in presence of Eosin Y as photosensitizer. Reduction cycle produces H2 from water as reported earlier, whereas oxidative cycle is responsible for CDC Reaction and formation of additional H2. Reproduced from ref. 77c with permission from [Royal Society of Chemistry], copyright [2019]. |
Fig. 10 The Pd-nanodot decorated MoS2 micro/nanosheet for an efficient visible light induced photocatalytic Suzuki–Miyaura coupling reaction. Reproduced from ref. 78 with permission from [Royal Society of Chemistry], copyright [2017]. |
Fig. 11 Preparation of MoS2/CZ300 photocatalyst and the suggested charge transfer mechanism of MoS2/CZ300 composite (a) under simulated sunlight (b) and visible light irradiation. Reproduced from ref. 79 with permission from [American Chemical Society], copyright [2018]. |
Several methods were employed to characterise the produced composite, and the results displayed that carbon was loaded onto the surface of ZnO instead of doped into the lattice in order to prevent charge carrier recombination. By thermal processing, the carbon content of C–ZnO can be altered by which enhances the effectiveness of charge separation. For calcination, the most favourable temperature of 300 °C determined. The separation of electron–hole pairs can be further improved by photodepositing MoS2 nanoparticles on the C–ZnO-300 sample, which traps electrons on the carbon layer. The ideal 1% MoS2/C-ZnO-300 composite exhibits the fastest NH3 generation rate under simulated sunlight irradiation at 245.7 μmol L−1 g−1 h−1, which is 9.3 and 4.0 times higher than that of ZnO and C–ZnO, respectively. However, when exposed to visible light, C–ZnO performs best, producing NH3 at a rate of 28.8 μmol L−1 g−1 h−1, indicating that the composite employs a different process. The carbon layer was thought to act as a photosensitizer by transferring electrons to ZnO or MoS2. The processes proposed have been confirmed by EIS and PC evaluations. The addition of a carbon layer and MoS2 nanoparticles to the ZnO increases the BET area. The increased surface area might be partially responsible for the higher photocatalytic performance. (Fig. 11a and b). There is another way to reduce N2 to NH3 via photocatalysis80,81.
Fig. 12 (a) Schematic illustration of the electrospun LZO nanofibers, followed by hydrothermal synthesis of MoS2@ LZO. (b) The band structure demonstrating the type II heterojunction for MoS2@LZO heterostructures with a reduced Fermi level under applied bias and solar illumination and schematic illustration of the PEC-NRR mechanism following the associative pathway for ammonia production. Reproduced from ref. 82 with permission from [American Chemical Society], copyright [2022]. |
Due to the electrical dispersion in LZO, the homogeneous assembly of MoS2@LZO prevents the photogenerated electron hole pair from recombining more quickly. Superior PEC-NRR performances can be seen as a result of the interaction between the localised electronic distribution in the MoS2@LZO heterostructures, oxygen vacancies in La2Zr2O7, excellent conductivity of the nanofibers, and increased exposure of surface sites for electron–hole generation in MoS2. The trapped photoelectron during N2 activation is donated by the reduction of N2 to NH3 at the oxygen vacancy, which serves as the active site for these reactions. The MoS2@LZO PEC catalyst exhibits a maximum ammonia yield rate of approximately 10.4 μg h−1 cm−2 and a higher faradaic efficiency of approximately 2.25% compared to pristine MoS2 and LZO. The heterostructure interface, which is rich in oxygen vacancies and offers a stable design with effective active sites, can be used to confirm this remarkable performance. Compared to the current MoS2 catalysts, MoS2@LZO demonstrates a higher ammonia yield, selective N2 adsorption, and excellent stability. The 2D-MoS2 nanoflakes, the La2Zr2O7 pyrochlore/fluorite structured nanofibers, and their heterostructure interfaces for enhanced ammonia production are presented with PEC-NRR mechanisms under ambient conditions. For various additional catalysis techniques to efficiently drive photoelectrocatalytic activity, the design strategy of inserting an n-type semiconductor over rich oxygen-vacant stable nanostructures to establish an interface can be used. (Fig. 12b).
Fig. 13 Schematic illustration of the energy band structure and electron–hole separation of MoS2/TiO2 composites. Reproduced from ref. 83 with permission from [Royal Society of Chemistry], copyright [2021]. |
Fig. 14 (a) Schematic representation of the band positions and potentials of PPy and MoS2. (b) Photocatalytic CO2 conversion into CH4 and CO gas via Z-scheme mechanism with rGO as redox mediator on the rGO-MoS2/PPy nanocomposite. (c) The conventional type II electron transfer for H2 production on the rGO-MoS2/PPy nanocomposite. Reproduced from ref. 84 with permission from [American Chemical Society], copyright [2020]. |
The synergistic effect of the MoS2, rGO, and PPy nanostructures, which encouraged the separation and migration of the photogenerated charges through the heterojunction interfaces and decreased electron–hole recombination, was attributed to the improvement in the photocatalytic performance of the highly active surface material rGO-MoS2/PPy-150. The PPy in the rGO-MoS2/PPy nanocomposite played multiple roles in improving the photocatalytic reduction of CO due to its superior conductivities, CO2 adsorption capabilities via its amine groups, and electron storage qualities. Better catalytic performance is possible due to the reduced PPy concentration in nanocomposite. Importantly, the strong recycling performance of rGO-MoS2/PPy-150 nanocomposite revealed the robustness and stability of photocatalyst (Fig. 14a–c).
1. MoS2-based materials will be extensively used in a variety of disciplines and could one day become commercially feasible because of their unique physical and chemical qualities. With this, MoS2 will be applied to various media and released into the environment, potentially posing risks to both human health and the ecosystem. Therefore, a detailed evaluation of the toxicity of MoS2-based compounds is required.
2. The application of MoS2-based photocatalysts has so far been confined to laboratory-based research; it has not yet been investigated how well MoS2 performs in real-world environments, and a perfect photocatalyst that can be used on a wide scale and in industry has not yet been developed.
3. Certain MoS2 composites become unstable when exposed to visible light. Therefore, more investigation is required to design materials based on MoS2 that are photosensitive.
4. A significant challenge pertains for the development of practical techniques for the generation of MoS2 in quantities appropriate for industrial applications. For large-scale production, additional research into the scaling up of synthetic approaches is therefore needed.
This journal is © The Royal Society of Chemistry 2022 |