Next Article in Journal
All-Dielectric Color Filter with Ultra-Narrowed Linewidth
Previous Article in Journal
Development of a Mobile Analytical Chemistry Workstation Using a Silicon Electrochromatography Microchip and Capacitively Coupled Contactless Conductivity Detector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Two-Dimensional MoS2: Structural Properties, Synthesis Methods, and Regulation Strategies toward Oxygen Reduction

1
State Key Laboratory of Advanced Technology for Materials Synthesis and Processing, Wuhan University of Technology, Wuhan 430070, China
2
Foshan Xianhu Laboratory of the Advanced Energy Science and Technology, Guangdong Laboratory, Xianhu Hydrogen Valley, Foshan 528200, China
*
Authors to whom correspondence should be addressed.
Micromachines 2021, 12(3), 240; https://doi.org/10.3390/mi12030240
Submission received: 29 January 2021 / Revised: 14 February 2021 / Accepted: 21 February 2021 / Published: 27 February 2021
(This article belongs to the Special Issue Oxygen/Hydrogen-Based Electrocatalysis and Devices)

Abstract

:
Compared with three-dimensional (3D) and other materials, two-dimensional (2D) materials with unique properties such as high specific surface area, structurally adjustable band structure, and electromagnetic properties have attracted wide attention. In recent years, great progress has been made for 2D MoS2 in the field of electrocatalysis, and its exposed unsaturated edges are considered to be active sites of electrocatalytic reactions. In this review, we focus on the latest progress of 2D MoS2 in the oxygen reduction reaction (ORR) that has not received much attention. First, the basic properties of 2D MoS2 and its advantages in the ORR are introduced. Then, the synthesis methods of 2D MoS2 are summarized, and specific strategies for optimizing the performance of 2D MoS2 in ORRs, and the challenges and opportunities faced are discussed. Finally, the future of the 2D MoS2-based ORR catalysts is explored.

1. Introduction

Due to the high energy conversion efficiency and environmental friendliness, hydrogen fuel cells have been regarded as one of the most important clean energy conversion devices to solve the energy crisis and environmental pollution. However, the main challenge in hydrogen fuel cells is to find a highly efficient catalyst for the oxygen reduction reaction (ORR) [1]. The complex multi-step electron transfer process and the slow mass diffusion of the reaction itself lead to slow ORR kinetics [2]. Hitherto, platinum (Pt)-based catalysts are still considered as the most efficient ORR catalysts, but the high cost and poor durability severely limit their wide applications [3]. Therefore, these disadvantages have triggered intense interest in non-Pt catalysts with both high ORR activity and stability. Currently, a variety of non-platinum metal-based ORR catalysts, such as defective carbon materials [4], heteroatom-doped metal-free carbon [5], transition metal/heteroatom-co-doped carbon [6,7], intermetallic compounds [8], and metal nitrides [9], have been synthesized. Among them, two-dimensional (2D) materials with atomic or molecular thickness and infinite planar length exhibit peculiar physicochemical properties, including high specific surface area, adjustable band structure, and electromagnetic properties, which make 2D materials become one of the most important non-Pt-based ORR catalysts [10,11,12,13].
A variety of 2D materials have been found for electrocatalysis. Among heteroatom-doped carbon [14], layered metal hydroxide (LDH) [15], layered metal oxide [16], 2D metal-organic frameworks (2D-MOFs) [17], 2D covalent organic frameworks (2D-COFs) [18], graphite carbonitrides (g-C3N4) [19], hexagonal boron nitride (h-BN) [20], black phosphorous (BP) [21], MXenes [22], and 2D transition metal dichalcogenides (TMDs) [23,24,25], TMDs are a class of materials with great application prospects and basic research value. The chemical formula of TMDs is MX2, where M represents transition metal elements, including group IV (Ti, Zr, or Hf), group V (V, Nb, or Ta), group VI (Mo, W), group VII (Tc, Re), or group X (Pd, Pt), and X refers to a chalcogen atom (S, Se). In a TMD, X-M-X units stack on each other to form a sandwich structure, and the layers are combined by van der Waals forces [26].
As the thickness decreases, 2D TMDs exhibit a series of special properties. During the conversion process of TMDs from bulk to single-layered 2D materials, significant bandgap transitions occur, resulting in excellent photoelectric performance. Besides, due to the high degree of structural controllability and anisotropy of 2D materials, 2D TMDs can be controlled to form various shapes and sizes, doped with different heteroatoms, and their surfaces can be modified with other materials to obtain a unique crystal structure [27]. More importantly, there are unsaturated coordination and dangling bonds at the edges of TMDs nanosheets. These special properties of TMDs nanosheets provide many inspirations for basic research in many fields including catalysis [28,29,30,31,32,33], transistor [34], energy storage [35,36,37,38,39] and sensor [40,41,42]. 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 [26]. 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 [43]. These make 2D MoS2 get more attention among 2D TMDs.
So far, although there have been many reviews about MoS2 electrocatalysis, applying 2D MoS2 in ORRs, especially a detailed summary of the entire process from synthesis to application, is a little concerning. Thus, this review focuses on describing the ORR application of 2D MoS2 and its composite materials. Our aim is to provide comprehensive and cutting-edge information to deeply understand the electrocatalytic ORR for MoS2-based catalysts and reveal the relationship between structure and performance. For different types of MoS2 materials, we will discuss their structural characteristics and internal structure–activity relationship. We believe that electrocatalysts based on MoS2 could help to solve the energy problems we face. Finally, based on the current research results, the difficulty and deficiency encountered in the research are proposed, and the outlook is also discussed.

2. Structural Properties of 2D MoS2

The properties of 2D MoS2 are highly dependent on their crystal phase, structure size and chemical composition. In the MoS2 structure, the coordination of S elements and the stacking state between layers can form different electronic characteristics. The most common crystal structures of MoS2 are 1T and 2H phases, where the number represents the number of layers of the S-Mo-S layer in a unit cell, and T and H represent the tetragonal crystal system, and the hexagonal crystal system, respectively [44]. As shown in Figure 1A, in the 2H phase, each Mo atom exists in the center of the hexagonal prism coordination structure and covalently bonded to six S atoms, and exhibits semiconductor properties. In the 1T phase, one Mo atom and six S atoms form a twisted octahedral coordination structure, with metallicity [45]. In MoS2, the S-Mo-S of each planar unit is regarded as a single layer, which is composed of two layers of S atoms and an intermediate layer of Mo atoms [46]. Due to the weak van der Waals force connection between the layers, the peeling between the layers is easy to occur, which makes it possible to synthesize single-layer MoS2 [47].
As shown in Figure 1B, the bulk material of MoS2 is an indirect-bandgap semiconductor, in which the top of the valence band is located at point G and the bottom of the conduction band is located at the midpoint of the axis of symmetry of G–K, while in single-layer MoS2, the top of the valence band and the bottom of the conduction band coincide with point K, indicating that single-layer MoS2 is a direct-bandgap semiconductor [48]. This indirect-to-direct transition of the bandgap from bulk to the single-layered 2D material is caused by the quantum confinement effect. The electronic properties of 2D MoS2 are closely related to the d electrons existing in the d orbital between the bonding band and the anti-bonding band of the Mo–S bond. In 2H-MoS2, the non-bonded d orbitals are fully occupied, resulting in semiconductor properties. In 1T-MoS2, the non-bonded d orbitals are partially filled, resulting in higher conductivity. As the d-orbital electron occupancy rate decreases, 2D MoS2 gradually changes from a semiconductor to a conductor, which endows the material with different physical and chemical properties [27]. When the thickness of MoS2 drops to the atomic level, a large number of exposed surfaces can undergo oxidation or reduction reactions with various reagents [49]. These surface atoms can escape in the reaction to form vacancy defects, simultaneously leading to the disorder of the nanostructure and a reduction in the coordination number of surface atoms. By adjusting the vacancy defects, we can further modify the electronic characteristics of MoS2. At the same time, unsaturated edges and dangling bonds often appear on the edges of 2D MoS2. These unique edge defects also provide possible active sites for catalytic reactions [2,50].
So far, many 2D MoS2-based materials have been applied to ORRs. At present, it is generally believed that the active site of 2D MoS2 in the ORR is the unsaturated Mo site exposed at the edge. Because the Mo site is positively charged due to the possible polarization effect of the surrounding negatively charged S atoms, it can, therefore, be easily combined with negatively charged O atoms [51]. Simultaneously, the vacancy defects on the basal plane also contribute to ORR performance. In general, an excellent ORR electrocatalyst must have favorable stability, small overpotentials, high current densities, and low manufacturing cost [52]. Although MoS2 has a low cost and high chemical stability, its large inert basal plane and poor electron transport properties limit its catalytic activity. However, it is undeniable that after adjusting the structure of 2D MoS2, it still possesses unique advantages and broad development prospects for ORRs. Here are some advantages of 2D MoS2 in ORRs:
(1) The uncoordinated metallic edge centers not only provide a certain intrinsic ORR activity but also give an opportunity for effective functionalization with different groups.
(2) Compared with bulk MoS2, nanostructured MoS2 has a smaller particle size and a larger specific surface area. The small size 2D MoS2 is more sensitive to the catalytic reactions compared with bulk MoS2. At the same time, the large surface area can anchor more active sites [53]. Besides, 2D MoS2 has a unique open structure, which makes ORR reactants easily access the active site.
(3) 2D MoS2 with a uniformly exposed lattice plane can supply a much simpler condition as an ideal platform for combing experiments and theoretical results, which is helpful to further understand the origin of ORR activity.

3. The Synthesis Methods

3.1. Liquid-Phase-Stripping Method

Liquid-phase stripping is a common method for preparing monolayer and multilayer materials, usually based on ultrasound assistance (in a solvent or aqueous solution) and subsequent centrifugal process [43,54]. As shown in Figure 2, this method is extremely compatible. Since the MoS2 layers are connected by weak van der Waals forces, the 2D MoS2 layer can be peeled from bulk by weakening the force between the layers. Recently, an ion intercalation method was created for layered materials that can attract guest molecules between layers and form clathrates. First, intercalation agents such as n-butyllithium [55,56] and iodine bromide are used to increase the interlayer spacing, weaken the interlayer force, and reduce the energy barrier for peeling. Then, the layered material is peeled off by ultrasonic waves [55], thermal shock [57], and other treatments. Ambrosi et al. obtained complete MoS2 nanosheets under mild conditions by using a variety of Li-based intercalants and explored the effects of different intercalants on the material structure and catalysis performance. Among such Li-based intercalants (tert-butyllithium (t-Bu-Li), n-butyllithium (n-Bu-Li), and methyllithium (Me-Li)), the stripping efficiency of n-Bu-Li and t-Bu-Li is the highest, and the obtained MoS2 nanosheets have higher catalytic performance [56]. However, this liquid-phase intercalation method is susceptible to changes in the reaction environment, causing structural deformation of some materials [58,59]. Another problem is the re-aggregation of the material during the removal of the intercalant.
To overcome the shortcomings of the intercalation method, such as time-consuming process and product instability, the latest stripping strategy is to place MoS2 in a solvent with a specific surface energy and then use ultrasonic waves to destroy the material crystallites to produce nanosheets. It was discovered that organic solvents such as N-cyclohexyl-2-pyrrolidone (CHP) [60] and N-methyl-2-pyrrolidone (NMP) [43,61] can effectively strip MoS2. With further exploration, it was found that using surfactants such as sodium cholate [62,63] as a stabilizer can have a synergistic effect with the solvent and effectively prevent the nanosheets from re-aggregating (Figure 3A). For the first time, Peter et al. employed a series of polymers (Polybutadiene-styrene (PBS), Polystyrene (PS), polyvinyl alcohol (PVA), Polyvinyl chloride (PVC), etc.) as stabilizers to replace surfactants, which proved that MoS2 could be stripped and could remain stable in a mixed solution of a variety of organic solvents and polymers (Figure 3B,C). More importantly, by a proposed simple model, they predicted that polymer stabilization is effective only when the solubility parameters of the nanosheets, polymer, and solvent are similar [64].

3.2. Hydrothermal/Solvothermal Method

The hydrothermal/solvothermal method usually involves transferring the Mo and S precursor to an autoclave and heating at different temperatures. MoS2 nanosheets with different structures can be obtained by adjusting the reactant composition, reaction temperature, time, and pH value [65,66]. This method is easy to control and causes little environmental pollution. In addition, external forces such as microwaves [67,68] and strong magnetic fields [69] are added based on hydrothermal/solvothermal methods with the deepening of research, which accelerates the synthesis of materials and promotes their combination with other components. As shown in Figure 4A, Hao et al. applied the microwave-assisted hydrothermal synthesis method to obtain flower-structured MoS2 nanosheets. At the same time, Ni ions were embedded on the MoS2 nanosheets to form a unique self-supporting structure. This unique structure effectively inhibited the stacking of MoS2 nanosheets [70].
The self-sacrificing template method is also employed to suppress the re-stacking of nanosheets. For example, taking g-C3N4 as a self-sacrificing template, Huang et al. first obtained a layered MoO3/g-C3N4 precursor with molybdate and g-C3N4 by hydrothermal synthesis and then processing at 900 °C. The g-C3N4 template was decomposed, and MoS2 nanosheets were finally obtained with a porous ultrathin structure (Figure 4B) [46].

3.3. Chemical Vapor Deposition (CVD) Method

Chemical vapor deposition (CVD) is an important synthesis method for 2D materials. Compared with the solvent method and the hydrothermal synthesis method, large-size MoS2 nanosheets can be prepared [71]. However, for large-scale synthesis, MoS2 often forms a rod shape instead of a sheet shape [71,72,73,74], so high-vacuum reaction conditions are required and precious metal materials such as gold are used as the substrate [75,76,77]. The harsh reaction conditions limit the further application of the CVD method for the synthesis of 2D MoS2. To overcome such difficulties, simpler synthesis methods have been further explored. Lee et al. took MoO3 and S powder as reactants to directly synthesize MoS2 nanosheets on Si/SiO2 substrates treated with graphene-like molecules (Figure 5A). Under moderate temperatures, a large area of MoS2 layers can be directly obtained on the surface of amorphous SiO2 through the reaction of S vapor and MoO3. Without using highly crystalline metal substrates or an ultrahigh-vacuum environment, the cost was greatly reduced [71].
Due to the bottom-up synthesis principle of CVD, the continuous monolayer MoS2 prepared by CVD is usually atomically neat, and its edges lack saturated atoms as main active sites [78], very unfavorable for the catalytic reaction. To improve this, Tan et al. first prepared porous gold as a template and then synthesized irregularly shaped MoS2 nanosheets at the holes of the template (Figure 5B). To further enhance the catalytic activity of MoS2 nanosheets, a certain out-of-plane strain on the gold template was introduced so that MoS2 was born on the curved inner surface of the template, which gave MoS2 a certain degree of strain and improved its hydrogen evolution reaction (HER) catalytic performance [76]. However, this method is still too expensive for mass production; therefore, to reduce the cost, cheaper nanoporous metal templates (such as copper, nickel) could be considered.

4. Regulation Strategies toward Oxygen Reduction

4.1. 2D MoS2 for ORRs

For the ORR, there are currently three mainstream reaction mechanisms: (1) dissociation mechanism, where the O-O bond of oxygen is directly broken when adsorbed on the catalyst surface and O* is reduced into OH* and H2O* in the next reaction; (2) association mechanism, where OOH* is first produced during the reaction and then split into O* and OH*; and (3) peroxygen mechanism, in which when the two-electron steps reach OOH* and HOOH*, the latter is decomposed into OH*. The formation of adsorbed oxidizing substances (e.g., O*, OH*, OOH*) caused by the aqueous electrolyte and the reaction itself is considered as a decisive process to determine the performance of the catalyst. All the above three mechanisms indicate that the ORR rate is actually limited by the formation of OOH* and the removal of OH*. The binding energy on the surface of the ORR intermediate should not be too strong or too weak [79]. This phenomenon can be represented by a volcano figure [80,81]. The catalyst located near the apex of the volcano (Figure 6B) theoretically has the best catalytic performance.
In 2D MoS2, when the size of the S-Mo-S three-layer flake is reduced laterally, the coordination number at the edges and corner atoms would become less, which provides rich active sites for electrocatalysis [82,83]. Through theoretical calculations, it can be found that edge-exposed unsaturated Mo atoms are the active sites of ORRs [84]. These unsaturated sites can effectively regulate the adsorption/desorption of reactants and intermediate products on the surface of the catalyst. Although it has been proved that MoS2 nanosheets exhibit a certain ORR performance, their reaction path is mostly a two-electron path. At the same time, the catalytic performance of 2D MoS2 is lower than that of the commercial Pt/C catalyst and even has a certain gap compared with the popular metal-nitrogen-carbon (M-N-C) catalyst. Two main factors lead to the low catalytic performance of MoS2. The first is that pristine 2D MoS2 has low conductivity, which is not conducive to the electron transport in ORR processes and thus limits the reaction rate. The second is that, as a 2D material, although pristine 2D MoS2 has a large specific surface area, its active sites exist only at the edges and defects of the material, and the huge basal plane is inactive, which greatly hinders the reaction. Thus, how to properly adjust the structure of MoS2 to increase the number of active sites and enhance conductivity becomes a key issue in enhancing the catalytic activity of MoS2.

4.2. Controlling Electronic Structure by Doping

4.2.1. Non-Metallic Element Doping

Studies have shown that incorporating non-metallic elements in MoS2 nanosheets can effectively regulate the ORR activity of catalysts due to the changed electronic structures around the doped atoms. It has been demonstrated that the introduction of heteroatoms (O, P, N, etc.) in MoS2 can adjust the electronic structure of materials, reduce the bandgap, improve conductivity, and provide catalytically active sites. As presented in Figure 7, Huang et al. used g-C3N4 as a self-sacrificial template, and through hydrothermal and subsequent heat treatment, MoS2 nanosheets with abundant edge defects were obtained, and then oxygen was introduced into MoS2 through H2O2 treatments. The sample with the best catalytic performance had an initial potential of 0.94 V and a half-wave electric potential of 0.80 V in 0.1M KOH. It is interesting to note that the ORR selectivity of O-MoS2 changed from a two-electron path to a four-electron path. They believed that this change was due to the highly electronegative O atom polarizing the unsaturated adjacent Mo atom. The Mo atom at the edge generated an additional positive charge and was preferentially adsorbed by oxygen molecules, accelerating the ORR process [51]. Through density functional theory (DFT) calculations, Xie et al. also reported that compared with the original 2H-MoS2 nanosheets (1.75 eV), oxygen-bound MoS2 nanosheets have a (1.30 eV) narrower bandgap (Figure 7D), indicating that the combination of oxygen and MoS2 nanosheets may generate more carriers and increase the intrinsic conductivity of the material [85]. This leads to enhanced ORR and HER activity.
In addition, the incorporation of low-electronegativity atoms such as N and P into MoS2 would significantly enhance ORR catalytic activity. However, there are different explanations for performance enhancement. As shown in Figure 8A,B Huang et al. doped P atoms into MoS2 nanosheets, which effectively improved the ORR performance of the MoS2 nanosheets, and explained the enhancement of ORR performance using boundary molecular orbital theory. Compared to the original MoS2 nanosheets, the energy levels of highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO) of P-doped MoS2 nanosheets are higher [46]. In semiconductors, increasing the frontier orbital energy is beneficial to electron donation [86], which makes oxygen adsorption and the formation of the intermediate product OH easier, accelerating the reaction rate.
Zhang et al. calculated the overpotential of N- or P-doped single-layer MoS2 in the ORR under acidic conditions and found that replacing S atoms in the MoS2 single layer with P or N atoms can introduce high spin density into the basal plane of MoS2, thereby improving its ability to activate O2 and making the ORR step more inclined to proceed through the more efficient 4e pathway [87]. It is worth noting that the ORR performance of P-doped MoS2 is worse than that of N-doped MoS2 due to the excessively high adsorption energy of the intermediate product of P-doped MoS2. However, this does not mean that P-doped MoS2 has no application prospects. As shown in Figure 8C, through DFT calculations, Liu et al. investigated the ORR mechanism and key active sites of P-doped single-layer MoS2 with different structures and the P content in an alkaline environment. It can be seen that double P-doped MoS2 with a higher P content had better performance than single P-doped MoS2. Its catalytic active site is the S2 atom, which is adjacent to the two P atoms, and the two adjacent P atoms lead to a reduction in the charge of the S2 site and provide a stronger hydrogen bond, boosting the adsorption of H2O and OH groups. Simultaneously, compared with single P-doped MoS2, double P-doped MoS2 had an appropriate intermediate product adsorption energy (Figure 8D). By DFT calculations, the optimal P doping amount of the catalyst for the ORR was 5.5%, close to that obtained in the experiment (4.7%) [88]. This work provides a reasonable explanation for the contribution of P elements to catalytic activity that has long puzzled researchers.

4.2.2. Metal Element Doping

Metallic elements can also be added to MoS2 to strengthen ORR catalytic activity. It has been reported that doping Cu, Co, and other transition metals in the S vacancy of MoS2 can effectively enhance ORR catalytic activity. Xiao et al. calculated the theoretical ORR catalytic performance of single-layer MoS2 doped with Co/Ni. It is known that doping metal atoms into the basal plane of single-layer MoS2 can adjust the adsorption energy of the reaction intermediates on the catalyst surface, thereby increasing catalysis activity. In Co/Ni-doped single-layer MoS2, Co/MoS2 was similar to the FeN4 active site in the M-N-C catalyst, and Ni/MoS2 was similar to the CoN4 active site. Therefore, the Co/Ni-doped single-layer MoS2 had higher ORR catalytic activity than the original MoS2 [89]. In addition, Urbanova et al. and He each incorporated metal elements such as Fe, Mn, Ti, and V into MoS2 and found a significant improvement in ORR activity (Figure 9A,B) [90,91]. Although much work has been done on the direction of metal doping in MoS2, the choice of metal types is a simple try-out method. Using DFT, Wang et al. systematically explored the ORR performance of single-layer MoS2 doped with various transition metals. The calculation result showed that transition metal atoms could be embedded in the S vacancy and could significantly change the electronic structure of the material. As shown in Figure 9C,D, they found that Cu-doped MoS2 had the best theoretical binding strength for ORR intermediates and therefore had the best ORR catalytic activity, indicating its high potential as an efficient single-atom catalyst for ORRs [92]. This work may serve as a useful guidance for further developing metal-doped MoS2-based ORR catalysts.
Since metal-S bonds are regarded as reactive sites in metal-doped transition metal dichalcogenides (TMDs), increasing the density of metal atoms on the surface of MoS2 as much as possible and reducing the metal aggregation during heating can be the key to improving the reactivity of the catalyst, but these are difficult to achieve. The metal loading on the surface of most metal-doped MoS2 catalysts is very low and difficult to precisely control. Therefore, how to overcome these difficulties is crucial to enhance the ORR performance of metal-doped MoS2.

4.3. Activating the Basal Plane to Improve the Catalytic Performance of ORRs

To obtain higher ORR performance, many studies have been devoted to optimizing the composition and structure of 2D MoS2 through post-processing methods. However, these methods will inevitably destroy the stability of the material and reduce conductivity [93,94]. In fact, the huge impact of the large-area inert basal plane of 2D MoS2 on catalytic performance has been neglected for a long time. Thus, how to effectively activate the basal plane of 2D MoS2 has become a research hotspot.

4.3.1. Strain Engineering Activates Inert Basal Plane

It has been found that introducing tensile or compressive stress into MoS2 can affect the ORR performance. The ultrathin flexible structure of 2D MoS2 makes it easier to generate and maintain strain. Through DFT calculations, Zhao et al. observed that tensile strain would shift the center of the p orbital of chalcogen on the surface of TMDs to move to the Fermi level. This shift can enhance the adsorption of materials for ORR intermediates, thereby effectively promoting the ORR process. At the same time, for 2D MoS2, the larger biaxial strain would cause a certain degree of lattice mismatch and then inevitably generate new vacancies on the basal plane. The newly generated basal surface vacancies cooperate with strain and further enhance its catalytic performance [95]. These results indicate that strain engineering is an effective way to adjust the ORR activity of 2D TMDs. Recently, a lot of methods have been developed for introducing strain into the material during synthesis, such as using a template with a certain curvature (nanospheres, porous curved metal) or transferring the material to a patterned substrate. For example, Li et al. transferred MoS2 to SiO2 nanocones by means of capillary pressure and added strain into MoS2 (Figure 10) [96]. However, the introduction of strain into 2D MoS2 in the experiment to enhance its ORR performance is still not widely used.

4.3.2. Stable Phase Change Activated Inert Basal Plane

Two-dimensional MoS2 has different phases, causing various catalytic activities. The 1T phase is a metal phase and has the highest conductivity. However, it is metastable at room temperature and would slowly change into the 2H phase spontaneously. Improving the performance of 2D MoS2 by controlling the phase transition has broad research prospects. According to previous reports, the 1T phase can be obtained in a single layer of MoS2 through different methods such as alkali metal intercalation-exfoliation and plasma hot-electron injection. In addition, the 1T phase can stably exist by doping Re, Tc, Mn, and other metal elements in nanosheets as electron donors [97,98,99,100]. The high electrical conductivity and catalytic activity make 1T-MoS2 a more effective electrocatalyst than 2H-MoS2. As exhibited in Figure 11, Sadighi prepared 1T-MoS2 nanosheets via the in situ liquid-redox intercalations of sodium ions and solvent exchange exfoliation. The 1T-MoS2 nanosheets showed an almost equivalent half-wave potential and enhanced the limit diffusion current as compared to Pt/C [101]. The 1T-phase MoS2 nanosheets prepared by the alkali metal intercalation method can spontaneously transform from the 1T phase to the thermodynamically stable 2H phase in the air [59]. This unstable property hinders catalytic stability. At the same time, the conditions of this preparation method are relatively harsh, and the 1T-phase/2H-phase ratio is not high. As a result, it is necessary to develop methods that can easily synthesize 1T-phase MoS2 with high stability.

4.4. Compounding Other Materials to Enhance Conductivity

It is well known that the low conductivity and aggregation of nanocrystalline sheets can reduce the catalytic activity of 2D MoS2. To avoid these cases, the combination of 2D MoS2 with materials possessing high conductivity to form a composite has attracted much attention. Among them, compounding 2D MoS2 with carbon materials to form a hybrid structure can effectively boost the conductivity of materials and reduce aggregation [102,103]. As presented in Figure 12A,B, by employing 4-iodophenyl-functionalized monolayer MoS2 as a template, Yuan et al. synthesized a series of MoS2-coupled sandwich-like conjugated microporous polymers (M-CMPs). After direct pyrolysis, the 2D MoS2/N-doped porous carbon (M-CMPs-T) hybrid with a large specific surface area and layered porous structure was finally obtained. Compared with porous nitrogen-doped carbon without MoS2, 2D porous carbon hybrids exhibited a higher half-wave potential and limit diffusion current in the ORR [104].
Zhao et al. synthesized hybrid nanorods using dopamine and ammonium through self-assembly followed by annealing, while an ultrathin monolayer of MoS2 was formed under the constraints of the carbon skeleton (Figure 12C,D). The unique tubular structure, large specific surface area, relatively high N dopant content, and oxygen-bonded MoS2 monolayer endowed the MoS2/ nitrogen-doped carbon hybrid nanotubes (C HNT) with high electronic conductivity and abundant ORR active sites. The as-prepared MoS2/C HNT possessed ORR activity close to commercial Pt/C under alkaline conditions [105]. Zhou et al. used hydrothermal and annealing treatment to attach MoS2 nanosheets to graphene oxide (GO), and the interconnected MoS2 nanosheets formed a layered structure that completely covered the GO surface. The material revealed excellent ORR performance due to the enhanced conductivity of MoS2. At the same time, graphene oxide effectively limited the aggregation of MoS2 nanosheets [106] (Figure 12E,F).

4.5. 2D MoS2 as a Co-Catalyst

Although 2D MoS2 can be directly used as an ORR catalyst, its performance is mostly unsatisfactory due to poor conductivity and rare active sites. Thus, to achieve higher catalytic performance, 2D MoS2 can be used as a co-catalyst through having a synergistic effect with other active substances. As shown in Figure 13A–C, Mao et al. used g-C3N4 as a self-sacrificial template to synthesize ultrathin Ni3S2/MoS2 nanosheets with rich heterogeneous interfaces. The heterostructure formed on the two-phase surface of Ni3S2 and MoS2 generated a large number of unique Mo-Ni-S catalytic active sites. The catalyst revealed higher ORR catalytic activity than the commercial Pt/C [107]. Kwon et al. synthesized an iron phthalocyanine (FePc) and 1T-phase MoS2 nanoplate heterostructure catalyst through a one-step hydrothermal reaction. Herein, iron phthalocyanine promoted the transition of nanosheets from the 2H phase to the 1T phase, increasing the electrical conductivity. The unique non-planar geometry of the Fe-N4 active site of FePc further enhanced the ORR catalytic activity of FePc-MoS2 (Figure 13D,E) [108]. Our group grew 2D MoS2 on a N-doped carbon skeleton derived from zeolitic imidazolate frameworks-8 (ZIF-8) and constructed an excellent multifunctional Mo/N/C@MoS2 catalyst toward HER, ORR and oxygen evolution reaction (OER). Its excellent catalytic performance originated from the synergistic effect between the abundant main active sites on the edges of MoS2, the Mo-N phase coupling center, and the N-induced active sites of nanocarbon frameworks. The porous 3D carbon skeleton structure also accelerated the diffusion rate of products and enhanced electrical conductivity [109].
Based on a strategy derived from ZIF-67, Li et al. synthesized a core-double-shell Co9S8@Co9S8@MoS2 heterostructure. The Co@Co9S8 precursor provided sufficient space to limit the growth of MoS2, which prevented the aggregation of MoS2 and improved the electrical conductivity of the material. The core-double-shell Co9S8@Co9S8@MoS2 heterostructure exhibited a synergistic interaction between Co9S8 and MoS2 and modified electronic structures. Such characteristics made Co9S8@Co9S8@MoS2 exhibit excellent electrocatalytic performance, close to commercial Pt/C [111]. As shown in Figure 13F, Vattikuti et al. deposited silver nanocrystals on the surface of 2D MoS2 nanosheets by a simple microwave-assisted method. Due to the Ag/MoS2 heterostructure and the increased specific surface area, the catalyst presented better ORR activity than pristine MoS2 in alkaline solutions [110].
When used as a co-catalyst, 2D MoS2 can provide a large-area substrate supporting other active materials (such as precious metals Pt, Ag) and reduce the dosage of active materials. In addition, the heterostructure formed by MoS2 and other materials can provide good catalytic activity. Although a variety of MoS2-based heterostructure catalysts with excellent ORR catalytic performance have been prepared, the exact reaction that occurs on the heterostructure and the internal mechanism of improving the catalytic performance are still not determined. At the same time, the tunable, controllable, and large-scale synthesis of heterostructures is hard to achieve.

5. Outlook

As a kind of electrocatalytic material with great potential, 2D MoS2 has shown great application prospects on the HER, but its development in ORRs is stagnant. Most of the 2D MoS2 materials possess low-electron-transport characteristics and a large area of the unused inert basal plane, which limits their applications. This review summarized the research progress in 2D MoS2 in ORRs in recent years; focused on the specific works on 2D MoS2, from synthesis methods to ORR performance optimization; and identified the key obstacles that 2D MoS2 faces in ORRs. To promote the ORR performance of such materials, electron transport and the increase in active sites are basic problems to resolve. Some strategies for resolving these problems, such as increasing exposed edges, heteroatom doping, constructing heterostructures, designing composite materials, and stable phase transitions to enhance electron and product transport, etc., were also discussed in this article. These strategies make it possible to develop high-performance 2D MoS2-based ORR catalysts. However, despite numerous encouraging outcomes, there are still many issues to be resolved.
(1) Controllable synthesis. The current bottom-up synthesis method requires relatively harsh conditions, and the top-down synthesis method would cause the aggregation of nanosheets, which lowers the quality of 2D MoS2. Striving to achieve large-scale synthesis of high-quality 2D MoS2 materials is still the basic requirement.
(2) At present, there is no consensus on the detailed mechanism of 2D MoS2 and 2D MoS2-based catalysts in the ORR, especially the specific contributions of elements such as O, N, and P. Thus, it is necessary to further analyze the relationship between structure composition and performance to guide future development.
(3) Effective usage of the large-area basal plane with reactive inertness is necessary. The current work mainly focuses on heteroatom doping and material recombination. However, fine control of heteroatom doping has not been well performed. Further exploring the synergy between MoS2 and other ORR active materials, and the internal mechanism of enhanced ORR activity, is the focus for future research.
(4) The intrinsic properties of MoS2 can be activated by a stable phase change and strain engineering. Thus, we should find more convenient methods to synthesize long-term stable 1T-phase MoS2 nanosheets to improve the adverse effects of spontaneous phase transition on the catalytic performance.

Author Contributions

Conceptualization and writing—original draft preparation, H.X.; investigation, Q.M. and H.B.; formal analysis, J.M. and L.C.; writing—original draft preparation and writing—review and editing, J.Z. and S.M.; supervision, S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 22075223).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Alaswad, A.; Omran, A.; Sodre, J.R.; Wilberforce, T.; Pignatelli, G.; Dassisti, M.; Baroutaji, A.; Olabi, A.G. Technical and Commercial Challenges of Proton-Exchange Membrane (PEM) Fuel Cells. Energies 2020, 14, 144. [Google Scholar] [CrossRef]
  2. Chen, Y.; Yang, K.; Jiang, B.; Li, J.; Zeng, M.; Fu, L. Emerging two-dimensional nanomaterials for electrochemical hydrogen evolution. J. Mater. Chem. A 2017, 5, 8187–8208. [Google Scholar] [CrossRef]
  3. Yoo, M.; Yu, Y.-S.; Ha, H.; Lee, S.; Choi, J.-S.; Oh, S.; Kang, E.; Choi, H.; An, H.; Lee, K.-S.; et al. A tailored oxide interface creates dense Pt single-atom catalysts with high catalytic activity. Energy Environ. Sci. 2020, 13, 1231–1239. [Google Scholar] [CrossRef]
  4. Zhu, J.; Huang, Y.; Mei, W.; Zhao, C.; Zhang, C.; Zhang, J.; Amiinu, I.S.; Mu, S. Effects of Intrinsic Pentagon Defects on Electrochemical Reactivity of Carbon Nanomaterials. Angew. Chem. Int. Ed. Engl. 2019, 58, 3859–3864. [Google Scholar] [CrossRef]
  5. Jiang, H.; Gu, J.; Zheng, X.; Liu, M.; Qiu, X.; Wang, L.; Li, W.; Chen, Z.; Ji, X.; Li, J. Defect-rich and ultrathin N doped carbon nanosheets as advanced trifunctional metal-free electrocatalysts for the ORR, OER and HER. Energy Environ. Sci. 2019, 12, 322–333. [Google Scholar] [CrossRef]
  6. Jin, H.; Zhou, H.; He, D.; Wang, Z.; Wu, Q.; Liang, Q.; Liu, S.; Mu, S. MOF-derived 3D Fe-N-S co-doped carbon matrix/nanotube nanocomposites with advanced oxygen reduction activity and stability in both acidic and alkaline media. Appl. Catal. B Environ. 2019, 250, 143–149. [Google Scholar] [CrossRef]
  7. Zhang, N.; Zhou, T.; Chen, M.; Feng, H.; Yuan, R.; Zhong, C.a.; Yan, W.; Tian, Y.; Wu, X.; Chu, W.; et al. High-purity pyrrole-type FeN4 sites as a superior oxygen reduction electrocatalyst. Energy Environ. Sci. 2020, 13, 111–118. [Google Scholar] [CrossRef]
  8. Du, X.X.; He, Y.; Wang, X.X.; Wang, J.N. Fine-grained and fully ordered intermetallic PtFe catalysts with largely enhanced catalytic activity and durability. Energy Environ. Sci. 2016, 9, 2623–2632. [Google Scholar] [CrossRef]
  9. Lou, Y.; Liu, J.; Liu, M.; Wang, F. Hexagonal Fe2N Coupled with N-Doped Carbon: Crystal-Plane-Dependent Electrocatalytic Activity for Oxygen Reduction. ACS Catal. 2020, 10, 2443–2451. [Google Scholar] [CrossRef]
  10. Guo, H.W.; Hu, Z.; Liu, Z.B.; Tian, J.G. Stacking of 2D Materials. Adv. Funct. Mater. 2020. [Google Scholar] [CrossRef]
  11. Zhang, Q.; Zhang, J.; Wan, S.; Wang, W.; Fu, L. Stimuli-Responsive 2D Materials Beyond Graphene. Adv. Funct. Mater. 2018, 28. [Google Scholar] [CrossRef]
  12. Ye, C.; Chao, D.; Shan, J.; Li, H.; Davey, K.; Qiao, S.-Z. Unveiling the Advances of 2D Materials for Li/Na-S Batteries Experimentally and Theoretically. Matter 2020, 2, 323–344. [Google Scholar] [CrossRef] [Green Version]
  13. Walsh, L.A.; Hinkle, C.L. van der Waals epitaxy: 2D materials and topological insulators. Appl. Mater. Today 2017, 9, 504–515. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Wan, J.; Yao, H.; Zhang, L.; Lin, K.; Wang, L.; Yang, N.; Liu, D.; Song, L.; Zhu, J.; et al. Few-layer graphdiyne doped with sp-hybridized nitrogen atoms at acetylenic sites for oxygen reduction electrocatalysis. Nat. Chem. 2018, 10, 924–931. [Google Scholar] [CrossRef]
  15. Qian, Y.; An, T.; Sarnello, E.; Liu, Z.; Li, T.; Zhao, D. Janus Electrocatalysts Containing MOF-Derived Carbon Networks and NiFe-LDH Nanoplates for Rechargeable Zinc–Air Batteries. ACS Appl. Energy Mater. 2019, 2, 1784–1792. [Google Scholar] [CrossRef]
  16. Lim, J.; Jin, X.; Jo, Y.K.; Lee, S.; Hwang, S.J. Kinetically Controlled Layer-by-Layer Stacking of Metal Oxide 2D Nanosheets. Angew. Chem. Int. Ed. Engl. 2017, 56, 7093–7096. [Google Scholar] [CrossRef]
  17. Xu, Y.; Huang, Z.; Wang, B.; Liang, Z.; Zhang, C.; Wang, Y.; Zhang, W.; Zheng, H.; Cao, R. A two-dimensional multi-shelled metal-organic framework and its derived bimetallic N-doped porous carbon for electrocatalytic oxygen reduction. Chem. Commun. 2019, 55, 14805–14808. [Google Scholar] [CrossRef] [PubMed]
  18. Zhan, X.; Chen, Z.; Zhang, Q. Recent progress in two-dimensional COFs for energy-related applications. J. Mater. Chem. A 2017, 5, 14463–14479. [Google Scholar] [CrossRef]
  19. Wang, Y.; Yang, W.; Chen, X.; Wang, J.; Zhu, Y. Photocatalytic activity enhancement of core-shell structure g-C3N4@TiO2 via controlled ultrathin g-C3N4 layer. Appl. Catal. B Environ. 2018, 220, 337–347. [Google Scholar] [CrossRef]
  20. He, Z.; Kim, C.; Lin, L.; Jeon, T.H.; Lin, S.; Wang, X.; Choi, W. Formation of heterostructures via direct growth CN on h-BN porous nanosheets for metal-free photocatalysis. Nano Energy 2017, 42, 58–68. [Google Scholar] [CrossRef]
  21. Liu, H.; Hu, K.; Yan, D.; Chen, R.; Zou, Y.; Liu, H.; Wang, S. Recent Advances on Black Phosphorus for Energy Storage, Catalysis, and Sensor Applications. Adv. Mater. 2018, 30, e1800295. [Google Scholar] [CrossRef]
  22. Zhang, Y.Z.; El-Demellawi, J.K.; Jiang, Q.; Ge, G.; Liang, H.; Lee, K.; Dong, X.; Alshareef, H.N. MXene hydrogels: Fundamentals and applications. Chem. Soc. Rev. 2020, 49, 7229–7251. [Google Scholar] [CrossRef] [PubMed]
  23. Han, T.; Liu, H.; Chen, S.; Chen, Y.; Wang, S.; Li, Z. Fabrication and Characterization of MoS2/h-BN and WS2/h-BN Heterostructures. Micromachines 2020, 11, 1114. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, M.; Zhang, L.; Huang, M.; Zhang, Q.; Zhao, X.; He, Y.; Lin, S.; Pan, J.; Zhu, H. One-step synthesis of a hierarchical self-supported WS2 film for efficient electrocatalytic hydrogen evolution. J. Mater. Chem. A 2019, 7, 22405–22411. [Google Scholar] [CrossRef]
  25. Wang, P.; He, H.; Pu, Z.; Chen, L.; Zhang, C.; Wang, Z.; Mu, S. Phosphorization engineering ameliorated the electrocatalytic activity for overall water splitting on Ni3S2 nanosheets. Dalton Trans. 2019, 48, 13466–13471. [Google Scholar] [CrossRef]
  26. Canadell, E.; LeBeuze, A.; El Khalifa, M.A.; Chevrel, R.; Whangbo, M.H. Origin of metal clustering in transition-metal chalcogenide layers MX2 (M = Nb, Ta, Mo, Re; X = S, Se). J. Am. Chem. Soc. 1989, 111, 3778–3782. [Google Scholar] [CrossRef]
  27. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  28. Carroll, G.M.; Zhang, H.; Dunklin, J.R.; Miller, E.M.; Neale, N.R.; van de Lagemaat, J. Unique interfacial thermodynamics of few-layer 2D MoS2 for (photo)electrochemical catalysis. Energy Environ. Sci. 2019, 12, 1648–1656. [Google Scholar] [CrossRef]
  29. Wang, J.; Yan, M.; Zhao, K.; Liao, X.; Wang, P.; Pan, X.; Yang, W.; Mai, L. Field Effect Enhanced Hydrogen Evolution Reaction of MoS2 Nanosheets. Adv. Mater. 2017, 29, 1604464. [Google Scholar] [CrossRef]
  30. Masurkar, N.; Thangavel, N.K.; Arava, L.M.R. CVD-Grown MoSe2 Nanoflowers with Dual Active Sites for Efficient Electrochemical Hydrogen Evolution Reaction. ACS Appl Mater Interfaces 2018, 10, 27771–27779. [Google Scholar] [CrossRef]
  31. Gopalakrishnan, D.; Lee, A.; Thangavel, N.K.; Reddy Arava, L.M. Facile synthesis of electrocatalytically active NbS2 nanoflakes for an enhanced hydrogen evolution reaction (HER). Sustain. Energy Fuels 2018, 2, 96–102. [Google Scholar] [CrossRef]
  32. Mao, J.; Wang, Y.; Zheng, Z.; Deng, D. The rise of two-dimensional MoS2 for catalysis. Front. Phys. 2018, 13, 138118. [Google Scholar] [CrossRef]
  33. Wei, Z.; Hsu, C.; Almakrami, H.; Lin, G.; Hu, J.; Jin, X.; Agar, E.; Liu, F. Ultra-high-aspect-ratio vertically aligned 2D MoS2-1D TiO2 nanobelt heterostructured forests for enhanced photoelectrochemical performance. Electrochim. Acta 2019, 316, 173–180. [Google Scholar] [CrossRef]
  34. Kim, T.; Kang, D.; Lee, Y.; Hong, S.; Shin, H.G.; Bae, H.; Yi, Y.; Kim, K.; Im, S. 2D TMD Channel Transistors with ZnO Nanowire Gate for Extended Nonvolatile Memory Applications. Adv. Funct. Mater. 2020, 30, 2004140. [Google Scholar] [CrossRef]
  35. Zhu, C.; Mu, X.; van Aken, P.A.; Yu, Y.; Maier, J. Single-layered ultrasmall nanoplates of MoS2 embedded in carbon nanofibers with excellent electrochemical performance for lithium and sodium storage. Angew. Chem. Int. Ed. Engl. 2014, 53, 2152–2156. [Google Scholar] [CrossRef]
  36. Sun, D.; Huang, D.; Wang, H.; Xu, G.-L.; Zhang, X.; Zhang, R.; Tang, Y.; Abd Ei-Hady, D.; Alshitari, W.; Saad Al-Bogami, A.; et al. 1T MoS2 nanosheets with extraordinary sodium storage properties via thermal-driven ion intercalation assisted exfoliation of bulky MoS2. Nano Energy 2019, 61, 361–369. [Google Scholar] [CrossRef]
  37. Mahankali, K.; Thangavel, N.K.; Gopchenko, D.; Arava, L.M.R. Atomically Engineered Transition Metal Dichalcogenides for Liquid Polysulfide Adsorption and Their Effective Conversion in Li-S Batteries. ACS Appl. Mater. Interfaces 2020, 12, 27112–27121. [Google Scholar] [CrossRef]
  38. Babu, G.; Masurkar, N.; Al Salem, H.; Arava, L.M. Transition Metal Dichalcogenide Atomic Layers for Lithium Polysulfides Electrocatalysis. J. Am. Chem. Soc. 2017, 139, 171–178. [Google Scholar] [CrossRef]
  39. Thangavel, N.K.; Gopalakrishnan, D.; Arava, L.M.R. Understanding Heterogeneous Electrocatalysis of Lithium Polysulfide Redox on Pt and WS2 Surfaces. J. Phys. Chem. C 2017, 121, 12718–12725. [Google Scholar] [CrossRef]
  40. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  41. Voshell, A.; Terrones, M.; Rana, M. Thermal and Photo Sensing Capabilities of Mono- and Few-Layer Thick Transition Metal Dichalcogenides. Micromachines 2020, 11, 693. [Google Scholar] [CrossRef]
  42. Masurkar, N.; Thangavel, N.K.; Yurgelevic, S.; Varma, S.; Auner, G.W.; Reddy Arava, L.M. Reliable and highly sensitive biosensor from suspended MoS2 atomic layer on nano-gap electrodes. Biosens. Bioelectron. 2021, 172, 112724. [Google Scholar] [CrossRef] [PubMed]
  43. Coleman, J.N.; Lotya, M.; O’Neill, A.; Bergin, S.D.; King, P.J.; Khan, U.; Young, K.; Gaucher, A.; De, S.; Smith, R.J.; et al. Two-dimensional nanosheets produced by liquid exfoliation of layered materials. Science 2011, 331, 568–571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Jayabal, S.; Saranya, G.; Wu, J.; Liu, Y.; Geng, D.; Meng, X. Understanding the high-electrocatalytic performance of two-dimensional MoS2 nanosheets and their composite materials. J. Mater. Chem. A 2017, 5, 24540–24563. [Google Scholar] [CrossRef]
  45. Toh, R.J.; Sofer, Z.; Luxa, J.; Sedmidubsky, D.; Pumera, M. 3R phase of MoS2 and WS2 outperforms the corresponding 2H phase for hydrogen evolution. Chem. Commun. 2017, 53, 3054–3057. [Google Scholar] [CrossRef] [Green Version]
  46. Huang, H.; Feng, X.; Du, C.; Song, W. High-quality phosphorus-doped MoS2 ultrathin nanosheets with amenable ORR catalytic activity. Chem. Commun. 2015, 51, 7903–7906. [Google Scholar] [CrossRef]
  47. Song, I.; Park, C.; Choi, H.C. Synthesis and properties of molybdenum disulphide: From bulk to atomic layers. RSC Adv. 2015, 5, 7495–7514. [Google Scholar] [CrossRef] [Green Version]
  48. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  49. Tao, H.; Gao, Y.; Talreja, N.; Guo, F.; Texter, J.; Yan, C.; Sun, Z. Two-dimensional nanosheets for electrocatalysis in energy generation and conversion. J. Mater. Chem. A 2017, 5, 7257–7284. [Google Scholar] [CrossRef]
  50. Tsai, C.; Chan, K.; Nørskov, J.K.; Abild-Pedersen, F. Rational design of MoS2 catalysts: Tuning the structure and activity via transition metal doping. Catal. Sci. Technol. 2015, 5, 246–253. [Google Scholar] [CrossRef]
  51. Huang, H.; Feng, X.; Du, C.; Wu, S.; Song, W. Incorporated oxygen in MoS2 ultrathin nanosheets for efficient ORR catalysis. J. Mater. Chem. A 2015, 3, 16050–16056. [Google Scholar] [CrossRef]
  52. Chen, D.; Liu, T.; Wang, P.; Zhao, J.; Zhang, C.; Cheng, R.; Li, W.; Ji, P.; Pu, Z.; Mu, S. Ionothermal Route to Phase-Pure RuB2 Catalysts for Efficient Oxygen Evolution and Water Splitting in Acidic Media. ACS Energy Lett. 2020, 5, 2909–2915. [Google Scholar] [CrossRef]
  53. Tang, L.; Meng, X.; Deng, D.; Bao, X. Confinement Catalysis with 2D Materials for Energy Conversion. Adv. Mater. 2019, 31, e1901996. [Google Scholar] [CrossRef] [PubMed]
  54. Nicolosi, V.; Chhowalla, M.; Kanatzidis, M.G.; Strano, M.S.; Coleman, J.N. Liquid Exfoliation of Layered Materials. Science 2013, 340, 568–571. [Google Scholar] [CrossRef] [Green Version]
  55. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef] [PubMed]
  56. Ambrosi, A.; Sofer, Z.; Pumera, M. Lithium intercalation compound dramatically influences the electrochemical properties of exfoliated MoS2. Small 2015, 11, 605–612. [Google Scholar] [CrossRef]
  57. Shih, C.J.; Vijayaraghavan, A.; Krishnan, R.; Sharma, R.; Han, J.H.; Ham, M.H.; Jin, Z.; Lin, S.; Paulus, G.L.; Reuel, N.F.; et al. Bi- and trilayer graphene solutions. Nat. Nanotechnol. 2011, 6, 439–445. [Google Scholar] [CrossRef]
  58. Eda, G.; Fujita, T.; Yamaguchi, H.; Voiry, D.; Chen, M.; Chhowalla, M. Coherent atomic and electronic heterostructures of single-layer MoS2. ACS Nano 2012, 6, 7311–7317. [Google Scholar] [CrossRef]
  59. Jimenez Sandoval, S.; Yang, D.; Frindt, R.F.; Irwin, J.C. Raman study and lattice dynamics of single molecular layers of MoS2. Phys. Rev. B Condens. Matter. 1991, 44, 3955–3962. [Google Scholar] [CrossRef] [PubMed]
  60. Cunningham, G.; Lotya, M.; Cucinotta, C.S.; Sanvito, S.; Bergin, S.D.; Menzel, R.; Shaffer, M.S.; Coleman, J.N. Solvent exfoliation of transition metal dichalcogenides: Dispersibility of exfoliated nanosheets varies only weakly between compounds. ACS Nano 2012, 6, 3468–3480. [Google Scholar] [CrossRef] [PubMed]
  61. O’Neill, A.; Khan, U.; Coleman, J.N. Preparation of High Concentration Dispersions of Exfoliated MoS2 with Increased Flake Size. Chem. Mater. 2012, 24, 2414–2421. [Google Scholar] [CrossRef]
  62. Lotya, M.; King, P.J.; Khan, U.; De, S.; Coleman, J.N. High-concentration, surfactant-stabilized graphene dispersions. ACS Nano 2010, 4, 3155–3162. [Google Scholar] [CrossRef]
  63. Smith, R.J.; King, P.J.; Lotya, M.; Wirtz, C.; Khan, U.; De, S.; O’Neill, A.; Duesberg, G.S.; Grunlan, J.C.; Moriarty, G.; et al. Large-scale exfoliation of inorganic layered compounds in aqueous surfactant solutions. Adv. Mater. 2011, 23, 3944–3948. [Google Scholar] [CrossRef]
  64. May, P.; Khan, U.; Hughes, J.M.; Coleman, J.N. Role of Solubility Parameters in Understanding the Steric Stabilization of Exfoliated Two-Dimensional Nanosheets by Adsorbed Polymers. J. Phys. Chem. C 2012, 116, 11393–11400. [Google Scholar] [CrossRef]
  65. Gao, M.R.; Liang, J.X.; Zheng, Y.R.; Xu, Y.F.; Jiang, J.; Gao, Q.; Li, J.; Yu, S.H. An efficient molybdenum disulfide/cobalt diselenide hybrid catalyst for electrochemical hydrogen generation. Nat. Commun. 2015, 6, 5982. [Google Scholar] [CrossRef] [PubMed]
  66. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 nanoparticles grown on graphene: An advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef] [Green Version]
  67. Li, Z.; Chen, Y.; Li, J.-F.; Chen, H.; Wang, L.; Zheng, S.; Lu, G. Systhesizing SnTe nanocrystals leading to thermoelectric performance enhancement via an ultra-fast microwave hydrothermal method. Nano Energy 2016, 28, 78–86. [Google Scholar] [CrossRef]
  68. López-Linares, J.C.; García-Cubero, M.T.; Lucas, S.; González-Benito, G.; Coca, M. Microwave assisted hydrothermal as greener pretreatment of brewer’s spent grains for biobutanol production. Chem. Eng. J. 2019, 368, 1045–1055. [Google Scholar] [CrossRef]
  69. Ding, W.; Hu, L.; Dai, J.; Tang, X.; Wei, R.; Sheng, Z.; Liang, C.; Shao, D.; Song, W.; Liu, Q.; et al. Highly Ambient-STable 1T-MoS2 and 1T-WS2 by Hydrothermal Synthesis under High Magnetic Fields. ACS Nano 2019, 13, 1694–1702. [Google Scholar] [CrossRef]
  70. Hao, C.; Wen, F.; Xiang, J.; Wang, L.; Hou, H.; Su, Z.; Hu, W.; Liu, Z. Controlled Incorporation of Ni(OH)2 Nanoplates Into Flowerlike MoS2 Nanosheets for Flexible All-Solid-State Supercapacitors. Adv. Funct. Mater. 2014, 24, 6700–6707. [Google Scholar] [CrossRef]
  71. Lee, Y.H.; Zhang, X.Q.; Zhang, W.; Chang, M.T.; Lin, C.T.; Chang, K.D.; Yu, Y.C.; Wang, J.T.; Chang, C.S.; Li, L.J.; et al. Synthesis of large-area MoS2 atomic layers with chemical vapor deposition. Adv. Mater. 2012, 24, 2320–2325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Balendhran, S.; Ou, J.Z.; Bhaskaran, M.; Sriram, S.; Ippolito, S.; Vasic, Z.; Kats, E.; Bhargava, S.; Zhuiykov, S.; Kalantar-Zadeh, K. Atomically thin layers of MoS2 via a two step thermal evaporation-exfoliation method. Nanoscale 2012, 4, 461–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Liu, H.F.; Wong, S.L.; Chi, D.Z. CVD Growth of MoS2-based Two-dimensional Materials. Chem. Vap. Depos. 2015, 21, 241–259. [Google Scholar] [CrossRef]
  74. Vilá, R.A.; Rao, R.; Muratore, C.; Bianco, E.; Robinson, J.A.; Maruyama, B.; Glavin, N.R. In situ crystallization kinetics of two-dimensional MoS2. 2D Mater. 2017, 5. [Google Scholar] [CrossRef]
  75. Helveg, S.; Lauritsen, J.V.; Laegsgaard, E.; Stensgaard, I.I.; Norskov, J.K.; Clausen, B.S.; Topsoe, H.; Besenbacher, F. Atomic-scale structure of single-layer MoS2 nanoclusters. Phys. Rev. Lett. 2000, 84, 951–954. [Google Scholar] [CrossRef] [Green Version]
  76. Tan, Y.; Liu, P.; Chen, L.; Cong, W.; Ito, Y.; Han, J.; Guo, X.; Tang, Z.; Fujita, T.; Hirata, A.; et al. Monolayer MoS2 films supported by 3D nanoporous metals for high-efficiency electrocatalytic hydrogen production. Adv. Mater. 2014, 26, 8023–8028. [Google Scholar] [CrossRef]
  77. Li, X.; Zhu, H. Two-dimensional MoS2: Properties, preparation, and applications. J. Mater. 2015, 1, 33–44. [Google Scholar] [CrossRef] [Green Version]
  78. Yu, Y.; Huang, S.Y.; Li, Y.; Steinmann, S.N.; Yang, W.; Cao, L. Layer-dependent electrocatalysis of MoS2 for hydrogen evolution. Nano Lett. 2014, 14, 553–558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Chen, W.; Huang, J.; Wei, J.; Zhou, D.; Cai, J.; He, Z.-D.; Chen, Y.-X. Origins of high onset overpotential of oxygen reduction reaction at Pt-based electrocatalysts: A mini review. Electrochem. Commun. 2018, 96, 71–76. [Google Scholar] [CrossRef]
  80. Katsounaros, I.; Cherevko, S.; Zeradjanin, A.R.; Mayrhofer, K.J. Oxygen electrochemistry as a cornerstone for sustainable energy conversion. Angew. Chem. Int. Ed. Engl. 2014, 53, 102–121. [Google Scholar] [CrossRef]
  81. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  82. Fu, Q.; Bao, X. Surface chemistry and catalysis confined under two-dimensional materials. Chem. Soc. Rev. 2017, 46, 1842–1874. [Google Scholar] [CrossRef]
  83. Lauritsen, J.V.; Kibsgaard, J.; Helveg, S.; Topsoe, H.; Clausen, B.S.; Laegsgaard, E.; Besenbacher, F. Size-dependent structure of MoS2 nanocrystals. Nat. Nanotechnol. 2007, 2, 53–58. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, T.; Gao, D.; Zhuo, J.; Zhu, Z.; Papakonstantinou, P.; Li, Y.; Li, M. Size-Dependent Enhancement of Electrocatalytic Oxygen-Reduction and Hydrogen-Evolution Performance of MoS2 Particles. Chemistry 2013, 19, 11939–11948. [Google Scholar] [CrossRef]
  85. Xie, J.; Zhang, J.; Li, S.; Grote, F.; Zhang, X.; Zhang, H.; Wang, R.; Lei, Y.; Pan, B.; Xie, Y. Controllable disorder engineering in oxygen-incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution. J. Am. Chem. Soc. 2013, 135, 17881–17888. [Google Scholar] [CrossRef] [PubMed]
  86. Li, L.; Wei, Z.; Chen, S.; Qi, X.; Ding, W.; Xia, M.; Li, R.; Xiong, K.; Deng, Z.; Gao, Y. A comparative DFT study of the catalytic activity of MnO2 (211) and (2-2-1) surfaces for an oxygen reduction reaction. Chem. Phys. Lett. 2012, 539–540, 89–93. [Google Scholar] [CrossRef]
  87. Zhang, H.; Tian, Y.; Zhao, J.; Cai, Q.; Chen, Z. Small Dopants Make Big Differences: Enhanced Electrocatalytic Performance of MoS2 Monolayer for Oxygen Reduction Reaction (ORR) by N– and P–Doping. Electrochim. Acta 2017, 225, 543–550. [Google Scholar] [CrossRef] [Green Version]
  88. Liu, C.; Dong, H.; Ji, Y.; Hou, T.; Li, Y. Origin of the catalytic activity of phosphorus doped MoS2 for oxygen reduction reaction (ORR) in alkaline solution: A theoretical study. Sci. Rep. 2018, 8, 13292. [Google Scholar] [CrossRef]
  89. Xiao, B.B.; Zhang, P.; Han, L.P.; Wen, Z. Functional MoS2 by the Co/Ni doping as the catalyst for oxygen reduction reaction. Appl. Surf. Sci. 2015, 354, 221–228. [Google Scholar] [CrossRef]
  90. Urbanova, V.; Lazar, P.; Antonatos, N.; Sofer, Z.; Otyepka, M.; Pumera, M. Positive and Negative Effects of Dopants toward Electrocatalytic Activity of MoS2 and WS2: Experiments and Theory. ACS Appl. Mater. Interfaces 2020, 12, 20383–20392. [Google Scholar] [CrossRef]
  91. He, T.; Xu, L.; Zhang, Y.; Huang, H.; Jiao, H. High-quality vanadium-doped MoS2 ultrathin nanosheets as an efficient ORR catalyst. New J. Chem. 2019, 43, 1611–1616. [Google Scholar] [CrossRef]
  92. Wang, Z.; Zhao, J.; Cai, Q.; Li, F. Computational screening for high-activity MoS2 monolayer-based catalysts for the oxygen reduction reaction via substitutional doping with transition metal. J. Mater. Chem. A 2017, 5, 9842–9851. [Google Scholar] [CrossRef]
  93. Laursen, A.B.; Kegnæs, S.; Dahl, S.; Chorkendorff, I. Molybdenum sulfides—Efficient and viable materials for electro—And photoelectrocatalytic hydrogen evolution. Energy Environ. Sci. 2012, 5. [Google Scholar] [CrossRef]
  94. Kibsgaard, J.; Chen, Z.; Reinecke, B.N.; Jaramillo, T.F. Engineering the surface structure of MoS2 to preferentially expose active edge sites for electrocatalysis. Nat. Mater. 2012, 11, 963–969. [Google Scholar] [CrossRef] [PubMed]
  95. Zhao, S.; Wang, K.; Zou, X.; Gan, L.; Du, H.; Xu, C.; Kang, F.; Duan, W.; Li, J. Group VB transition metal dichalcogenides for oxygen reduction reaction and strain-enhanced activity governed by p-orbital electrons of chalcogen. Nano Res. 2019, 12, 925–930. [Google Scholar] [CrossRef]
  96. Li, H.; Contryman, A.W.; Qian, X.; Ardakani, S.M.; Gong, Y.; Wang, X.; Weisse, J.M.; Lee, C.H.; Zhao, J.; Ajayan, P.M.; et al. Optoelectronic crystal of artificial atoms in strain-textured molybdenum disulphide. Nat. Commun. 2015, 6, 7381. [Google Scholar] [CrossRef] [PubMed]
  97. Enyashin, A.N.; Yadgarov, L.; Houben, L.; Popov, I.; Weidenbach, M.; Tenne, R.; Bar-Sadan, M.; Seifert, G. New Route for Stabilization of 1T-WS2 and MoS2 Phases. J. Phys. Chem. C 2011, 115, 24586–24591. [Google Scholar] [CrossRef] [Green Version]
  98. Yadgarov, L.; Rosentsveig, R.; Leitus, G.; Albu-Yaron, A.; Moshkovich, A.; Perfilyev, V.; Vasic, R.; Frenkel, A.I.; Enyashin, A.N.; Seifert, G.; et al. Controlled doping of MS2 (M = W, Mo) nanotubes and fullerene-like nanoparticles. Angew. Chem. Int. Ed. Engl. 2012, 51, 1148–1151. [Google Scholar] [CrossRef]
  99. Lin, Y.C.; Dumcenco, D.O.; Huang, Y.S.; Suenaga, K. Atomic mechanism of the semiconducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotechnol. 2014, 9, 391–396. [Google Scholar] [CrossRef]
  100. Lin, Y.C.; Dumcenco, D.O.; Komsa, H.P.; Niimi, Y.; Krasheninnikov, A.V.; Huang, Y.S.; Suenaga, K. Properties of individual dopant atoms in single-layer MoS2: Atomic structure, migration, and enhanced reactivity. Adv. Mater. 2014, 26, 2857–2861. [Google Scholar] [CrossRef] [Green Version]
  101. Sadighi, Z.; Liu, J.; Zhao, L.; Ciucci, F.; Kim, J.K. Metallic MoS2 nanosheets: Multifunctional electrocatalyst for the ORR, OER and Li-O2 batteries. Nanoscale 2018, 10, 22549–22559. [Google Scholar] [CrossRef]
  102. Lee, C.; Ozden, S.; Tewari, C.S.; Park, O.K.; Vajtai, R.; Chatterjee, K.; Ajayan, P.M. MoS2 -Carbon Nanotube Porous 3D Network for Enhanced Oxygen Reduction Reaction. ChemSusChem 2018, 11, 2960–2966. [Google Scholar] [CrossRef]
  103. Chang, K.; Mei, Z.; Wang, T.; Kang, Q.; Ouyang, S.; Ye, J. MoS2/graphene cocatalyst for efficient photocatalytic H2 evolution under visible light irradiation. ACS Nano 2014, 8, 7078–7087. [Google Scholar] [CrossRef] [PubMed]
  104. Yuan, K.; Zhuang, X.; Fu, H.; Brunklaus, G.; Forster, M.; Chen, Y.; Feng, X.; Scherf, U. Two-Dimensional Core-Shelled Porous Hybrids as Highly Efficient Catalysts for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. Engl. 2016, 55, 6858–6863. [Google Scholar] [CrossRef] [Green Version]
  105. Zhao, C.; Zhang, Y.; Chen, L.; Yan, C.; Zhang, P.; Ang, J.M.; Lu, X. Self-Assembly-Assisted Facile Synthesis of MoS2-Based Hybrid Tubular Nanostructures for Efficient Bifunctional Electrocatalysis. ACS Appl. Mater. Interfaces 2018, 10, 23731–23739. [Google Scholar] [CrossRef] [PubMed]
  106. Zhou, J.; Xiao, H.; Zhou, B.; Huang, F.; Zhou, S.; Xiao, W.; Wang, D. Hierarchical MoS2–rGO nanosheets with high MoS2 loading with enhanced electro-catalytic performance. Appl. Surf. Sci. 2015, 358, 152–158. [Google Scholar] [CrossRef]
  107. Mao, J.; Liu, P.; Du, C.; Liang, D.; Yan, J.; Song, W. Tailoring 2D MoS2 heterointerfaces for promising oxygen reduction reaction electrocatalysis. J. Mater. Chem. A 2019, 7, 8785–8789. [Google Scholar] [CrossRef]
  108. Kwon, I.S.; Kwak, I.H.; Kim, J.Y.; Abbas, H.G.; Debela, T.T.; Seo, J.; Cho, M.K.; Ahn, J.P.; Park, J.; Kang, H.S. Two-dimensional MoS2/Fe-phthalocyanine hybrid nanostructures as excellent electrocatalysts for hydrogen evolution and oxygen reduction reactions. Nanoscale 2019, 11, 14266–14275. [Google Scholar] [CrossRef]
  109. Amiinu, I.S.; Pu, Z.; Liu, X.; Owusu, K.A.; Monestel, H.G.R.; Boakye, F.O.; Zhang, H.; Mu, S. Multifunctional Mo-N/C@MoS2 Electrocatalysts for HER, OER, ORR, and Zn-Air Batteries. Adv. Funct. Mater. 2017, 27. [Google Scholar] [CrossRef]
  110. Prabhakar Vattikuti, S.V.; Nagajyothi, P.C.; Devarayapalli, K.C.; Yoo, K.; Dang Nam, N.; Shim, J. Hybrid Ag/MoS2 nanosheets for efficient electrocatalytic oxygen reduction. Appl. Surf. Sci. 2020, 526. [Google Scholar] [CrossRef]
  111. Li, J.; Li, G.; Wang, J.; Xue, C.; Li, X.; Wang, S.; Han, B.; Yang, M.; Li, L. A novel core–double shell heterostructure derived from a metal–organic framework for efficient HER, OER and ORR electrocatalysis. Inorg. Chem. Front. 2020, 7, 191–197. [Google Scholar] [CrossRef]
Figure 1. (A) Metal coordination, top view, and side view of single-layer 2H-MoS2 and 1T-MoS2. Reproduced with permission from [44] Copyright © 2021 The Royal Society of Chemistry. (B) Energy dispersion (energy versus wavevector k) in (1) bulk, (2) quadrilayer, (3) bilayer, and (4) monolayer MoS2 from left to right. Reproduced with permission from [48] Copyright © 2021 The American Chemical Society.
Figure 1. (A) Metal coordination, top view, and side view of single-layer 2H-MoS2 and 1T-MoS2. Reproduced with permission from [44] Copyright © 2021 The Royal Society of Chemistry. (B) Energy dispersion (energy versus wavevector k) in (1) bulk, (2) quadrilayer, (3) bilayer, and (4) monolayer MoS2 from left to right. Reproduced with permission from [48] Copyright © 2021 The American Chemical Society.
Micromachines 12 00240 g001
Figure 2. Schematic description of the main liquid exfoliation mechanisms.
Figure 2. Schematic description of the main liquid exfoliation mechanisms.
Micromachines 12 00240 g002
Figure 3. (A) Transmission electron microscope (TEM), high-resolution TEM (HRTEM) and high angle annular dark field scanning transmission electron microscope (HAADF STEM) image of MoS2 nanosheet exfoliation in sodium cholate solution. Reproduced with permission from [63] Copyright © 2021 WILEY-VCH. (B) TEM images of nanosheets dispersed in polymer/solvent solutions. Each image is labeled using the following convention: nanosheet/solvent/polymer. Reproduced with permission from [64] Copyright © 2021 American Chemical Society. (C) Schematic diagram of polymer stabilization.
Figure 3. (A) Transmission electron microscope (TEM), high-resolution TEM (HRTEM) and high angle annular dark field scanning transmission electron microscope (HAADF STEM) image of MoS2 nanosheet exfoliation in sodium cholate solution. Reproduced with permission from [63] Copyright © 2021 WILEY-VCH. (B) TEM images of nanosheets dispersed in polymer/solvent solutions. Each image is labeled using the following convention: nanosheet/solvent/polymer. Reproduced with permission from [64] Copyright © 2021 American Chemical Society. (C) Schematic diagram of polymer stabilization.
Micromachines 12 00240 g003
Figure 4. (A) Schematic illustration for the formation of self-supporting three-dimensional hierarchical nanostructure MoS2@Ni(OH)2 nanocomposites. Reproduced with permission from [70] Copyright © 2021 WILEY-VCH. (B) Proposed synthetic protocol for ultrathin P-MoS2 nanosheets. Reproduced with permission from [46] Copyright © 2021 The Royal Society of Chemistry.
Figure 4. (A) Schematic illustration for the formation of self-supporting three-dimensional hierarchical nanostructure MoS2@Ni(OH)2 nanocomposites. Reproduced with permission from [70] Copyright © 2021 WILEY-VCH. (B) Proposed synthetic protocol for ultrathin P-MoS2 nanosheets. Reproduced with permission from [46] Copyright © 2021 The Royal Society of Chemistry.
Micromachines 12 00240 g004
Figure 5. (A) (1) Schematic illustration of the experimental setup. (2) Optical microscopy images of MoS2 layers grown on a substrate respectively treated with reduced graphene oxide solution. The inset shows the magnified optical microscopy of the MoS2 films, where the seed is observed at the center of each star-shaped sheet. (3) Atomic Force Microscope (AFM) image of a monolayer MoS2 film on a SiO2/Si substrate (pre-treated with reduced graphene oxide (rGO)). (4) The thickness of the MoS2 layer is 0.72 nm from the AFM cross-sectional profile along the line indicated. Reproduced with permission from [71] Copyright © 2021 WILEY-VCH. (B) Monolayer MoS2@nanoporous gold (NPG) toward catalytic hydrogen evolution reaction (HER). (1) Schematic diagram of the fabrication process of monolayer MoS2@nanoporous gold (NPG) hybrid materials by a nanoporous metal-based CVD approach. (2) Schematic HER catalyzed by the monolayer MoS2@NPG hybrid material. Reproduced with permission from [76] Copyright © 2021 WILEY-VCH.
Figure 5. (A) (1) Schematic illustration of the experimental setup. (2) Optical microscopy images of MoS2 layers grown on a substrate respectively treated with reduced graphene oxide solution. The inset shows the magnified optical microscopy of the MoS2 films, where the seed is observed at the center of each star-shaped sheet. (3) Atomic Force Microscope (AFM) image of a monolayer MoS2 film on a SiO2/Si substrate (pre-treated with reduced graphene oxide (rGO)). (4) The thickness of the MoS2 layer is 0.72 nm from the AFM cross-sectional profile along the line indicated. Reproduced with permission from [71] Copyright © 2021 WILEY-VCH. (B) Monolayer MoS2@nanoporous gold (NPG) toward catalytic hydrogen evolution reaction (HER). (1) Schematic diagram of the fabrication process of monolayer MoS2@nanoporous gold (NPG) hybrid materials by a nanoporous metal-based CVD approach. (2) Schematic HER catalyzed by the monolayer MoS2@NPG hybrid material. Reproduced with permission from [76] Copyright © 2021 WILEY-VCH.
Micromachines 12 00240 g005
Figure 6. (A) Proposed oxygen reduction reaction (ORR) mechanisms. Reproduced with permission from [80] Copyright © 2021 WILEY-VCH. (B) Volcano-type relationship for the ORR activity versus the oxygen-binding energy. Reproduced with permission from [81] Copyright © 2021 The American Chemical Society.
Figure 6. (A) Proposed oxygen reduction reaction (ORR) mechanisms. Reproduced with permission from [80] Copyright © 2021 WILEY-VCH. (B) Volcano-type relationship for the ORR activity versus the oxygen-binding energy. Reproduced with permission from [81] Copyright © 2021 The American Chemical Society.
Micromachines 12 00240 g006
Figure 7. (A) Proposed synthetic protocol for incorporation of oxygen into ultrathin MoS2 nanosheets. (B) ORR performance test for O-MoS2 and pristine MoS2 nanosheets as well as 20% Pt/C in oxygen-saturated 0.1 M KOH. Reproduced with permission from [51] Copyright © 2021 The Royal Society of Chemistry. (C) Calculated density of states (DOS) of the oxygen-incorporated MoS2 slab (top) and the pristine 2H-MoS2 slab (bottom). The orange shading clearly indicates the decrease in bandgap after oxygen incorporation. (D) The charge density distributions of the valence band (left) and conduction band (right) near the oxygen atom in oxygen-incorporated MoS2 ultrathin nanosheets. Reproduced with permission from [85] Copyright © 2021 The American Chemical Society.
Figure 7. (A) Proposed synthetic protocol for incorporation of oxygen into ultrathin MoS2 nanosheets. (B) ORR performance test for O-MoS2 and pristine MoS2 nanosheets as well as 20% Pt/C in oxygen-saturated 0.1 M KOH. Reproduced with permission from [51] Copyright © 2021 The Royal Society of Chemistry. (C) Calculated density of states (DOS) of the oxygen-incorporated MoS2 slab (top) and the pristine 2H-MoS2 slab (bottom). The orange shading clearly indicates the decrease in bandgap after oxygen incorporation. (D) The charge density distributions of the valence band (left) and conduction band (right) near the oxygen atom in oxygen-incorporated MoS2 ultrathin nanosheets. Reproduced with permission from [85] Copyright © 2021 The American Chemical Society.
Micromachines 12 00240 g007
Figure 8. (A) X-ray diffraction (XRD) patterns of the MoO3/g-C3N4 intermediate and ultrathin P-MoS2 nanosheets and TEM and high resolution-TEM images of ultrathin P-MoS2 nanosheets. (B) ORR electrocatalysis for P-MoS2-0.1, P-MoS2-0.2, P-MoS2-0.3, MoS2, and 20% Pt/C catalysts. Reproduced with permission from [46] Copyright © 2021 The Royal Society of Chemistry. (C) Top view of the optimized structure of (a) pristine MoS2 nanosheets, (b) single P-doped MoS2 nanosheets, and (c) double P-doped MoS2 nanosheets. (D) Top and side views of the optimized adsorption configurations of (a) O2-adsorbed P-MoS2, (b) 2O-adsorbed P-MoS2, and (c) 2O-adsorbed 2P-MoS2. The Hirshfeld charge values of the corresponding atoms are also given. (E) Optimized structure configuration of reactant, intermediates, and product (14) and the corresponding reaction pathways of the ORR on a 2P-MoS2 sheet in an alkaline environment. Reproduced with permission from [88] Copyright © 2021 American Association for the Advancement of Science.
Figure 8. (A) X-ray diffraction (XRD) patterns of the MoO3/g-C3N4 intermediate and ultrathin P-MoS2 nanosheets and TEM and high resolution-TEM images of ultrathin P-MoS2 nanosheets. (B) ORR electrocatalysis for P-MoS2-0.1, P-MoS2-0.2, P-MoS2-0.3, MoS2, and 20% Pt/C catalysts. Reproduced with permission from [46] Copyright © 2021 The Royal Society of Chemistry. (C) Top view of the optimized structure of (a) pristine MoS2 nanosheets, (b) single P-doped MoS2 nanosheets, and (c) double P-doped MoS2 nanosheets. (D) Top and side views of the optimized adsorption configurations of (a) O2-adsorbed P-MoS2, (b) 2O-adsorbed P-MoS2, and (c) 2O-adsorbed 2P-MoS2. The Hirshfeld charge values of the corresponding atoms are also given. (E) Optimized structure configuration of reactant, intermediates, and product (14) and the corresponding reaction pathways of the ORR on a 2P-MoS2 sheet in an alkaline environment. Reproduced with permission from [88] Copyright © 2021 American Association for the Advancement of Science.
Micromachines 12 00240 g008
Figure 9. (A) Typical linear sweep voltammetry (LSV) curves toward ORRs with the corresponding values of onset potentials (B). Reproduced with permission from [90] Copyright © 2021 The American Chemical Society. (C) Schematic of a transition atom that embeds into the S vacancy of the MoS2 monolayer. Reproduced with permission from [92] Copyright © 2021 The Royal Society of Chemistry. (D) Schematic of Gibbs free energy diagrams of the ORR on various transition metal embedded MoS2 monolayers via the OOH association mechanism in an acidic medium. Based on literature data from [92].
Figure 9. (A) Typical linear sweep voltammetry (LSV) curves toward ORRs with the corresponding values of onset potentials (B). Reproduced with permission from [90] Copyright © 2021 The American Chemical Society. (C) Schematic of a transition atom that embeds into the S vacancy of the MoS2 monolayer. Reproduced with permission from [92] Copyright © 2021 The Royal Society of Chemistry. (D) Schematic of Gibbs free energy diagrams of the ORR on various transition metal embedded MoS2 monolayers via the OOH association mechanism in an acidic medium. Based on literature data from [92].
Micromachines 12 00240 g009
Figure 10. Assembly of an artificial-atom crystal via spatially patterned biaxial strain within a monolayer of MoS2. Reproduced with permission from [96] Copyright © 2021 Nature Publishing Group.
Figure 10. Assembly of an artificial-atom crystal via spatially patterned biaxial strain within a monolayer of MoS2. Reproduced with permission from [96] Copyright © 2021 Nature Publishing Group.
Micromachines 12 00240 g010
Figure 11. (A) Illustration of the synthetic procedure. (B) Electrocatalytic activities of 1T-MoS2, 1T-NaxMoS2, 1T-MoS2/carbon nanotube (CNT), 2H-MoS2, 2H-MoS2/CNT, and Pt/C in O2-saturated 0.1 M KOH. Reproduced with permission from [101]. Copyright © 2021 The Royal Society of Chemistry.
Figure 11. (A) Illustration of the synthetic procedure. (B) Electrocatalytic activities of 1T-MoS2, 1T-NaxMoS2, 1T-MoS2/carbon nanotube (CNT), 2H-MoS2, 2H-MoS2/CNT, and Pt/C in O2-saturated 0.1 M KOH. Reproduced with permission from [101]. Copyright © 2021 The Royal Society of Chemistry.
Micromachines 12 00240 g011
Figure 12. (A) Idealized formula scheme depicting the chemical exfoliation of bulk MoS2 and subsequent functionalization with 4-iodophenyl substituents under the formation of MoS2-I as well as the preparation of MoS2-templated conjugated microporous polymers and the corresponding MoS2/nitrogen-doped porous carbon hybrids. (B) Electrochemical evaluation of catalysts in alkaline media of the MoS2-based porous carbon nanosheets. Reproduced with permission from [104] Copyright © 2021 WILEY-VCH. (C) (1), (2) Scanning electron microscope (SEM), (3), (4) TEM images of (NH4)2MoS4/PDA nanotubes; (5) TEM image and (6) HRTEM image of MoS2/ nitrogen-doped carbon hybrid nanotubes (C HNT). (D) Electrochemical evaluation of catalysts in alkaline media of the Pt/C, CPDA, bulk MoS2, and MoS2/C HNT. Reproduced with permission from [105] Copyright © 2021 American Chemical Society. (E) Field emission-SEM images of hierarchical MoS2− reduced graphene oxide (rGO) nanosheets. (F) LSV curves of MoS2–graphene oxide (GO) (solid curves) and MoS2 microparticles (dashed line) regarding the ORR in O2-saturated 0.1 M KOH at various rotating speeds. Reproduced with permission from [106] Copyright © 2021 Elsevier B.V.
Figure 12. (A) Idealized formula scheme depicting the chemical exfoliation of bulk MoS2 and subsequent functionalization with 4-iodophenyl substituents under the formation of MoS2-I as well as the preparation of MoS2-templated conjugated microporous polymers and the corresponding MoS2/nitrogen-doped porous carbon hybrids. (B) Electrochemical evaluation of catalysts in alkaline media of the MoS2-based porous carbon nanosheets. Reproduced with permission from [104] Copyright © 2021 WILEY-VCH. (C) (1), (2) Scanning electron microscope (SEM), (3), (4) TEM images of (NH4)2MoS4/PDA nanotubes; (5) TEM image and (6) HRTEM image of MoS2/ nitrogen-doped carbon hybrid nanotubes (C HNT). (D) Electrochemical evaluation of catalysts in alkaline media of the Pt/C, CPDA, bulk MoS2, and MoS2/C HNT. Reproduced with permission from [105] Copyright © 2021 American Chemical Society. (E) Field emission-SEM images of hierarchical MoS2− reduced graphene oxide (rGO) nanosheets. (F) LSV curves of MoS2–graphene oxide (GO) (solid curves) and MoS2 microparticles (dashed line) regarding the ORR in O2-saturated 0.1 M KOH at various rotating speeds. Reproduced with permission from [106] Copyright © 2021 Elsevier B.V.
Micromachines 12 00240 g012
Figure 13. (A) Proposed synthetic protocol for ultrathin Ni3S2/MoS2 nanosheets. (B) TEM image of Ni3S2/MoS2-0.2 nanosheets. (C) Electrochemical evaluation of catalysts in alkaline media. Reproduced with permission from [107] Copyright © 2021 The Royal Society of Chemistry. (D) HRTEM and HAAD-STEM images of the FePc-MoS2 hybrid complex. (E) Electrochemical evaluation of catalysts of FePc-MoS2 in alkaline media. Reproduced with permission from [108] Copyright © 2021 The Royal Society of Chemistry. (F) Representation of synthetic procedure of a Ag/MoS2 nanohybrid. Reproduced with permission from [110] Copyright © 2021 Elsevier B.V.
Figure 13. (A) Proposed synthetic protocol for ultrathin Ni3S2/MoS2 nanosheets. (B) TEM image of Ni3S2/MoS2-0.2 nanosheets. (C) Electrochemical evaluation of catalysts in alkaline media. Reproduced with permission from [107] Copyright © 2021 The Royal Society of Chemistry. (D) HRTEM and HAAD-STEM images of the FePc-MoS2 hybrid complex. (E) Electrochemical evaluation of catalysts of FePc-MoS2 in alkaline media. Reproduced with permission from [108] Copyright © 2021 The Royal Society of Chemistry. (F) Representation of synthetic procedure of a Ag/MoS2 nanohybrid. Reproduced with permission from [110] Copyright © 2021 Elsevier B.V.
Micromachines 12 00240 g013
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, H.; Zhu, J.; Ma, Q.; Ma, J.; Bai, H.; Chen, L.; Mu, S. Two-Dimensional MoS2: Structural Properties, Synthesis Methods, and Regulation Strategies toward Oxygen Reduction. Micromachines 2021, 12, 240. https://doi.org/10.3390/mi12030240

AMA Style

Xu H, Zhu J, Ma Q, Ma J, Bai H, Chen L, Mu S. Two-Dimensional MoS2: Structural Properties, Synthesis Methods, and Regulation Strategies toward Oxygen Reduction. Micromachines. 2021; 12(3):240. https://doi.org/10.3390/mi12030240

Chicago/Turabian Style

Xu, Hanwen, Jiawei Zhu, Qianli Ma, Jingjing Ma, Huawei Bai, Lei Chen, and Shichun Mu. 2021. "Two-Dimensional MoS2: Structural Properties, Synthesis Methods, and Regulation Strategies toward Oxygen Reduction" Micromachines 12, no. 3: 240. https://doi.org/10.3390/mi12030240

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